International Handbook of Mathematics Teacher Education: Volume 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 419

International Handbook of Mathematics Teacher Education: Volume 1

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
International Handbook of
Mathematics Teacher Education
(2nd Edition)
Series Editor:

Olive Chapman
University of Calgary
Calgary, Alberta
Canada

This second edition of the International Handbook of Mathematics Teacher Education


builds on and extends the first edition (2008) in addressing the knowledge, teaching
and learning of mathematics teachers at all levels of teaching mathematics and
of mathematics teacher educators, and the approaches/activities and programmes
through which their learning can be supported. It consists of four volumes based on
the same themes as the first edition.

Volume 1: Knowledge, Beliefs, and Identity in Mathematics Teaching and


Teaching Development
Despina Potari, National and Kapodistrian University of Athens, Athens, Greece and
Olive Chapman, University of Calgary, Calgary, Canada (eds.)
paperback: 978-90-04-41886-8, hardback: 978-90-04-41885-1,
ebook: 978-90-04-41887-5

Volume 2: Tools and Processes in Mathematics Teacher Education


Salvador Llinares, University of Alicante, Alicante, Spain and Olive Chapman,
University of Calgary, Calgary, Canada (eds.)
paperback: 978-90-04-41897-4, hardback: 978-90-04-41895-0,
ebook: 978-90-04-41896-7

Volume 3: Participants in Mathematics Teacher Education


Gwendolyn M. Lloyd, Pennsylvania State University, Pennsylvania, USA and
Olive Chapman, University of Calgary, Calgary, Canada (eds.)
paperback: 978-90-04-41922-3, hardback: 978-90-04-41921-6,
ebook: 978-90-04-41923-0

Volume 4: The Mathematics Teacher Educator as a Developing Professional


Kim Beswick, University of New South Wales, Sydney, Australia and Olive Chapman,
University of Calgary, Calgary, Canada (eds.)
paperback: 978-90-04-42420-3, hardback: 978-90-04-42419-7,
ebook: 978-90-04-42421-0

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
International Handbook
of Mathematics Teacher
Education: Volume 1

Knowledge, Beliefs, and Identity in Mathematics


Teaching and Teaching Development

(Second Edition)

Edited by

Despina Potari and Olive Chapman

අൾංൽൾඇ_ൻඈඌඍඈඇ

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
Cover illustration: Photograph by John Mason

All chapters in this book have undergone peer review.

The Library of Congress Cataloging-in-Publication Data is available online at


http://catalog.loc.gov

ISBN 978-90-04-41886-8 (paperback)


ISBN 978-90-04-41885-1 (hardback)
ISBN 978-90-04-41887-5 (e-book)

Copyright 2020 by Koninklijke Brill NV, Leiden, The Netherlands.


Koninklijke Brill NV incorporates the imprints Brill, Brill Hes & De Graaf,
Brill Nijhoff, Brill Rodopi, Brill Sense, Hotei Publishing, mentis Verlag,
Verlag Ferdinand Schöningh and Wilhelm Fink Verlag.
All rights reserved. No part of this publication may be reproduced, translated,
stored in a retrieval system, or transmitted in any form or by any means, electronic,
mechanical, photocopying, recording or otherwise, without prior written
permission from the publisher.
Authorization to photocopy items for internal or personal use is granted by
Koninklijke Brill NV provided that the appropriate fees are paid directly to The
Copyright Clearance Center, 222 Rosewood Drive, Suite 910, Danvers, MA 01923,
USA. Fees are subject to change.

This book is printed on acid-free paper and produced in a sustainable manner.

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CONTENTS

Preface vii

List of Figures and Tables ix

Mathematics Teaching and its Development: Looking into Teacher


Knowledge, Beliefs and Identity: An Introduction 1
Despina Potari

Part 1: Mathematics Teacher Knowledge and Its Relation to Teaching

1. Mathematical Subject Knowledge for Teaching Primary School


Mathematics: Evidence and Models for Professional Development 15
Mike Askew and Hamsa Venkat

2. Building Teachers’ Capacity in Mathematics Authentic Assessment 43


Kim Koh and Olive Chapman

3. Mathematics Conceptual Knowledge for Teaching: Helping Prospective


Teachers Know Mathematics Well Enough for Teaching 77
Yeping Li, JeongSuk Pang, Huirong Zhang and Naiqing Song

4. Researching Mathematical Knowledge in Teaching 105


Tim Rowland

Part 2: Mathematics Teacher Beliefs about Mathematics and Its Teaching

5. Mathematics Teachers’ Cultural Beliefs: The Case of Lesson Study 131


Maria G. Bartolini Bussi, Silvia Funghi and Alessandro Ramploud

6. Mathematical Creativity in the Classroom: Teachers’ Conceptions and


Professional Development 155
Esther S. Levenson

Part 3: The Interplay of Mathematics Teacher Identity, Beliefs and Knowledge

7. Beliefs and Pedagogical Content Knowledge for Teachers


of Mathematics 185
Kim Beswick and Helen Chick

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CONTENTS

8. Developing Professional Knowledge and Identities of Non-Specialist


Teachers of Mathematics 211
Merrilyn Goos, Anne Bennison, Stephen Quirke, Niamh O’Meara
and Colleen Vale

Part 4: Mathematics Teaching and Its Development

9. Mathematics Teachers Committed to Equity: A Review of


Teaching Practices 243
Marta Civil, Roberta Hunter and Sandra Crespo

10. Inquiry-Based Practice in University Mathematics Teaching


Development 275
Barbara Jaworski

11. Teacher Decision Making: Developments in Research and Theory 303


Despina Potari and Konstantinos Stouraitis

Part 5: From Mathematics Teaching Practices to Teacher Education

12. Exemplifying with Variation and Its Development in Mathematics


Teacher Education 329
Jill Adler and Craig Pournara

13. Mathematics Teaching Practices and Practice-Based Pedagogies:


A Critical Review of the Literature Since 2000 355
Charalambos Y. Charalambous and Seán Delaney

14. Seeing Mathematics through the Lens of Children’s Mathematical


Thinking: Perspective on the Development of Mathematical
Knowledge for Teaching 391
Randolph A. Philipp, John (Zig) Siegfried and Eva Thanheiser

Index 419

vi

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PREFACE

It is an honor to follow Terry Wood, series editor of the first edition of the
International Handbook of Mathematics Teacher Education, as series editor of this
second edition of the Handbook. As Terry indicated, she, Barbara Jaworski, Sandy
Dawson and Thomas Cooney played key roles in opening up the field of mathematics
teacher education “to establish mathematics teacher education as an important and
legitimate area of research and scholarship” (Wood, 2008, p.vii). The field has grown
significantly since the late 1980s “when Barbara Jaworski initiated and maintained
the first Working Group on mathematics teacher education at PME” (p. vii) and over
the last 10 years following the first edition of the Handbook. So, the editorial team,
I and the four volume editors (Kim Beswick, Salvador Llinares, Gwendolyn Lloyd,
and Despina Potari), of this second edition is honored to present it to the mathematics
education community and to the field of teacher education in general. It builds on
and extends the topics/ideas in the first edition while maintaining the themes for each
of the volumes. Collectively, the authors looked back beyond and within the last 10
years to establish the state-of-the-art and continuing and new trends in mathematics
teacher and mathematics teacher educator education, and looked forward regarding
possible avenues for teachers, teacher educators, researchers, and policy makers to
consider to enhance and/or further investigate teacher and teacher educator learning
and practice, in particular. The volume editors provide introductions to each volume
that highlight the subthemes used to group related chapters and provide meaningful
lenses to see important connections within and across chapters. Readers can also
use these subthemes to make connections across the four volumes, which, although
presented separately, include topics that have relevance across them since they are
all situated in the common focus regarding mathematics teachers.
I extend special thanks to the volume editors for their leadership and support
in preparing this handbook. I feel very fortunate to have had the opportunity to
work with them on this project. Also, on behalf of myself and the volume editors,
sincere thanks to all of the authors for their invaluable contributions and support in
working with us to produce a high-quality handbook to inform and move the field of
mathematics teacher education forward.

Volume 1, Knowledge, Beliefs, and Identity in Mathematics Teaching and


Teaching Development, edited by Despina Potari, examines teacher knowledge,
beliefs, identity, practice and relationships among them. These important aspects
of mathematics teacher education continue to be the focus of extensive research
and policy debate globally. Thus, as the first volume in the series, it appropriately
addresses central topics/issues that provide an excellent beginning to engage in the
field of mathematics education through the handbook.

vii

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PREFACE

REFERENCE
Sullivan, P. & Wood, T. (Eds.). (2008). International handbook of mathematics teacher
education: Vol. 1. Knowledge and beliefs in mathematics teaching and teaching
development. Rotterdam, the Netherlands: Sense Publishers.

Olive Chapman
Calgary, AB
Canada

viii

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
FIGURES AND TABLES

FIGURES

1.1. Diagrams with open array for equation 32


2.1. Adapted mathematical tasks framework 71
3.1. Three possible models for ÷ 89
3.2. A model that differs from (A) in Figure 3.1 and connects with
the complex fraction division algorithm 90
5.1. The complexity of relationships in mathematics education (adapted
from Xie & Carspecken, 2008, p. 17) 137
7.1. The division of fractions item 196
8.1. Ball’s model of Mathematical Knowledge for Teaching
(Ball et al., 2008, p. 403) 214
10.1. Inquiry in three layers 278
10.2. Inquiry in three layers at university level 279
11.1. The activity system (adapted from Engeström, 2001a) 313
12.1. Constitutive elements of Mathematical Discourse in Instruction
(adapted from Adler & Ronda, 2015) 333
12.2. Mathematics tasks for quadratic equations and trigonometric notation 340
12.3. Learner example set involving application of the distributive law 342
12.4. Teacher example set for application of the distributive law 342
12.5. Teachers’ extensions of the given example set 343
12.6. Example set for learner task to match equations and graphs 344
12.7. Teacher task for function matching task 345
12.8. Learner task for operations on integers 346
12.9. Successive adaptations of the object of learning and example sets by
teachers 347
13.1. Flowchart of the literature search and annotation 357
13.2. How techniques are nested within practices of different grain
sizes which are nested within the domain of “Assessing Student
Thinking” (based on Sleep & Boerst, 2012) 363
14.1. Prospective elementary school teachers looking at school
mathematical issues 392
14.2. Prospective elementary school teachers looking at school
mathematical issues through the lens of children’s mathematical
thinking see richer and different mathematics 393
14.3. An approach to the topic of fraction division as an integration of
concepts, procedures, reasoning, and problem solving 395
14.4. A solely procedural approach to the topic of fraction division 395

ix

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
FIGURES AND TABLES

14.5. Andrew’s work in solving 63 – 25 403


14.6. Recomposing the minuend or subtrahend and compensating for the
change to show Andrew’s reasoning 404
14.7. Circles of Caring. A model of growth, by way of children’s
mathematical thinking, from prospective elementary school
teachers’ caring about children to their caring about mathematics
(Philipp et al., 2007, p. 441) 410
14.8. Subtraction of 321 – 80 using an equal-addition algorithm 411
14.9. Many paths exist for developing mathematical knowledge for
teaching, represented in this figure by the paths ending with
arrows. Some paths provide higher views offering a more
comprehensive view of the landscape, and others do not.
The path that combines mathematics with children’s mathematical
thinking, represented by the solid path, is a particularly productive
path yielding a comprehensive view of the landscape 413

TABLES

2.1. An example of mathematics authentic assessment task


(adapted from Lim, 2011) 47
2.2. Definitions of the six mathematical competencies in the item
analysis scheme (Turner et al., 2015) [as cited in Pettersen &
Braeken, 2019, p. 408] 49
2.3. Cognitive demands or complexity of mathematical tasks 50
2.4. Authentic intellectual quality criteria for mathematics authentic
assessment tasks 52
2.5. An example of patchwork text assessment approach
(Koh et al., 2015) 63
2.6. PT-A’s grade 7 authentic mathematics task 65
3.1. Summary results of sampled elementary school teachers’
mathematics conceptual knowledge for teaching in fraction
division 99
4.1. The Knowledge Quartet (adapted from Rowland et al., 2005) 116
4.2. The Knowledge Quartet – Dimensions and contributory codes 120
5.1. Comparison between sections of the Chinese lesson plan and
the Italian lesson plan 143
6.1. Types of relationships between creativity and excellence 167
6.2. Inferred beliefs related to each category 168
7.1. Framework for Pedagogical Content Knowledge (PCK)
for teaching school mathematics 191

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
FIGURES AND TABLES

8.1. Elements of the numeracy model developed by Goos and colleagues 227
9.1. Characterization of exclusionary/inclusive mathematics teaching
practices (Louie, 2017, p. 496) 247
10.1. Teaching Triad analysis of teaching-learning in the ESUM project 289
10.2. Teaching Triad analysis of activity in the SYMBoL Project 291
10.3. Teaching Triad analysis of learning in the Catalyst Project 295

xi

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

MATHEMATICS TEACHING AND ITS


DEVELOPMENT: LOOKING INTO TEACHER
KNOWLEDGE, BELIEFS AND IDENTITY
An Introduction

This introduction to Volume 1 of this second edition of the International Handbook


of Mathematics Teacher Education starts with a short discussion of the nature of
mathematics teaching and its development and on the attempts that have been made
to search for its quality and the underlying reasons that promote this quality. In that
perspective, mathematics teacher knowledge, beliefs and identity become important
as a way to study mathematics teaching focusing on the mathematics teacher as
the main actor. Following this discussion, the fourteen chapters of the volume are
presented emphasising how they contribute to our understanding of the complexity
of mathematics learning, teaching and teacher education.

***

In Volume 1 of the first edition of this International Handbook of Mathematics


Teacher Education, Peter Sullivan (2008) described the content as presenting
research and theoretically informed perspectives on Knowledge and Beliefs
in Mathematics Teaching and Teaching Development. The chapters together
address the “what” of mathematics teacher education, meaning knowledge
for mathematics teaching and teaching development and consideration of
associated beliefs. (p. 1)
In this second edition of the handbook, Volume 1 extends this work, as discussed
at the end of this introduction chapter, to include teacher identity and recent
developments in research on these areas related to mathematics teacher education.
Mathematics teaching and its development has always been the focus of
mathematics education research. The study of mathematics has been initiated with
a focus on specific interventions and the development of materials and approaches
(e.g., Lunkenbein, 1977; Stacey, Helme, Acher, & Condon, 2001), moving to
the study of classroom interactions (Cobb, Wood, & Yackel, 1990) and to the
current description about teaching practices that aim to characterise the quality of

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_001

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

mathematics teaching and use a practice-based pedagogy to teacher education (see


Charalambous & Delaney, Chapter 13, this volume).
Numerous characterisations have been made about the nature of teaching that
imply different theoretical perspectives and ideologies of mathematics teaching.
For example, constructs like exploratory teaching or investigative teaching, used
mainly in the decades of the 1980s and 1990s, indicate a rather constructivist view
of teaching and learning where the focus is on the close relation between teaching
actions and students’ learning in the context of classroom interactions (Jaworski,
1994). The notion of equitable teaching that has become more central currently
implies a socio-cultural or socio-political perspective that focuses on the social,
cultural and political structures that allow, limit, or prevent the mathematical
engagement of all students in the classroom (Bartell et al., 2017). The dichotomy
between “traditional” versus “reform” teaching has been one of the most commonly
used characterisations of mathematics teaching, implying a specific type of teaching
that is desirable and facilitates mainly the development of students’ mathematical
thinking. To this direction, notions of “good teaching” (Wilson, Cooney & Stinson,
2005), “effective teaching” (Anthony & Walshaw, 2009), “inquiry teaching”
(see the chapter of Jaworski in this volume), and “equitable teaching” (see Civil,
Hunter, & Crespo, Chapter 9, this volume), have been used to describe the quality
of mathematics teaching mainly in relation to its impact on students’ mathematical
learning. Research also has showed that these notions are culturally driven, so what
is effective mathematics teaching for example in China is not the same as in Germany
(Cai, Kaiser, Perry, & Ngai-Ying Wong, 2009).
Research on mathematics teaching has been accompanied by research on
mathematics teacher professional development to support the shifts in teaching
perspectives. For example, there is some evidence that professional development
contexts in which teachers plan and reflect on mathematics teaching, usually
working with other teachers and researchers, have promoted the development
of teaching. This is reflected in some of the chapters included in this volume of
the handbook that report approaches to development of mathematics teaching
through the development of teachers’ professional learning. These approaches are
associated with, for example, a lesson study context as in Bartolini-Bussi, Funghi
and Ramploud (Chapter 5, this volume), practice-based pedagogies as discussed in
Charambous and Delaney (Chapter 13, this volume), and professional development
courses as in Adler and Pournara (Chapter 12, this volume), where teachers work
with exemplification both as learners of mathematics and as teachers.
The teacher has been the central focus of all efforts of mathematics education
researchers to study and improve the quality of mathematics teaching. In particular,
mathematics teacher beliefs and knowledge have been the focus of many studies
during the last three decades in an attempt to interpret why mathematics teaching is
effective or not and to look for ways of developing it. Several of these studies argued
that there is a close relationship between the teacher’s beliefs about mathematics,
teaching and learning and the teacher’s instructional practice; thus, changing teacher

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING AND ITS DEVELOPMENT

beliefs to more inquiry views of mathematics, teaching and learning would improve
the quality of mathematics teaching (e.g., Cross, 2009; Ponte & Chapman, 2006;
Thompson, 1984).
In order to address a relationship, similar to that of teacher belief, between teacher
knowledge and practice, different frameworks have been developed to describe
dimensions of mathematics teacher knowledge that seemed to be related to the
quality of mathematics teaching (e.g., the Ball and colleagues’ framework) (Ball,
Thames, & Phelps, 2008) and the Rowland’s (Chapter 2, this volume) or Li, Pang,
Zhang, and Song’s framework (Chapter 3, this volume). Some studies also view
teacher knowledge and beliefs as interrelated epistemologically (see Beswick &
Chick, Chapter 7, this volume) and others attempt to see the interplay of knowledge
and beliefs with mathematics teaching (Charalambous, 2015). Different attempts
have also been made in teacher education and professional development contexts to
improve teacher knowledge and beliefs (e.g., Miyakawa & Winslow, 2013; Koh &
Chapman, Chapter 2, this volume, in relation to the design of authentic assessment
mathematical tasks and Levenson (Chapter 6, this volume) in terms of mathematical
creativity).
Although knowledge and beliefs have been the focus of many studies as one
important way of looking into the classroom teacher and on how mathematics
teaching can improve, some recent studies have shown that there is not a cause-
effect relation between teachers’ knowledge, beliefs and mathematics teaching.
For example, Skott (2009) challenges the relationship between teachers’ beliefs
and the quality of mathematics teaching and more recently elaborates further on
the limitations of researching teachers’ beliefs to interpret teachers’ instructional
decisions and actions (Skott, 2015). Askew and Venkant in this volume also argue
that in our field we do not have strong research evidence that the development of the
different dimensions of teachers’ mathematical knowledge have a positive impact on
mathematics teaching and students’ learning. Thus, current research has recognized
that mathematics teaching is complex and cannot be studied and developed without
taking into account the social, cultural and institutional contexts from which the
teacher draws in his/her daily teaching.
Valuing the importance of context is more indicative in the research on teacher
identity as addressed in a number of reviews in the area (Graven & Heyd-Metzuyanim,
2019). For example, Ntow and Adler (2019) use the construct of mathematics teaching
identity to link professional development contexts to the choice of resources and
the enactment of teaching in the mathematics classroom. Overall, research has used
several frameworks of identity that indicate different interrelations between teacher
identity and mathematics teaching, with a focus on mathematics teaching through
teachers’ multiple identities in their participation in different practices (Losano,
Fiorentini, & Villarreal, 2018). Teacher identity is explicitly discussed and related to
teacher knowledge and mathematics teaching in Goos, Bennison, Quirke, O’Meara,
and Vale (Chapter 8, this volume).

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

These complex relations between teacher beliefs, knowledge, identity and


mathematics teaching are addressed in different ways in the fifteen chapters of
this volume of the handbook. The chapters offer to a large extent the state of the
art in these particular areas and they discuss and give insights about the nature of
mathematics teaching practices that may promote meaningful mathematics learning
for students with different needs. These practices are linked to the directions and
ways that the teacher needs to be supported to enact these practices. The volume is
structured in five sections related to the main focus of the fourteen chapters included
in it, as follows:
Part 1: Mathematics teacher knowledge and its relation to teaching.
Part 2: Mathematics teacher beliefs about mathematics and its teaching.
Part 3: The interplay of mathematics teacher identity, beliefs and knowledge.
Part 4: Mathematics teaching and its development.
Part 5: From mathematics teaching practices to teacher education.
Next is a brief overview of the main ideas discussed in the chapters in reference
to each of these sections followed by a conclusion of how they extend the work in
Volume 1 of the first edition of this handbook.

CONTRIBUTIONS IN THIS VOLUME

Mathematics Teacher Knowledge and Its Relation to Teaching

The chapter of Mike Askew and Hamsa Venkat offers a critical view about the nature
of mathematical subject knowledge needed for teaching mathematics at primary
school and its impact on mathematics teaching and learning. In their literature
review, they found different terms that have been used to characterize the quality of
teacher subject matter knowledge such as “conceptual understanding,” “specialized
content knowledge,” “profound understanding of fundamental mathematics,” and
“connected mathematical knowledge.” They point out that a number of research
studies search for relations between these forms of knowledge to students’
mathematical learning. However, they argue that in our research community, the
situation concerning the impact of teachers’ mathematical knowledge on students
learning is still complex and beyond a cause-effect relation. Contextual factors
need to be taken into account and existing conceptual frameworks in the area of
teachers’ subject matter knowledge do not appropriately address them. Going more
deeply on philosophical aspects of the nature of mathematical knowledge developed
by the mathematicians and unpacked by the classroom teacher, they challenge the
distinction between pedagogical content knowledge and subject knowledge and the
need to do this. Adopting a sociocultural perspective, they argue that mathematical
knowledge is not something that the teachers carry but it is blended in the practice
of teaching. In terms of teacher education and professional development, they claim
that it is important for the teachers to understand how mathematical knowledge is
established and develop a particular sensibility towards mathematics. The chapter

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING AND ITS DEVELOPMENT

helps us to realise the need of research on teacher subject knowledge to shift its
focus from the individual to the collective, placing the teacher in the position of the
classroom, in the school, and in general, in social interactions with other teachers
and resources.
Kim Koh and Olive Chapman, in their chapter, discuss different frameworks used
for analysing mathematical tasks and in particular authentic assessment tasks. They
reviewed research on mathematics teachers’ knowledge of mathematical tasks that
indicates mathematics teachers’ difficulties to recognise the different attributes of
the tasks and to enact them in the classroom. However, they report professional
development attempts where teachers used existing frameworks to analyse tasks
mainly in terms of their cognitive demands that show that they supported teachers
to overcome their difficulties. The main focus of this chapter is on authentic
assessment tasks aiming to assess students’ mathematical understanding and their
engagement in mathematical processes proposed by reform curricula. The authors
propose a framework of the principles of designing authentic assessment tasks
and provide two examples from their own studies with prospective and practicing
mathematics teachers working collaboratively in analysing and designing authentic
assessment tasks. These cases indicate teachers’ improvement of their capacity to
design authentic assessment tasks but also of their mathematical knowledge for
teaching. The chapter addresses an area where research is rather limited and helps us
understand the important role of designing and analysing instructional or assessment
tasks in the development of mathematical knowledge of teaching.
ȉKH FKDSWHU RI <HSLQJ /L -RHQJ6XN 3DQJ +XLURQJ =KDQJ DQG 1DLTLQJ 6RQJ
discusses the notion of “Mathematics conceptual knowledge for teaching,” that is, a
connected way of mathematics teacher knowledge based on the work of Liping Ma.
Comparing this construct to other existing ones in the research literature, they argue
that mathematics conceptual knowledge for teaching emphasizes the connections
between the subject matter knowledge and pedagogical content knowledge. Their
framework characterizes three aspects of teacher knowledge that are interrelated. In
the chapter, they present the results of two surveys given to elementary prospective
teachers in Mainland China and South Korea regarding their knowledge on the
division of fractions. The results from the surveys indicated the strong knowledge
of prospective teachers in both countries but in South Korea the knowledge related
to the teaching of division of fraction was stronger. The authors interpret these
differences on the basis of the teacher education programs existing in each country.
The chapter contributes to the development of our understanding of mathematics
teacher knowledge in relation to the teacher education policy and culture in a country.
The chapter of Tim Rowland discusses the theoretical framework on the
Knowledge Quartet, emphasizing the history of its development, methodological
issues that the author and his team encountered, and decisions taken in the design
of the study and in particular in the data analysis. This unpacking of the process
indicates the difficulties that a researcher faces when he wants to study subject matter
knowledge in teaching. The 17 codes that the research team initially identified were

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

grouped in the four elements of the Knowledge Quartet (foundation, transformation,


connection and contingency). Rowland discusses extensions of the codes through
the use of the Knowledge Quartet by other researchers. Finally, he brings some
examples from other studies reporting how the Knowledge Quartet has been used
emphasizing also its developmental character as a reflection tool for teachers and
as a framework that can be used in disciplines beyond mathematics. This chapter
helps us to understand methodological issues that are related to the study of teachers’
subject matter knowledge in a rather developmental way.

Mathematics Teacher Beliefs about Mathematics and Its Teaching

The chapter of Maria Bartolini Bussi, Silvia Funghi and Alessandro Ramploud
focuses on the lesson study model of collaboration and on the role of cultural beliefs
in the way that it has been operationalized around the world. The authors challenge
the transposition of the lesson study model from one country to the other if the
culture from which it originates is not taken into account. By analyzing prospective
and practicing teachers’ written reports in a questionnaire, they show that three
groups of teachers (one practicing and two prospective) held certain beliefs about
the lesson study process that explain the different ways in comparison to the Chinese
and Japanese that Italian teachers operationalize lesson study. The chapter develops
our understanding of the role of culture in mathematics teacher collaboration
contexts and make us aware about the need to take into account the role of culture in
supporting mathematics teachers to develop mathematics teaching.
The chapter of Esther Levenson focuses on research on prospective and practicing
teachers’ conceptions about mathematical creativity in school classrooms and on
how these develop in the context of professional development. On the basis of a
systematic literature review, Levenson discusses teachers’ conceptions of creativity
in terms of the nature of creativity, the characteristics of a creative person, and the
creative environment. She also indicates through an example from her own research
that mathematics teachers see mathematical creativity and mathematical excellence
interrelated in a rather complex way. Finally, she shows teacher education efforts
have been developed both for prospective and practicing teachers that promote their
awareness of the importance of developing students’ mathematical creativity and
offer them ways that this can be done. These teacher education practices engage
teachers to solve, analyse and design creative tasks. However, research in this area
shows that studies on teachers’ conceptions about mathematical creativity need to
be linked to what is taking place in a mathematics classroom by adopting more
qualitative methodological approaches. Moreover, teacher education programs
and professional development initiatives need to be designed and their impact on
developing mathematically creative students in the mathematics classroom needs
to be systematically studied. The chapter helps us to understand that the research
direction in our field that the development of mathematical creativity can be a
learning goal for all the students requires mathematics teachers to change their views

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING AND ITS DEVELOPMENT

and teaching practices in order to fulfil this goal. This is a big challenge both for
the mathematics teachers and mathematics teacher educators. The chapter offers us
ideas about how this can be facilitated.

The Interplay of Mathematics Teacher Identity, Beliefs and Knowledge

The chapter of Kim Beswick and Helen Chick offers a perspective of viewing
beliefs and knowledge in an integrated way while many studies have treated them
separately. In particular, they propose a framework that blends knowledge and
beliefs and they illustrate how they used this framework in a study where teachers
worked together discussing professional tasks that present students’ unexpected
solutions. The authors challenge existing frameworks on teacher knowledge that do
not consider, for example, affective dimensions of students’ learning. Their approach
is psychological, but as they claim, studying knowledge and beliefs in a context of
teacher collaboration brings to the fore social dimensions as well. The chapter offers
us a more systemic way of understanding teacher knowledge and beliefs and teacher
education practices that could encourage shifts that could improve mathematics
teaching and learning.
The chapter of Merrilyn Goos, Anne Bennison, Stephen Quirke, Niamh O’ Meara
and Collen Vale investigates teacher knowledge and identity in the case of non-
specialist teachers of mathematics. The authors present and compare the Deborah
Ball and colleagues’ model with the Knowledge Quartet of Tim Rowland identifying
similarities and differences. They then discuss identity as it is approached by
different perspectives and provide a systematic literature review on teacher identity
and its development. Through a description of the situation of out-of-field teaching
in four countries, they give reasons that have led to non-specialist teachers who
teach mathematics in these countries. Most studies show that these teachers lack
professional knowledge, but the authors extend this by discussing that there are
clear links between teacher knowledge and teacher professional identity. The authors
then bring an example from numeracy to illustrate the difficulty that the out-of-
school teachers face to embed numeracy into subjects other than mathematics and to
establish a new identity as a teacher who will embed numeracy in parallel with their
discipline identity. They then discuss professional development attempts with out-
of-field teachers. The chapter helps us to understand the close relationship between
knowledge and identity and also the dynamic character of identity development and
the multiple identities of the non-specialist teachers.

Mathematics Teaching and Its Development

The chapter of Marta Civil, Roberta Hunter and Sandra Crespo addresses the nature of
mathematics teaching and its development that aims for equity in terms of gender, race,
and ethnicity. The authors’ review of the literature identified three main perspectives
in equity in mathematics education: Inclusive Mathematics Teaching, Culturally

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

Sustaining Mathematics Teaching and Social Justice Mathematics Teaching. The


three approaches share commonalities but also place different emphasis on the role
of the teacher and the classroom interaction. The authors discuss these perspectives
as they have been considered in the literature and offer three examples of studies
that show the need for integrating these perspectives to get a better understanding of
the complexity of mathematics teaching aimed at equity and also to develop ways of
supporting teachers in their attempts to create environments supporting equity. The
chapter helps us to become aware of this complexity and of teaching practices that
can meet the diversity of students’ needs.
The chapter of Barbara Jaworski focuses on inquiry-based practice in university
mathematics teaching and its development. She uses the theory of community of
inquiry and, in particular, the construct of critical alignment and the three layers of
the inquiry process to describe mathematics teaching and its development. Jaworski
views teacher resources and practices as embedded in the activity of mathematics
teaching and in the teacher’s participation in projects that offer collective
experiences where inquiry is a central focus. The study of mathematics teaching
and its development at the university level bring new challenges and directions as
the mathematical activity is quite abstract, creating difficulties to students and the
instructional practices (lectures, tutorials, projects) and pose a lot of challenges to
the teachers. The chapter ends with addressing the need for moving from small
scale studies focusing on mathematics teaching and its development to large scale
contexts. In this situation, the issue of “transfer” gets an important meaning and it
requires changes in the way that the developmental process can operate.
Despina Potari and Konstantinos Stouraitis address teacher decision making,
pointing out its character, the sources on which the teacher bases his decisions,
and its links to other dimensions of noticing as well as to teacher knowledge. The
authors distinguish between decisions on the moment during mathematics teaching
and decisions made in the process of selecting curriculum materials and planning
teaching. They argue that decision making has been mainly examined in the context
of mathematics classroom and the focus is mainly on the individual teacher. Finally,
they offer an example from their own research where they used Activity Theory to
interpret the decision-making process of a group of teachers working collaborative
in reflecting on their pilot enactment of a new mathematics curriculum. The authors
offer this as an example of showing the dynamic nature of decision making that
captures social and historical dimensions.

From Mathematics Teaching Practices to Teacher Education

The chapter of Jill Adler and Craig Pournara addresses the importance of the use
of examples and variation in mathematics classroom and how exemplification
can be developed as a teaching practice in professional development contexts.
Through a current literature review, they point out that exemplification is important
in mathematics education and that it is connected to specialized knowledge

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING AND ITS DEVELOPMENT

needed for teaching. They present how exemplification has been promoted in a
longitudinal professional development project in South Africa in their attempts to
address the question of what and how exemplification is attended in mathematics
teacher education. In their education course, they engage teachers both as learners
of mathematics and as teachers by asking them to face tasks where examples are
varied in a systematic way on the basis of variation theory. The chapter’s focus
on a specific mathematical practice, exemplification, helps us to get insight of its
importance in mathematics teaching and mathematics teacher education. It also
offers us mathematics teacher education practices such as the variation of examples
that can prove effective to support teachers in designing and handling examples in
the mathematics classroom.
The chapter of Charalambos Charalambus and Sean Delaney offers a systematic
literature review of studies since 2000 that attempt to characterize core or high
leverage practices of mathematics teaching and how these are promoted in the
context of mathematics teacher education. It shows the large number of studies that
try to decompose teaching into distinct practices in different ways, often giving a
rather non consistent picture. The chapter moves to the notion of practice-based
pedagogies that has been used mainly in the United States in research on mathematics
teacher education and offers ways that this has been exploited in mathematics teacher
education. The chapter ends with recommendations of directions that research to
mathematics teaching and its development needs to take. As can be seen from the
chapter, the quality of mathematics teaching in relation to its impact on all students’
learning still remains an open issue. The decomposition of teaching and teacher
education offers ways to study teaching through the identified instructional practices.
However, it leaves open the study of mathematics teaching and mathematics teacher
education in terms of its complexity where both the classroom interaction and the
wider social and cultural contexts are taken into account.
The chapter of Randolph Phillipp, John (Zig) Siegfried and Eva Thanheiser
focuses on the development of elementary prospective teacher’s mathematical
understanding through studying students’ mathematical thinking. The authors
exemplify four assumptions that they consider critical in the way that understanding
mathematics and studying students’ mathematical thinking are related through an
example of division of fractions. The assumptions are related to epistemological
and pedagogical beliefs such as: mathematics is characterized by rich connections;
children learn from inside and outside of school experiences and can develop rich and
FUHDWLYHPDWKHPDWLFDOUHDVRQLQJH൵HFWLYHWHDFKLQJUHTXLUHVIRFXVLQJRQVWXGHQWV¶
mathematical reasoning; and teachers’ learning is a life-long process. The authors’
review of the research literature and discussion about the professional knowledge
for teachers of mathematics focuses on the development of understanding from the
noticing of students’ mathematical thinking. Through many examples of students’
ideas, they show how important mathematical ideas emerge and justify their argument
of the importance of focusing on these ideas through students’ mathematical thinking
and reasoning. They also share approaches that they have used with prospective

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI

teachers involving students’ thinking as the basis for developing deep mathematical
understanding. The chapter helps us to understand on one hand the importance of
blending mathematics and pedagogy on mathematics teacher education and on the
other how this can be realised in our work with mathematics teachers.

ȂOVING AHEAD
Volume 1 of this second edition of the International Handbook of Mathematics
Teacher Education extends the work reported in the first edition of the Handbook
in several directions. There is an increasing interest in the quality of mathematics
teaching and in the mathematics teaching practices that can meet the needs of all
the students in the classroom. Teacher knowledge, beliefs and identity are studied
in a more interrelated way and there is further realization of the complexity of
mathematics teaching. The quality of mathematics teaching has started to be
discussed more in terms of students’ cognitive, affective and cultural needs and
research considers the new challenges for mathematics teachers in the direction of
developing ways of supporting them. Peter Sullivan, in his introductory chapter of
Volume 1 of the first edition of the Handbook (Sullivan, 2008), listed a number of
teacher actions and decisions that may characterize effective mathematics teaching.
He argues that we need to learn more about different aspects of teacher knowledge
about “pedagogy, student management, interpersonal relationships, historical
perspectives, cultural influences and differences, social disadvantage, linguistic
challenges and so on” (p. 8). The chapters of this volume add to our understanding
of this complexity and offer ways that we could support teachers of mathematics
to handle it.

REFERENCES
Anthony, G. & Walshaw, M. (2009). Characteristics of effective teaching of mathematics: A
vision from the west. Journal of Mathematics Education, 2(2), 147–164.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What
makes it special? Journal of Teacher Education, 59(5), 389–407.
Bartell, T., Wager, A., Edwards, A., Battey, D., Foote, M., & Spencer, J. (2017). Toward a
framework for research linking equitable teaching with the standards for mathematical
practice. Journal for Research in Mathematics Education, 48(1), 7–21.
Cai, J., Kaiser, G., Perry, B., & Wong, N. (2009). Effective mathematics teaching from
teachers’ perspectives. Rotterdam, The Netherlands: Sense Publishers.
Charalambous, C. Y. (2015). Working at the intersection of teacher knowledge, teacher beliefs,
and teaching practice: A multiple-case study. Journal of Mathematics Teacher Education,
18(5), 427–445.
Cobb, P., Wood, T., & Yackel, E. (1990). Classrooms as learning environments for teachers
and researchers. In R. B. Davis, C. A. Maher, & N. Noddings (Eds.), Journal for
research in mathematics education: Constructivist views on the teaching and learning of

10

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING AND ITS DEVELOPMENT

mathematics (Monograph No. 4, pp. 125–146). Reston, VA: National Council of Teachers
of Mathematics.
Cross, D. I. (2009). Alignment, cohesion, and change: Examining mathematics teachers’
belief structures and their influence on instructional practices. Journal of Mathematics
Teacher Education, 12, 325–346.
Graven, M., & Heyd-Metzuyanim, E. (2019). Mathematics identity research: the state of the
art and future directions. ZDM Mathematics Education. https://doi.org/10.1007/s11858-019-
01050-y
Jaworski, B. (1994). Investigating mathematics teaching: A constructivist enquiry. London:
Falmer Press.
Losano, L., Fiorentini, D., & Villarreal, M. (2018). The development of mathematics teacher’s
professional identity during her first year teaching. Journal of Mathematics Teacher
Education, 21, 287–325.
Lunkenbein, D. (1997). Rationalising teaching interventions - A working model of a process
of research in mathematics teaching. Educational Studies in Mathematics, 8(3), 271–293.
Miyakawa, T., & Winslow, C. (2013). Developing mathematics teacher knowledge: The
paradidactic infrastructure of “open lesson” in Japan. Journal of Mathematics Teacher
Education, 16(3), 185–209.
Ntow, F. D., & Adler, J. (2019). Identity resources and mathematics teaching identity: An
exploratory study. ZDM Mathematics Education. https://doi.org/10.1007/s11858-019-
01025-z
Ponte, J. P., & Chapman, O. (2006). Mathematics teachers’ knowledge and practices. In A.
Gutierrez & P. Boero (Eds.), Handbook of research on the psychology of mathematics
education: Past, present and future (pp. 461– 494). Rotterdam, The Netherlands: Sense
Publishers.
Skott, J. (2009). Contextualising the notion of ‘belief enactment’. Journal of Mathematics
Teacher Education, 12(1), 27–46.
Skott, J. (2015). The promises, problems, and prospects of research on teachers’ beliefs.
International Handbook of Research on Teachers’ Beliefs, 1, 37–54.
Sullivan, P. (2008). Knowledge for teaching mathematics: An introduction. In P. Sullivan
& T. Wood (Eds.), The International handbook of mathematics teacher education:
Knowledge and beliefs in mathematics teaching and teaching development (Vol. 1,
pp. 1–9), Rotterdam, The Netherlands: Sense Publishers.
Thompson, A. G. (1984). The relationship of teachers’ conceptions of mathematics and
mathematics teaching to instructional practice. Educational Studies in Mathematics, 15(2),
105–127.
Wilson. P. S., Cooney, T. J., & Stinsion, D. W. (2005). What constitutes good mathematics
teaching and how it develops: Nine high school teachers’ perspectives. Journal of
Mathematics Teacher Education, 8(2), 83–111.

Despina Potari
Department of Mathematics
National and Kapodistrian University of Athens

11

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PART 1
MATHEMATICS TEACHER KNOWLEDGE AND ITS
RELATION TO TEACHING

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW AND HAMSA VENKAT

1. MATHEMATICAL SUBJECT KNOWLEDGE FOR


TEACHING PRIMARY SCHOOL MATHEMATICS
Evidence and Models for Professional Development

This chapter examines the research evidence – theoretical and empirical – that
purports to elaborate the models of mathematical discipline knowledge that primary
(elementary) school teachers need to teach effectively and so improve student
learning. Key questions addressed include whether or not such models have brought
the mathematics education community close to understanding the actual content of
the subject knowledge that primary teachers might need and whether it is helpful
to continue to try and distinguish content knowledge from pedagogical content
knowledge.

INTRODUCTION

Debate as continued for years over the answer to the question of what, exactly,
mathematical knowledge primary (elementary) school teachers need in order to
teach effectively (‘primary’ taken as schooling for children aged five to 11, and
the terms primary and elementary are used interchangeably here). Over a century
ago Dewey (1904, 1964) argued that teachers need to be familiar with the nature of
inquiry in particular domains. But it was not always thus: in medieval universities
no distinction was made between knowledge of a discipline and knowing how to
teach it – if you knew the former, it was assumed that you would be able to teach it
(McNamara, Jaworski, Rowland, Hodgen, & Prestage, 2002).
Today, apprenticeship models of learning still draw no distinction between being
an expert craftsperson and being able to induct an apprentice into that craft. With,
however, the introduction of, almost, universal primary schooling has evolved
the separation of knowledge and practice in a discipline from the teaching of that
discipline. Teaching moved from being part of ongoing practices within disciplines
to becoming a practice in its own right. With this separation of doing from teaching,
discipline knowledge, rather than being able to be taken as a given, is problematized
as teachers, in primary schools in particular, need to impart knowledge about
disciplines that they are not necessarily part of. With the severing of knowing (doing)
and teaching questions then emerge about the extent and form of the discipline
knowledge required for teaching.

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_002

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

This history, together with common sense, suggests that an effective teacher of
mathematics would need to have a good understanding of the discipline, so it is
hardly surprising that researchers are interested in the mathematics that primary
school teachers may need to know. But is that assumption correct? The evidence for
exactly what primary school teachers (both prospective and practicing) actually need
to know about mathematics is still far from being clearly elaborated. This chapter
explores something of the current state of the debates around the mathematical
subject knowledge needed for teaching and suggests some directions towards which
the argument may go.
Although the main focus of the chapter is on practicing teachers, we also draw
on the research into developing the knowledge of prospective teachers as this has
been widely studied. Looking at what experienced primary school teachers need to
know may help to clarify not only what might need to be addressed in professional
development, but also what might be focused on with prospective teachers. Our gaze
thus encompasses both of prospective and practicing teachers, taking ‘professional
development’ as something that occurs both pre- and in-service.
The chapter first looks at the evidence for the strength of findings that primary
teachers’ mathematics subject knowledge has a direct link to the quality of teaching.
On the basis that the picture of any such link is a complex one we then turn to
examine whether research has come any closer to identifying the precise nature of
the knowledge that primary school teachers might need. Here again, we find the
picture is far from clear, so in the second half of the chapter we turn to examine
alternative perspectives on the issue. First, rather than try to pin down the content of
the mathematical subject knowledge that primary teachers might need, we suggest
that energy might be better put into looking at ways in which primary teachers
might develop their mathematical understanding – that a more dynamic view of
the development of mathematical subject knowledge needs to be worked with,
rather than treating knowledge as a body of ‘static’ content. From that perspective,
if primary school teachers’ mathematical content knowledge is to be treated as a
‘work in progress,’ then to promote that teachers might be encouraged to develop a
particular sensibility towards mathematics, a theme we take up in the penultimate
section. Finally, we look at arguments for a paradigm shift in this research – away
from the study of individual’s knowledge, to treating knowledge as a collective
enterprise.

MATHEMATICS SUBJECT KNOWLEDGE: THE LINK TO QUALITY OF


TEACHING AND LEARNING

The year 1986 provides a marker for the start of serious studies into the nature of subject
knowledge for teaching as that year marked the publication of the seminal paper on
knowledge for teaching; Lee Shulman’s ‘Those who understand: knowledge growth
in teaching’ (1986). Also published that year was the third edition of the Handbook
of Research on Teaching (Wittrock, 1986), in which, as Connolly, Clandinin, and He

16

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

(1997, p. 666) note, there were only two, relatively minor, references to research into
teacher knowledge. From the perspective of 1997, Connolly et al. suggest that since
1986 the field had ‘exploded’ and this explosion has not diminished in the years
since: this chapter is necessarily selective of the literature.
As well as the huge expansion of studies over the last 30 years or so, another
significant change continues to foreground the importance of teacher subject
knowledge. That change, in many parts of the world, is marked by increased political
and policy involvement in the setting out and defining of the subject knowledge
for teaching both generally and specifically for mathematics (Poulson, 2001).
Defining and codifying of subject knowledge for teaching is taken by many policy
makers as central to improving standards of pupil learning. In England, for example,
much policy effort and public funding has gone into attempts to improve teachers’
subject knowledge: centrally determined curricula have been set out for prospective
teacher educators to follow, prospective teachers have had their subject knowledge
‘audited’, and online teacher tests of ‘numeracy’ have been developed. It is worth
asking whether the research findings into the importance of subject knowledge
warrant such efforts.
Across the studies into teacher knowledge, two specific questions recur: What is
the relationship between teachers’ mathematics subject knowledge and the teaching
and learning of mathematics in primary schools? and What sort of mathematics
subject knowledge do primary school teachers need to know?
In a fully rational world, the second question might only be addressed once the
first had been clearly answered. But if, as indicated, common-sense suggests that
there must be a relationship between discipline knowledge and the ability to teach,
then work can proceed on the second question prior to the first being answered.
Indeed, despite mixed answers to the first question, work continues unabated on the
second. For example, in a major study into young children learning mathematics that
continues to be influential, Kilpatrick and colleagues (National Research Council,
2001) addressed these questions in reverse order, looking first at “knowledge of
mathematics” and then “teachers’ mathematical knowledge and student achievement”
(pp. 372–377). With regard to knowledge of mathematics, the authors claim that
content knowledge is “the cornerstone of teaching for proficiency” and “improving
teachers’ mathematical knowledge and their capacity to use it to do the work of
teaching is crucial in developing students’ mathematical proficiency” (p. 372).
Just over a page later, however, in examining the links between teacher content
knowledge and student achievement, they report: “For the most part, the results have
been disappointing: Most studies have failed to find a strong relationship between the
two” (p. 373).
A similar conclusion is arrived at in a North American review of research that
examined what kind of subject matter preparation is needed for prospective teachers
(Wilson, Floden, & Ferrini-Mundy, 2001). Drawing on rigorous selection for
the inclusion of studies, the authors could find only seven studies, four of which
included mathematics, and of these four only one addressed the mathematics subject

17

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

knowledge of elementary school teachers (Darling-Hammond, 2000). Wilson and


her colleagues report that the conclusions from these studies in establishing the
impact of subject knowledge are contradictory and “undermine the certainty often
expressed about the strong link between college study of a subject matter area and
teacher quality” (p. 6). The important thing to note here is the phrase ‘college study’–
that study of courses designed for students studying mathematics but not preparing
to be teachers, may not be the best preparation.
Much of the research on teachers’ subject knowledge pays homage to Shulman’s
work on forms of knowledge for teaching, but less often acknowledged is the
cautiousness that Shulman and his colleagues expressed over generalizing the range
of their work to primary school teachers (Grossman, Wilson, & Shulman, 1989;
Shulman, 1987; Wilson, Shulman, & Richert, 1987). In his 1987 paper, Shulman
expresses his belief that findings from the research carried out with secondary school
teachers may well apply to primary teachers but was “reluctant to make that claim
too boldly” (p. 4). A little later, Grossman and colleagues (1989) pointed out that
imputing implications from their work in secondary schools “for elementary school
teaching should be drawn cautiously” (p. 28) given the considerable differences
between teaching only one subject in secondary schools and the demands of teaching
several subjects in primary schools.
Despite such caveats, one thing emerging from the early research into subject
knowledge is that many prospective and practicing primary school teachers have, or
express, a lack of confidence in their mathematical knowledge. Wragg, Bennett, and
Carré (1989) surveyed teachers from 400 primary schools in Great Britain and found
a self-reported lack of confidence in their knowledge of the mathematics required
to teach the national curriculum (although these same teachers expressed an even
greater lack of confidence about science). In her study of teaching 4- to 7-year-
olds, Aubrey (1997) found that teachers claimed not to have extensive knowledge
of mathematics. Many other studies have found similar results (Bennett & Turner-
Bisset, 1993; Rowland, Barber, Heal, & Martyn, 2003; Simon & Brown, 1996).
These early studies, particularly, but not exclusively from the United States of
America and the United Kingdom, tend to point to a deficit model of teachers’
knowledge – what primary school teachers do not know, rather than what they actually
know – resulting in, at least, two outcomes. Firstly, attending to what primary teachers
did not know fanned the flames of a ‘crisis’ in the knowledge base of the teaching
forces. Even later on, findings were still couched in terms of the negative: Stacey et
al. (2003), for example, report that “only 80% of the sample tested as experts” (p.
205). While it might be reasonable to expect 100% to be proficient, at 80% the pot is
certainly much more full than empty. Second, it was imputed that improving teachers’
subject knowledge would necessarily improve teaching and learning (Askew, Brown,
Rhodes, Wiliam, & Johnson, 1997a; Brown, Askew, Baker, Denvir, & Millett, 1998).
However, such a conclusion needs to be treated cautiously.
Askew and colleagues (Askew & Brown, 1997; Brown et al., 1998) have argued
that while there may be evidence that teachers lack subject knowledge when this is

18

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

assessed outside the context of the classroom, whether or not this is a hindrance in
practice is more difficult to establish. For example, in their study of effective teachers
of numeracy (number sense) Askew and colleagues observed only two occasions
where teachers appeared to be hampered by a lack of mathematical knowledge,
out of 86 mathematics lessons observed in total. They concluded that while some
teachers of younger children may have problems with their understanding of subject
knowledge, it was not clear how much this actually impacted on their effectiveness
(Askew et al., 1997a, p. 59).
Similarly, Bennett, and Carre (1993) found that prospective primary teachers
who were mathematics specialists did display greater subject knowledge than
those specialising in other areas (music and early years). However, when they
observed all these prospective teachers actually teaching, the researchers found
little difference in their practices that distinguished the mathematics specialists
from the non-specialists teaching mathematics. These observations took place
whilst the teachers were still in their teacher education program, so we do not
know if there was a longer-term impact of the mathematics specialists’ greater
subject knowledge; it may be that novice teachers share a common set of concerns,
irrespective of their subject specialism. Concerns centred around becoming
familiar with the curriculum, planning lessons, managing 30 or more children, and
building productive classroom relationships. Such concerns initially may over-
ride attention to the particularities of subject content; in moving from being novice
teachers to experienced ones, specialized content knowledge may begin to have
an impact.
It would seem, nevertheless, that a certain threshold of discipline knowledge
is necessary for effective teaching. Whilst a major longitudinal study carried out
in the 1960s could find no association between teachers’ study of higher-level
mathematics and their students’ achievement, the researcher later argued that
teachers need to attain a certain level of mathematical understanding, but that
beyond a certain level further study of mathematics did not lead to increased
student gains (Begle, 1979). A later study found associations between student
attainment and the number of mathematics courses that their teachers had studied,
but only up to a certain number of courses (although this was with secondary
school teachers, there is no reason to suppose the results would be different for the
primary level) (Monk, 1994).
A major difficulty in establishing the impact of discipline knowledge is the
means by which teachers’ mathematical knowledge has been identified and
quantified. Until recently, studies have often relied on proxies such as highest level
of formal qualification in mathematics or courses taken rather than looking ‘inside
the black box’ at the mathematics that teachers actually draw on when teaching.
The use of college courses studied as a proxy for subject knowledge partly explains
why Hattie’s 2018 update of his original meta-analyses of research (Hattie, 2012)
ranks teacher subject knowledge at 213 out of 252 attributes, with an effect size
on learning as only 0.11. Hattie interprets this figure (i.e., 0.11) as likely only to

19

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

have small positive impact on student achievement, compared with the top factor –
teacher estimates of student achievement – having an effect size of 1.62, interpreted
as potentially able to accelerate student achievement considerably. Caution needs
to be exercised in suggesting that subject matter does not count, on the basis
of this evidence. What needs to be interrogated is what, in such meta-analyses,
gets ‘counted.’ As noted, college courses taken by teachers are well established
as not being good predictors of student learning, but rather than dismiss subject
knowledge as unimportant, we need better, quantifiable, measures of it (https://
visible-learning.org/hattie-ranking-influences-effect-sizes-learning-achievement/).
The fact that proxies such as courses studied have not demonstrated a link between
subject matter knowledge and teaching expertise or pupil outcomes is hardly
surprising as there are at least two problems.
Firstly, we need to question the assumption that examination results are an accurate
reflection of someone’s level of understanding for teaching. Success on formal
mathematics examinations may be attained through a base of procedural rather
than conceptual understanding, and thus not be measuring the sort of mathematics
that underpins successful teaching. This conjecture is supported by research carried
out in England. Observations of lessons conducted by teachers with higher formal
mathematical qualifications tended to be more procedural in their content. These
same teachers, in interview, expressed difficulty in understanding why some pupils
had problems with mathematics. Further, there was a slight negative association
between the gains over the course of a year that the pupils made on a specially
designed number assessment and the highest level of mathematical qualification
of their teachers: the higher the level of qualification, the lower the gains the
pupils made (Askew, Brown, Rhodes, Wiliam, & Johnson, 1997b). Second, even
if qualifications are an accurate measure of understanding, are they the ones that
teachers will need to draw on in the classroom? As discussed below, only recently
researchers have begun to try and ‘unpack’ the knowledge specifically needed from
primary school teaching.
Hoover, Mosvold, Ball, and Lai (2016) reviewed the literature related to
mathematical knowledge for teaching published between 2006 and 2013, in English
language journals. From an initial set of 349 articles identified, 190 were included
in the final selection. The authors note that many of these were small-scale studies,
and many used either no measures of teachers’ knowledge or used non-standardized
measures. Most of the research was on primary school teachers, and most was
carried out in North America. The field, while continuing to grow, thus still seems
to be disparate and fragmented in terms of useable knowledge. The key question is
whether or not we are closer to understanding the quality and content of mathematical
knowledge for teaching, to which we now turn.
To summarize, there is some agreement in the literature on prospective and
practicing primary school teachers’ mathematical knowledge in the conclusion
that a certain lack of knowledge of mathematics is associated with less successful
teaching and lower student attainment. The flip-side of this argument – that more

20

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

subject knowledge is linked to better teaching and learning – is historically less


well established, although research is beginning to examine this in more detail (See
Rowland, Martyn, Barber, & Heal, 2000). Recent studies examining this have begun
to examine more closely the nature and content of discipline knowledge needed by
primary school teachers, and it is to these that we now turn.

MATHEMATICS SUBJECT KNOWLEDGE: FROM WHETHER TO WHAT?

In this section, we look at ways that various writers have classified mathematics
subject knowledge for primary teachers. The work of Ball and her colleagues
(e.g., 2008) has been particularly influential in the study of the mathematics
for teaching and most classifications bear strong resemblances to frameworks
that they have set. In an early work, Ball (1990) argues for attention to the
distinction between knowledge of mathematics – the various meanings attached
to representations and procedures – and knowledge about mathematics – the
means by which ‘truth’ is established within the discipline. This distinction echoes
Lampert’s (1986) separation of procedural and principled knowledge, the former
involving ‘knowing that’ and includes the rules and procedures of mathematics
(‘knowledge of mathematics’ in Ball’s terms). Principled knowledge, for Lampert
is more conceptual – the knowing why of mathematics. Other similar distinctions
are between the substantive (facts and concepts) and the syntactic (nature of
knowledge growth in the field through inquiry and, in the case of mathematics,
proof) (Shulman, 1986).
Even earlier was Skemp’s (1976) setting out of instrumental or relational
understanding; argued for through the metaphor of a map. For Skemp, instrumental
understanding was analogous to being given step-by-step instructions for getting
from A to B, with relational understanding being more akin to having a map; so, if
one gets lost then one has the wherewithal to figure out the way back to the right
path. Many people’s understanding of traditional algorithms would be described
as instrumental in that they know the steps to follow to find an answer, but not
relational as they would have difficulty reconstructing the algorithm if forgotten or
would not be able to explain why it works.
In a similar vein, Thompson, Philipp, Thompson, and Boyd (1994) discuss the
difference between calculational and computational perspectives, although this
distinction is more akin to beliefs about the curriculum than about mathematical
knowledge per se. It could be perfectly possible for a teacher to have a rich and
varied understanding of the mathematics curriculum and yet still hold that learning
basic computational skills is the goal of primary school mathematics teaching.
Common across all such models is the importance of understanding not only factual
knowledge and central concepts but also the organising principles of a discipline – a
point we return to in the final section.
While distinctions between conceptual/procedural, instrumental/relational and
so forth seem reasonable, closer examination reveals some difficulties in setting

21

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

out exactly what these distinctions represent. Suppose a teacher knows how to
carry out a multiplication algorithm accurately but cannot articulate how it works.
Is the ability to correctly carry out the calculation any less ‘conceptual’ than being
able to explain how it works? Research at King’s College, London, supported the
conjecture that procedural knowledge of mathematics is not sufficient. As part
of a five-year longitudinal study of teaching and learning numeracy in primary
schools (the Leverhulme Numeracy Research Programme, Millett, Brown, &
Askew, 2004), researchers interviewed a group of teachers about a variety of
mathematical problems. The teachers’ responses indicated that they could, on the
whole, arrive at correct answers, but when it came to probing the rationales behind
these answers, it became clear that their solutions were based on instrumental/
procedural approaches. The difficulty was that while teachers were able to obtain
correct answers, they were not able to ground their approaches to finding solutions
in knowledge of the generalities of mathematics that they might underpin effective
teaching.
Ball, Thames, and Phelps (2008) addresses this point through making the
distinction between common content knowledge (CCK) and specialized content
knowledge (SCK), the former being the sort of mathematical knowledge that
teachers might be expected to share with the population in general and the latter the
sort of mathematical knowledge that might be necessary for teaching. In the case
of specialized content knowledge, it might be expected that teachers do need to
know how the multiplication algorithm works, in the expectation that they may need
to draw on such knowledge in order to be able to decide if an alternative method
devised by a student would have general applicability. That, however, carries an
assumption that the curriculum goes beyond teaching the standard algorithm and
expands to include encouraging student creativity. In countries with a more limited
curriculum, such as South Africa (see DBE, 2011), the emphasis is more on teaching
the one standard method, and in such circumstance specialized content knowledge
is going to be locally and culturally determined rather than universal. Ball (1991a,
1991b) also identified teachers with good understanding of mathematics but who
still adopted ‘transmission’ style teaching approaches rather than work on crafting
student explanations, suggesting that looking at the mathematics that teachers draw
on in mathematics lessons could be as much a consequence of their beliefs about the
role of mathematics in the curriculum as about their mathematical subject knowledge
per se.
Thus, a number of distinctions have been drawn up which are now largely taken
as descriptive; the most notable distinction being between discipline knowledge
and pedagogical knowledge. Within discipline knowledge the distinction is made
between ‘facts, concepts and procedures’ – the established cannon of mathematical
knowledge – and the way that the community of mathematicians has come
to establish this knowledge. Within pedagogical knowledge there is a similar
separation of knowledge of how to teach particular mathematical topics (didactics)
and knowledge of how the individual learner might develop understanding

22

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

(psychology). In Davis and Simmt’s (2006) terms these are assumed distinctions
between established/dynamic and collective/individual, arguing that the distinctions
between “formal disciplinary knowledge and instructional knowledge … between
established collective knowledge and dynamic individual understandings … are
inherently problematic” (p. 293). This point is returned to later.
Using a series of case studies and quantitative data to examine how mathematical
knowledge for teaching is associated with the mathematical quality of instruction,
Hill et al.’s (2008) findings report “a significant, strong, and positive association
between levels of [mathematics knowledge for teaching] and the mathematical
quality of instruction” (p. 430) whilst still noting that a number of other factors either
supported or hindered knowledge in practice.
Where attention has turned to looking in detail at the content of teachers’ subject
knowledge the focus is often on the quality of understandings rather than the actual
content of them. As Freudenthal (1975) expressed it, it may be that more nuanced
understandings are required. Ball and Bass (2003) similarly argue that mathematics-
for-teaching knowledge is not a case of knowing more than or to a greater ‘depth’
than that expected of students; it needs to be qualitatively different.
Ma (1999) expressed qualitative differences through her construct of ‘profound
understanding of fundamental mathematics.’ Such understanding was exhibited by the
Chinese teachers in her study, whilst the United States teachers presented knowledge
that was lacking in conceptual underpinning and ‘unconnected’. Powerful though
Ma’s evidence is, it still only partially sets out what such ‘profound understanding of
fundamental mathematics’ might look like. Her detailed qualitative work addresses
a small number of the topics in the primary school mathematics curriculum. In
addition, the teachers involved in Ma’s study were specialist mathematics teachers
and so not representative of the majority of primary school teachers.
Nevertheless, the metaphor of teachers needing to have ‘connected’ mathematical
knowledge resonates through the research. Askew and colleagues’ (1997a) study of
effective teachers of numeracy in primary schools engaged teachers in constructing
‘concept maps’ of understanding number (Novak & Gowin, 1984). In interviews,
teachers were asked to list as many topics in learning about number in the primary
years as they could. Each topic was noted on a separate ‘stickie’ note. The teachers
then ‘mapped’ these out onto a larger sheet of paper. Once the locations of these topic
‘landmarks’ had been chosen, the teachers drew arrows connecting topics: arrows
could be one or two headed to indicate directions of links, and topics could have
multiple connections. Finally, and importantly, they were asked to label the arrows
to provide concise descriptions of the nature of the connections. The completed
concept maps were analyzed in terms of the numbers of connections identified,
the range of the connections and the quality of the descriptions of the links. The
researchers found that there was a strong association between the complexity of
the maps that teachers produced and the average gains in scores that their classes
attained on a numeracy assessment over the course of a year. Where teachers had
more ‘connected’ maps, the gains for classes were higher (Askew et al., 1997a).

23

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

More recently, Hough, O’Rode, Terman, and Weisglass (2007) looked at how
concept maps could also be used with prospective teachers to help them think about
connections in mathematics.
In many studies, conclusions about teachers’ knowledge are based on interviews
rather than observation of actual classroom practice. To address this, Ball and her
colleagues turned to taking an approach more grounded in practice by looking at the
demands made in classrooms (Ball, Hill, & Bass, 2005). The question is whether
or not working on areas such as those identified by Ball et al. are likely to improve
teaching in contexts far removed from those originally studied? And it may be that
the mathematical behaviours demonstrated by effective teachers are still proxies for
something else – a mathematical sensibility – that cannot be reduced to a list of
mathematical topics. Transfer across the curriculum is also an issue. For example, as
noted above, an item from the Ball research asks teachers to judge which of several
‘non-standard’ vertical algorithms would work for any multiplication. If teachers
are able to answer this correctly, does that indicate a similar ability to unpack non-
standard methods for, say, division?
One of the few reports to set out the actual range of subject knowledge that
prospective elementary school teachers might be expected to be confident in is
presented by the Conference Board of the Mathematical Sciences (2001). In terms
of number and operations, the authors argue that
Although almost all teachers remember traditional computation algorithms,
their mathematical knowledge in this domain generally does not extend much
further. … In fact, in order to interpret and assess the reasoning of children
learning to perform arithmetic operations, teachers must be able to call upon a
richly integrated understanding of operations, place value, and computation in
the domains of whole numbers, integers, and rationals. (p. 58)
When, however, it comes to adumbrating the learning needed, there is little that looks
different from what one might find in a typical primary school mathematics syllabus.
Teachers, for example, are expected to develop “a strong sense of place value in
the base-10 number system” which includes “recognizing the relative magnitude of
numbers” (p. 58). Readers are left to decide for themselves exactly what constitutes
a ‘strong sense’.
Part of the problem is the codification of networks of knowledge into discrete lists
of items, as Davis and Simmt (2006) note. Working with a group of teachers to elicit
the range of representations that might be grouped together as multiplication, Davis
and Simmt conclude that “multiplication was not the sum of these interpretations.
… we conjecture that access to the web of interconnections that constitute a concept
is essential for teaching” (p 301), thus challenging attempts to delineate and list the
elements of mathematical knowledge:
What is multiplication? has no ‘best’ or ‘right’ answer. Responses, rather, are
matters of appropriateness or fitness to the immediate situation. The underlying

24

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

notion of adequacy that is at work here stands in contrast to the pervasive


assumption that mathematical constructs are unambiguous and clearly defined.
(Davis & Simmt, 2006, p. 302)
Formulating content lists is prey to another difficulty: the assumption that the
traditional curriculum will continue in future years to be the bedrock of primary
school mathematics. For example, will it continue to be the case that central to the
curriculum is “children learning to perform arithmetic operations” and, as Ball, Hill,
and Bass (2005) assert, that they need to be “able to use a reliable algorithm to
calculate an answer” (p. 17)? Some curricula, for example in Australia, have already
tossed onto the curriculum scrap-heap the ‘reliable algorithm’ for long division, with
the view that new technologies have produced more efficient and reliable methods
for carrying out such calculations. It may be the case that other algorithms are
regarded similarly in the future. Assuming that the primary school curriculum need
not change very much in the near future may become a self-fulfilling prophecy if
too much effort goes into specifying the knowledge that teachers need to know.
Rather than acquire a ‘body’ of mathematical knowledge, perhaps primary school
teachers need something else – a mathematical sensibility –that would enable them
to deal with existing curricula but also be open to change. For example, in a large-
scale study, Sullivan, Clarke, and Clarke (2009) analysed questionnaire responses
from 107 primary and secondary school teachers, concluding that a teacher’s
mathematical knowledge for teaching may limit their ability to design productive
learning experiences. The question such a study raises is whether most teachers
actively engage in designing tasks – if curriculum resources provide good tasks, to
what extent does mathematical knowledge impact upon the implementation of such
tasks.
Charalambous (2010) suggests that mathematical knowledge does have an impact.
In examining how the mathematical knowledge of two primary school teachers
appears to influence the unfolding of classroom tasks, tentatively hypothesising
that better mathematical knowledge may be linked to more effective use of
representations, explanations and working with students’ understanding, with, in
contrast, weaker mathematical knowledge leading to more procedurally, rule-based
teaching.
The links between mathematical subject knowledge for teaching, teaching itself
and consequently learning outcome is thus not simple. While there is evidence that
poor mathematical knowledge is linked to poor teaching and teachers with good
mathematical knowledge demonstrate better teaching practices, Hill, Umland,
Litke, and Kapitula (2012) report considerable variation in the quality of teaching
and learning outcomes of teachers, demonstrating that mathematical knowledge sits
between these extremes.
The picture gets even more complicated when trying to find the chain of connection
between mathematical knowledge, classroom teaching and learning outcomes.
Picking up on the links between knowledge, practice and learning outcomes from

25

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

the 190 articles selected in the earlier literature review, Mosvold and Hoover (2017)
identified only 12 that presented findings on how mathematical knowledge actually
influenced practice. The only one of these studies that presented quantitative data is
the Sullivan, Clarke, and Clarke study (2009). Mosvold and Hoover conclude that it
may actually be more helpful to focus on the doing of teaching and begin to develop
shared understandings of the actual work of teaching. This could be a valuable
move forward in our knowledge. But it raises further questions about to what extent
good practices are universal and what aspects of teaching are more context specific.
Mathematics education research is not immune from critiques being levelled
at psychological research more generally, that is, it is based on studies of WEIRD
(Western, educated, industrialized, rich, and democratic) populations (Rad,
Martingano, & Ginges, 2018). Any science that is representative of the range of
mathematics teaching must also be representative of the range of conditions in which
it takes place. We would agree with Mosvold and Hoover’s (2017) suggestion for the
need for large scale impact studies. But the question of what conceptual frameworks
and measurement would need to be developed is a thorny issue: the work to date on
mathematical knowledge for teaching is replete with frameworks and measures, but
low on usable knowledge. Part of the issue here is the position taken on the origins
of subject matter, to which we now turn.

SUBJECT KNOWLEDGE FOR TEACHING: DISCOVERED OR CREATED?

An issue arising from early work into subject matter knowledge was a lack of
theorizing of how discipline knowledge might inform and come to play out in
practice. As Leinhardt and Smith (1985) point out, there was a lack of problematizing
the issue of transfer: “No one asked how subject matter was transformed from the
knowledge of the teacher into the content of instruction” (p. 8). But this assumes
that there are different knowledge ‘packages’ to be transported and transformed. It is
built upon an objectivist epistemology.
Shulman’s models of knowledge have elements of an objectivist epistemology:
knowledge comprises objects located in the minds of individual teachers. While
philosophers of mathematics question an objectivist epistemology of mathematics
(Ernest, 1998) research that continues to treat subject knowledge as objects may
fix us in a teacher-centred pedagogy, with the assumption that the main source of
learning comes about through the pre-existing knowledge of the teacher.
Davis and Simmt (2006) argue that the practices of research mathematicians involve
creating concise expressions of mathematics, through ‘compressing’ information,
which in the final stage of their work in the process of producing mathematics, things
look different. Teachers, in contrast, have the opposite task and need to be “adept
at prying apart concepts, making sense of the analogies, metaphors, images, and
logical constructs that give shape to a mathematical construct” (p. 301). Ball and Bass
(2000) regard this ability to ‘unpack’ mathematics as an aspect of pedagogical content
knowledge. This raises the question of how easy and/or necessary it is to separate

26

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

subject knowledge from pedagogic knowledge. If mathematical constructs are shaped


through “analogies, metaphors, image, and logical constructs” then are these part of
pedagogical content knowledge or part of subject knowledge? And does it matter?
The answer comes down to a philosophical position on the epistemology of
mathematics. If one believes that there are idealized mathematical forms that exist
independently of representations, illustrations, examples and so forth, that there is
a definite signifier/signified distinction (Walkerdine, 1988), then such ‘unpacking’
(or re-packaging) is going to be seen as a pedagogic skill. If, on the other hand, one
views mathematics as a ‘language game’ (Wittgenstein, 1953) only brought into
being through representations, illustrations, examples and not existing outside these,
then this is an aspect of subject knowledge as much as pedagogic.
As McNamara (1991) argues, all mathematics is a form of representation.
Similarly, Askew (1999) questions whether attempts to classify different types of
teacher knowledge and to separate these from beliefs is possible, or desirable, as
all such propositional statements are brought into being through discourse and any
classification into discrete categories can only be established within social relations.
Barwell (2013) argues that the discourse of Shulman’s model is infused with
a representational epistemology that implies that individuals ‘have’ internal and
underlying concepts that are then represented in some form of external representation.
The follow through to this epistemological position is that there is a ‘body’ of
knowledge needed for teaching mathematics that exists independently of the actual
practices of teaching, and that such knowledge can be revealed through teacher
tests or revealed in classroom teaching. Barwell argues that this representational
epistemology still underpins much of the research following Shulman, and that
although there are moves towards a situated perspective on mathematical knowledge
for teaching, escaping from the discourse of knowledge as a possession of the teacher
is difficult, for example, as revealed by talk of ‘gaps’ in teacher knowledge (Hodgen,
2004) or studies purporting to demonstrate that teachers have stronger or weaker
knowledge. A key question missing from much of research, Barwell argues, asks
‘What, then, is knowledge?’ (p. 597)
Thus, distinctions between subject knowledge, pedagogic knowledge, semantic/
syntactic, product/process can be regarded as being constructed within the discourse
of research literature, rather than being discovered as independently existing
‘objects. In line with Vygotsky’s observation, psychology creates the very objects
that it investigates.
The search for method becomes one of the most important paradoxes of the
entire enterprise of understanding the uniquely human forms of psychological
activity. In this case, the method is simultaneously prerequisite and product,
the tool and the result of the study. (Vygotsky, 1978, p. 65)
Vygotsky challenges the view that the method of inquiry in psychology is separate
from the results of that inquiry, the traditional ‘tool for result’ position (Newman &
Holzman, 1993). Instead

27

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

As “simultaneously tool-and-result,” method is practiced, not applied.


Knowledge is not separate from the activity of practicing method; it is not
“out there” waiting to be discovered through the use of an already made tool.
… Practicing method creates the object of knowledge simultaneously with
creating the tool by which that knowledge might be known. Tool-and-result
come into existence together; their relationship is one of dialectical unity,
rather than instrumental duality. (Holzman, 1997, p. 52)
Tool-and-result means that no part of the practice can be removed and looked at
separately. Like the classic vase and faces optical illusion neither the faces nor
the vase can be removed and leave the other. There is no mathematical discipline
knowledge that can be removed from the way that it has been studied and looked
at separately. There is no content knowledge separate from pedagogic knowledge.
There is simply the practice of working together on the problem. This ‘practice
of method’ is entire in itself and ‘(t)he practice of method is, among other
things, the radical acceptance of there being nothing social (-cultural-historical)
independent of our creating it.’ (Newman & Holzman, 1997, p. 107)
The most important corollary of the ‘practice of method’ is, for Newman and
Holzman, the priority of creating and performing over cognition: “We are convinced
that it is the creating of unnatural objects – performances–which is required for
ongoing human development (developing)” (Newman & Holzman, 1997, p. 109). In
a similar vein, Adler (1998) refers to knowing being a dynamic and contextualized
process, in contrast to ‘knowledge’ as more static and abstract. The key distinction
here is that knowing is only manifested in practice and unfolds within a social
practice. The dynamics of this unfolding are different from those of ‘testing’ teachers’
subject knowledge in contexts that differ greatly from being in the classroom: one-
to-one interviews, multiple choice questions and so forth. We might question the
assumption that knowledge can be ‘retrieved’ outside the context in which it is used:
there is no retrieval, there is only practice.
Tool-and-result presents a double challenge to the research on teacher subject
knowledge: firstly, that the research itself constructs objects of knowledge, and,
second, that knowledge in classrooms emerges within ongoing discourse. Vygotsky’s
theory dissolves the notion that teachers ‘carry’ a store of mathematical knowledge
that they ‘apply’ in classrooms, and which mediates between the established cannon
of mathematical knowledge and the emergent mathematics of the classroom.
Davis and Simmt (2006) argue that distinctions constructed along the lines
of product and process are not helpful and that, from a complexity perspective,
the accepted cannon of codified mathematical knowledge and the emergent
mathematical understandings of learners are ‘self-similar’. This does not mean that
either is reducible to the other, but any difference between the two is more a matter
of scale than quality. Arguing that ‘learning’ needs to be extended beyond being
regarded as only happening in the heads of individuals, Davis and Simmt consider

28

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

mathematics as a learning system. Any distinction between discipline knowledge


and knowledge of how learners develop mathematical understanding is a matter
of timescale: mathematical discipline is only relatively stable in comparison to
mathematics in classrooms. It becomes impossible to separate subject knowledge
from pedagogic knowledge: “we must consider both teachers’ knowledge of
established mathematics and their knowledge of how mathematics is established
as inextricably intertwined” (Davis & Simmt, 2006, p. 300). This is not to argue
for a shift to a view of ‘learner-centred’ classrooms. As Davis and Simmt points
out, the question is not one of whether or not classrooms are learner-centred or
teacher-centred but whether or not they are mathematics centred. The argument
for primary school teachers to engage in mathematical activity may be more to do
with helping them better understand how mathematical knowledge is established
than with the substance of subject knowledge that they actually acquire. That could
entail encouraging teachers to develop a particular sensibility towards mathematics,
which we now turn to.

A FRAMEWORK FOR WORKING ON MATHEMATICAL


SENSIBILITY WITH TEACHERS

Rather than expecting primary school teachers to enter the profession knowing
all the subject knowledge that they might ever need to draw on, it may be more
reasonable to expect them to learn new aspects of the discipline, as and when they
need to. As Sullivan (2003) argues “so long as teachers have the orientation to learn
any necessary mathematics, and the appropriate foundations to do this, then prior
knowledge of particular aspects of content may not be critical” (p. 293). How then
might we encourage this sort of orientation? What sort of foundations might it need?
Possible key elements for both the initial and continuing professional development
of primary school teachers are usefully considered by adapting a framework
from Whitehead (1929, reprinted 1967). He argued that the curriculum should be
a process, a cycle of exploration and inquiry, a blend of “romance, precision and
generalization” (p. 17). Although Whitehead suggested that these three elements
should, to an extent, follow each other in that order, we argue that teachers would
need to develop them integratively; although for the sake of clarity here in the order
of precision, generalization and romance the integration of the three should be
kept in mind. In order to elaborate on the meaning of these three aspects, we take
multiplicative reasoning as a paradigm case.

Precision

The Oxford English Dictionary defines precision as “definite, exact, accurate and
free from vagueness.” This definition places accuracy as part of precision but not the
whole of it. Accuracy does, however, provide a useful starting point for considering
precision, as accuracy is arguably the part of the mathematics curriculum with which

29

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

most primary school teachers are most familiar. Indeed, for many, it is the main
purpose of the mathematics curriculum: getting right answers (and, unfortunately
for some learners, getting them right quickly). It is also part of the curriculum dear
to the hearts of policy makers. Any perceived attempts to diminish precision of the
mathematical curriculum are met with cries of lowering of standards and ‘fuzzy’
teaching. Hence the continuing debate in some parts of the world on whether or
not ‘standard algorithms’ should be taught and the announcement in England in
2018 of all 9-year-olds having to undertake a times tables test (https://www.gov.uk/
government/news/every-11-year-old-child-to-know-times-tables-by-heart).
But we argue that accuracy on its own is not enough: accuracy without insight is
limiting. As discussed earlier, as part of a large, longitudinal study, Millett, Askew,
and Simon (2004) interviewed 12 teachers about a number of mathematical tasks.
In the main, these teachers were able to find accurate answers to the tasks. But
their approaches to finding answers revealed a lack of insight into how and why
correct answers can come about, as a question asked about factors and divisibility
illustrates. Inscribed, at the time of the research, in the English national curriculum
for mathematics was the expectation that children were taught rules of divisibility.
To explore their subject knowledge of divisibility, the teachers in this research were
presented with a selection of numbers and asked which single digit numbers were
divisors of the number that was given (calculators were available). For example,
given 165, the teachers had to decide which single digit numbers were factors of it.
The teachers all were confident that 165 would not be divisible by 2 or 4 (“it’s odd”)
and that ending in 5 meant it was divisible by 5. They also said that knew there was a
rule for checking divisibility by 3, although some needed help to recall it. However,
whether or not 6 was a factor of 165 was not immediately apparent to them. Ten of
the twelve teachers needed to carry out dividing by 6 to check whether or not there
was a remainder. Asked whether knowing that 165 was not divisible by 2 could help
in deciding if 6 was a factor, they reasoned along the lines of “no, because the fact
that it was divisible by 3 might have made 6 a factor.”
The final number presented was 32 x 52 x 7. Given this, all 12 teachers multiplied
out the product to 1575 and then checked each digit in turn. When they had
established that 3, 5, 7 and 9 were factors they were asked whether, looking back
at the original expression, they might have been able to predict any of these results.
No, was the general response. None of the teachers appeared particularly surprised
at the result that, having started with a combination of 3s, 5s and 7, the only single-
digit factors that they found were 3, 5, 7 and 9. The sense of why they could not have
predicted this outcome was summed up by the teacher who said: “No, because you
never know that the way the 3s, 5s and 7 were multiplied together may have led to
it being divisible by, say, 2.” Over and above this, these teachers displayed a marked
lack of curiosity about the connection between the initial product and their answers.
This issue of being mathematically curious is one that we return to below.
But precision is more than accuracy. For instance, teachers need to be aware of
the importance of precise language in describing mathematical action. In a lesson

30

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

observed as part of this same research programme, a teacher was modelling division
with a class of 8- and 9-year-olds. She had a number of children standing holding
empty boxes and was modelling how many boxes would be filled if 7 cubes from
42 were put into each box; the model was of division as quotition (measurement).
However, throughout the modelling the teacher kept up a running commentary about
how the cubes were being ‘shared out’ amongst the boxes: the language of division
as partition. While problems like “How many bags are needed to put 42 apples into
bags of 7?” and “How many apples will be in each bag if 42 are shared between 7
bags?” can both ultimately be represented by 42 ÷ 6, children are likely initially to
solve these in different ways. Teachers need to appreciate the need for precision
and not muddling the two models through inappropriate marrying up of words and
actions.

Generalization

Without awareness of the move into the general, teaching and learning primary
school mathematics is unlikely to move beyond an emphasis on getting correct
answers, on precision. For example, consider exploring with children whether or
not this equation is true: 12 x 30 = 24 x 15. Working from a base of precision,
one way to answer this is to test it out; calculating the product on each side of the
equation will reveal the equation to be correct. Inspection of the numbers involved
may suggest a connection between them. This could be tested out with other similar
examples, arriving at a conjecture that doubling one number and halving the other
preserves the answer. This empirical approach to generalizing does not establish
why the conjecture holds true, or whether or not it will continue to hold. And it is
unlikely to provoke curiosity into whether or not this can be generalized further –
would tripling one number and ‘thirding’ the other also work? Trying out further
specific examples might mean a generalization is reached, but it would be a rule-
without-reason.
Consider, instead, modelling this with an array. A few simple diagrams, as shown
in Figure 1.1, with an open array quickly establishes the veracity of the equation and
opens the way to other partitionings of the array.
This analytic approach to generalizing (Schmittau, 2003) goes beyond the
‘specialize (through lots of examples) generalize model’ to seeing the ‘general-
in-the-specific’ (Mason & Pimm, 1984). It is a precise argument, but not based
on the precision of answers. The move to the general is a short one – the actual
dimensions of the rectangles are immaterial and the argument can, literally, be seen
to hold whatever dimensions are used to label the sides, and however many equally
sub-areas the original is divided up into. Generalizing is an adjunct to precision.
The use of the array goes beyond establishing the specific result. It opens up the
possibility of other constructions: do we only have to slice our rectangle into two
pieces? What about three slices? Or four? Curiosity is opened up and curiosity
leads to romance.

31

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

Figure 1.1. Diagrams with open array for equation

Romance

Mathematics has beauty and romance. It’s not a boring place to be, the
mathematical world. It’s an extraordinary place; it’s worth spending time there.
(Marcus du Sautoy, Professor of Mathematics, University of Oxford, speaking
at a conference)
Beauty may be considered a quality pertaining to the other – the quality of mathematics
that gives pleasure to the beholder. Or it may be in the eye of the beholder – how the
subject finds beauty in mathematics. Whichever, there is a ‘distance’ between subject
and object. Beauty can be admired from afar. Romance, however, implies intimacy,
a certain reciprocity, an entering in to a relationship with the other. Entering into a
romantic relationship with another person entails care and curiosity, qualities that
can equally be applied to romance with mathematics.

Romance and Care

Noddings (1992) questions whether it is too anthropomorphic to talk of caring for


an abstract discipline like mathematics. Her basic position is that of establishing
caring relationships, so a caring relationship with mathematics may indeed

32

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

be problematic and that “strictly speaking, one cannot form a relation with
mathematics” (p. 20). After all, how can mathematics care ‘back’? Yet, she argues,
we can talk meaningfully about caring for mathematics, that “oddly, people do
report a form of responsiveness from ideas and objects. The mathematician Gauss
was “seized” by mathematics” (p. 20), Bertrand Russell is said to have described
mathematics as his chief source of happiness and Littlewood said of Ramanujan
that every positive integer was one of his personal friends. Noddings goes on to
argue that
teachers should talk with students about the receptivity required in caring about
mathematics. People can become engrossed in mathematics, hear it “speak to
them,” be seized by its puzzles and challenges. It is tragic to deprive students
of this possibility. (1992, p. 152)
Initial or continuing education for teachers often engages them in mathematical
inquiry, but to what extent does such work engross, seize or challenge teachers
with mathematics? Davis and Simmt (2003) suggest that a common purpose
for engaging teachers in mathematical inquiry is predominantly pedagogical, so
that teachers are better placed to model to students what it means to engage with
mathematical problems and processes. This, they argue is problematic because
treating teaching as
a modelling activity seems to be rooted in the assumption of radical separations
among persons in the classroom. The teacher models, the learner mimics, but
their respective actions are seen to be separable and to spring from different
histories, interests and so on. (Simmt, Davis, Gordon, & Towers, 2003, pp.
178–179)
Often teachers engaged in mathematical inquiry talk less in terms of how the
experience will help them model for their students but more of re-entering the
experience of being a learner – the joys and frustrations of engaging in such work.
The emphasis is on the learner experience rather than the mathematics. Either
response – modelling or empathy – not only separates teachers and students, it
also carries the risk of teachers maintaining a distance from the mathematics. For
a caring relationship with the discipline to develop, the focus of mathematical
inquiry with primary school teachers needs to be on the mathematics, not just the
pedagogical.
Building on Noddings’ work, it is important to note the distinction between caring
for and caring about. It probably is unreasonable to expect all primary school teachers
to care for mathematics – people develop different appetites for different subjects.
But as teachers they do have a duty to care about mathematics: to recognize and
acknowledge the role that mathematics has played and continues to play in shaping
the world we live in. One step in promoting ‘caring about’ is the development of
curiosity.

33

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

Romance and Curiosity

When a National Curriculum was first established in England, funding was provided
for many primary school teachers to engage in 20 days of mathematical professional
development. Local funding for this was conditional on teachers working on their
discipline knowledge. As the 20 days released from school were spread over two
terms there was sufficient time to establish relationships with teachers that allowed
for working on aspects of mathematics that they might usually shy away from.
On one such course that one of the authors (Askew) conducted, the following
incident occurred towards the end when the teachers and tutor got to know each
other quite well. Askew recalls the episode:
Working on a particular inquiry I had given out some calculators. One of the
teachers, Ursula, called me over.
“My cheap calculator that I bought at a garage has got a 1/x button on, so why
doesn’t this expensive one?”
“Well, it does,” I replied “it’s that x-1 button.”
This seemed an opportunity to explore powers and so I stopped everyone to go
through an argument as to why x-1 is equated with 1/x.
While the teachers’ nods during my explanation suggested that they were
following my argument, afterwards there was a lot of muttering at Ursula’s
table.
“Is there anything you are not clear about?” I enquired.
“No, we follow your argument,” Ursula replied. “But we were just saying to
each other, ‘why would anyone ever want to do that in the first place?’”
This incident acts as a touchstone for several issues. First it highlights that
teachers’ emotional relationship with mathematics cannot be separated from their
intellectual, cognitive, knowledge of the subject. This was a group of teachers who
had begun to work with the mathematics intellectually – they were willing to ‘play
the game’ of developing mathematical ideas. But they were not yet engaged with
the game in the sense of deriving satisfaction from the pleasure of playing it, as
evidenced by the asking of why anyone would want to do that. Hodgen (2004)
highlights the importance of desire and imagination in developing and transforming
teachers’ relationships with mathematics.
We talk about engagement with mathematics as if the playing around with ideas
in and of itself is sufficient to make teachers and pupils want to carry on with the
play but as Simmt et al. (2003) put it, teachers have an ‘obligation’ to be curious
about mathematics. Teachers are encouraged to be curious about students’ responses
to mathematics, but this needs to be counterbalanced with a curiosity about
mathematics itself, as the following vignette illustrates.

34

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

Askew, visiting a class, was greeted by the teacher, who, in slightly exasperated
terms, asked: “Why, just when you think the children have got it, they can just
as easily lose it?” Asked to elaborate on what the ‘it’ was, the teacher talked of a
boy who could correctly say whether or not a number up to 8 was odd or even but
had answered that nine was even. It had not occurred to the teacher to ask why he
thought this (precision being most important presumably). When Askew asked the
boy why he thought nine was even, he demonstrated how nine tallies could be set
out into three ‘even’ (that is, same sized) sets. While part of being curious here is
about being interested in the pupil’s thinking, there is also the curiosity of wondering
if it is possible to construct a mathematics where nine could be considered even.
Where traditional teaching tends to focus on teaching of procedures, an orientation
underpinned by curiosity might consider instead ways of stretching, or extending,
procedures to encompass increasingly general sets of problems. Curiosity and
romance therefore feedback into generalisation. Such mathematical curiosity, when
present “compels teacher attendance to student articulations, it opens up closed
questions, and … can trigger similar contributions from learners” (Simmt et al.,
2003, p. 181). Simmt and her colleagues go on to argue that curiosity is not an
innate proclivity, but can be learned, to some extent at least and that this learning of
curiosity comes about through collective activity – a theme returned to in the final
section of this chapter.
One of the delicious aspects of entering into a new romantic relationship is finding
out about the history of the other. Being curious about the history of mathematics is
another aspect of subject knowledge that we would argue is necessary. Knowledge
of the history of mathematics can help teachers appreciate that there is no one single
story, no ‘truth’ of the way that mathematics has developed. For example, the story
of ‘Pascal’s’ triangle challenges the popularly held belief that mathematics is the
result of individual activity and inspiration. Although key theories are attributed to
individuals – Newton, Pascal, Pythagoras –these individuals were part of ongoing
communities, collectives. Not only did they stand on the shoulders of giants, they
rubbed shoulders with their peers. The development of mathematics is a collective
endeavour.
The story of the development of mathematics as one of the emergence and
invention of ideas, either to solve problems or simply through ‘playing’ with
mathematical objects, can challenge dominant views of the linearity of learning
mathematics. Such a perspective may help teachers to appreciate that there is no
‘truth’ about the way that students learn mathematics. As Davis and Sumara argue,
the distinction between established knowledge (the curriculum) and knowledge that
learners are developing or product versus process may not be the most appropriate
distinction. That the actual distinction may be more a matter of scale than quality:
it is the time span of development that is different rather than the substance of the
knowledge per se (Davis & Sumara, 2006). Appreciating the history of mathematics
problematizes the distinction between the established canonical body of knowledge
and the tentative knowledge of learners.

35

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

This points to the need to shift the discussion away from the individualistic
view of teachers, to a recognition that knowledge and learning is collective rather
than individual. We have to shift the attention away from the what of mathematics
subject knowledge that teachers ‘need’ to the why of what are they learning about
the way that they do mathematics. What does it mean to be part of a community of
mathematicians rather than an isolated acquirer of mathematical knowledge? Before
examining this question, we look at some of the theoretical and epistemological
stances that might underpin much of the work in this area to date. Through making
these explicit, directions for future work begin to emerge.

FROM THE INDIVIDUAL TO THE COLLECTIVE

Finally, we examine whether the enterprise of looking at/for prospective primary


school teachers’ subject knowledge needs to move in different directions. To
move away from the individualistic/cognitive focus that dominates so much of the
literature towards a collective/situated perspective. That classrooms are constituted
as they are with a single teacher to a large group of students is an historically
contingent norm rather than a necessary condition for the organisation of teaching
and learning. As such, the location of the ‘problem’ of subject knowledge within the
heads of individual teachers is a result of this contingency. Rather than continuing
to seek to solve the problem by working on what is ‘in’ teachers heads perhaps
a more productive approach may be to look to how schooling is constituted, to
accept that the knowledge of mathematics for teaching, particularly in primary
schools, is likely to be distributed across a group of teachers and to work with this.
Although studies of classroom communities have begun to take social, situated
and collective perspectives (Lerman, 2000), studies of teachers’ subject knowledge
remain, by and large, within the paradigm of individual cognition. Even those
studies that have begun to look at the mathematics knowledge that teachers actually
use in classrooms are still driven by a desire to ‘extract’ such knowledge from its
history and context and codify and canonize it. Adler (1998) argues that learning to
become a mathematics teacher means becoming able to talk within the discourses
of mathematics teaching and to talk about such discourses. Thus, there is more
than simply learning new knowledge and such talk needs to be developed through
collective engagement.
Much of the research into subject knowledge is predicated (tacitly) on the
assumption that it is necessary for one individual (the teacher) to be able to make
sense of another individual’s (the pupil) cognitions. With a shift to recognizing the
importance of the social in developing individual understandings, this ‘head-to-head’
model is only part of the picture. Perhaps even the focus on individual pupils needs
to be questioned: “The ‘learning system’ that the teacher can most directly influence
is not the individual student, but the classroom collective” (Davis & Simmt, 2003,
p. 164).

36

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

If the classroom collective rather than the individual student should be the focus
of the teacher, then should the teaching collective rather than the individual teacher
be the focus of the researcher? This is as much a political question as a theoretical
one. It is notable that teacher development in places like the United States of
America and the United Kingdom focuses on the development of the individual, in
line with the dominant political culture of ‘self-actualization’. Other traditions of
development, for example lesson study in Japan (Lewis, 2002) already attend more
to the collective than the individual.
Attending to the distributed nature of discipline knowledge means paying greater
attention to the communities within which teachers are located. Millet, Brown, and
Askew (2004), examine the importance of the professional community of teachers
within schools. Some of the schools tracked over five years in the course of the
Leverhulme Numeracy Research Programme were able successfully to ‘share’
mathematical expertise through setting up teams of teachers with collective
responsibility for mathematics across the school.
In research studying the effectiveness of 90 primary school teachers in teaching
mathematics, Askew and colleagues (Askew et al., 1997b) identified one school
where the pupil gains over the course of a year on an assessment of numeracy were
consistently high across all classes. This was despite the fact that not all teachers in
the school demonstrated particularly strong discipline knowledge. One factor that
conjectured as contributing to these consistent gains was the strength of support
provided by two teachers who shared the responsibility for mathematics across
the school and who had complementary strengths. One had a strong mathematical
background gained through studying a science degree. The other had studied the
psychology and pedagogy of primary mathematics over several years of ongoing
involvement in professional development.
One claim arising from the research into teacher knowledge is that lack of subject
knowledge leads to over-reliance on textbooks in the classroom. It may be that rather
than lack of subject knowledge, use of textbooks is the result of teachers seeking
‘surrogate’ classroom partners. In the absence of another adult being physically
present in the room, the voice of the textbook may provide the next best thing. The
issue may not be one of helping individual teachers become better ‘equipped’ to
scale the peaks of mathematics lessons but acknowledging their need for fellow
climbers.
This is not to deny the place of the individual, but to recognize that individual
cognition is part of a wider network. As Davis and Simmt express it:
mathematical knowing is rooted in our biological structure, framed by bodily
experiences, elaborated within social interactions, enabled by cultural tools, and
part of an ever-unfolding conversation of humans and the biosphere. (p. 315)
It seems there is plenty to keep researchers in this field going for many more
years yet.

37

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

REFERENCES
Adler, J. (1998). Lights and limits: Decontextualizing Lave and Wenger to theorize knowledge of teaching
and of learning school mathematics. In A. Watson (Ed.), Situated cognition and the learning of
mathematics. Oxford: Centre for Mathematics Education Research, University of Oxford, Department
of Educational Studies.
Askew, M. (1999). Teachers, orientations and contexts: Repertoires of discourse in primary mathematics
(Unpublished doctoral thesis). London: King’s College, University of London.
Askew, M., & Brown, M. (1997). Effective teachers of numeracy in UK primary schools: Teachers’
beliefs, practices and pupils’ learning. Paper presented at the European Conference on Educational
Research (ECER 97), Johann Wolfgang Goethe Universitat, Frankfurt, Germany.
Askew, M., Brown, M., Rhodes, V., Wiliam, D., & Johnson, D. (1997a). Effective teachers of numeracy:
Report of a study carried out for the Teacher Training Agency. London: King’s College, University
of London.
Askew, M., Brown, M., Rhodes, V., Wiliam, D., & Johnson, D. (1997b). The contribution of professional
development to effectiveness in the teaching of numeracy. Teacher Development, 1(3), 335–355.
Aubrey, C. (1997). Mathematics teaching in the early years: An investigation of teachers’ subject
knowledge. London: Falmer.
Ball, D. (1990). The mathematical understandings that prospective teachers bring to teacher education.
The Elementary School Journal, 90(4), 449–466.
Ball, D. (1991a). Teaching mathematics for understanding: What do teachers need to know about subject
matter? In M. Kennedy (Ed.), Teaching academic subjects to diverse learners. New York, NY:
Teachers College Press.
Ball, D. L. (1991b). Research on teaching mathematics: Making subject matter part of the equation. In
J. Brophy (Ed.), Advances in research on teaching. Teachers’ knowledge of subject matter as it relates
to their teaching practice (Vol. 2, pp. 1–48). Greenwich: JAI press.
Ball, D. L., & Bass, H. (2000). Interweaving content and pedagogy in teaching and learning to teach:
Knowing and using mathematics. In J. Boaler (Ed.), Multiple perspectives on mathematics teaching
and learning (pp. 83–104). Westport, CT: Ablex Publishing.
Ball, D. L., & Bass, H. (2003). Toward a practice-based theory of mathematical knowledge for teaching.
In B. Davis & E. Simmt (Eds.), Proceedings of the 2002 Annual Meeting of the Canadian Mathematics
Education Study Group (pp. 3–14). Edmonton: CMESG/GCEDM.
Ball, D. L., Hill, H. C., & Bass, H. (2005). Knowing mathematics for teaching: Who knows mathematics
well enough to teach third grade, and how can we decide? American Educator, 29(1), 14–46.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Barwell, R. (2013). Discursive psychology as an alternative perspective on mathematics teacher
knowledge. ZDM-The International Journal on Mathematics Education, 45(4), 595–606.
Begle, E. G. (1979). Critical variables in mathematics education: Findings from a survey of the empirical
literature. Washington, DC: Mathematical Association of America and National Council of Teachers
of Mathematics.
Bennett, N., & Carre, C. (Eds.). (1993). Learning to teach. London: Routledge.
Bennett, N., & Turner-Bisset, R. (1993). Case studies in learning to teach. In N. Bennett & C. Carre
(Eds.), Learning to teach. London: Routledge.
Brown, M., Askew, M., Baker, D., Denvir, H., & Millett, A. (1998). Is the national numeracy strategy
research-based? British Journal of Educational Studies, 46(4), 362–385.
Charalambous, C. (2010). Mathematical knowledge for teaching and task unfolding: An exploratory
study. The Elementary School Journal, 110(3), 247–278.
Conference Board of the Mathematical Sciences. (2001). The mathematical education of teachers.
Washington, DC: American Mathematical Society and Mathematical Association of America.
Connolly, F. M., Clandinin, D. J., & He, M. F. (1997). Teachers’ personal practical knowledge on the
professional knowledge landscape. Teaching and Teacher Education, 13(7), 665–674.

38

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

Darling-Hammond, L. (2000). Teacher quality and student achievement: A review of state policy
evidence. Education Policy Analysis Archives, 8(1), 1–44.
Davis, B., & Simmt, E. (2003). Understanding learning systems: Mathematics education and complexity
science. Journal for Research in Mathematics Education, 34(2), 137–167.
Davis, B., & Simmt, E. (2006). Mathematics-for-teaching: An ongoing investigation of the mathematics
that teachers (need to) know. Educational Studies in Mathematics, 61(3), 293–319.
Davis, B., & Sumara, D. (2006). Complexity and education: Inquiries into learning, teaching and
research. Mahwah, NJ: Lawrence Erlbaum Associates.
DBE. (2011). Curriculum and Assessment Policy Statement (CAPS): Foundation phase mathematics,
grades R-3. Pretoria: Department for Basic Education.
Dewey, J. (1964). The relation of theory to practice in education. In R. Archambault (Ed.), John Dewey
on education (pp. 313–338). Philadelphia, PA: University of Pennsylvania. (Original work published
in 1904)
Ernest, P. (1998). Social constructivism as a philosophy of mathematics. New York, NY: State University
of New York Press.
Freudenthal, H. (1975). Mathematics as an educational task. Dordrecht: Reidel.
Grossman, P. L., Wilson, S. M., & Shulman, L. E. (1989). Teachers of substance: Subject matter
knowledge for teaching. In M. C. Reynolds (Ed.), Knowledge base for the beginning teacher. New
York, NY: Pergamon.
Hattie, J. (2012). Visible learning for teachers: Maximizing impact on learning. Oxon: Routledge.
Hill, H. C., Blunk, M. L., Charalambous, C. Y., Lewis, J. M., Phelps, G. C., Sleep, L., & Ball, D. L. (2008).
Mathematical knowledge for teaching and the mathematical quality of instruction: An exploratory
study. Cognition and Instruction, 26(4), 430–511.
Hill, H. C., Umland, K. L., Litke, E., & Kapitula, L. (2012). Teacher quality and quality teaching:
Examining the relationship of a teacher assessment to practice. American Journal of Education, 118,
489–519.
Hodgen, J. (2004). Identity, motivation and teacher change in primary mathematics: A desire to be a
mathematics teacher. Proceedings of the British Society for Research into Learning Mathematics,
24(1), 31–36.
Holzman, L. (1997). Schools for growth: Radical alternatives to current educational models. Mahwah,
NJ: Lawrence Erlbaum Associates.
Hoover, M., Mosvold, R., Ball, D. L., & Lai, Y. (2016). Making progress on mathematical knowledge for
teaching. The Mathematics Enthusiast, 13(1–2), 3–34.
Hough, S., O’Rode, N., Terman, N., & Weissglass, J. (2007). Using concept maps to assess change
in teachers’ understandings of algebra: A respectful approach. Journal of Mathematics Teacher
Education, 10(1), 23–41.
Lampert, M. (1986). Knowing, doing, and teaching multiplication. Cognition and Instruction, 3(4), 305–342.
Leinhardt, G., & Smith, D. A. (1985). Expertise in mathematics instruction: Subject matter knowledge.
Journal of Educational Psychology, 77(3), 247–271.
Lerman, S. (2000). The social turn in mathematics education research. In J. Boaler (Ed.), Multiple
perspectives on mathematics teaching and learning (pp. 19–44). Westport, CT: Ablex Publishing.
Lewis, C. C. (2002). Lesson study: A handbook of teacher-led instructional change. Philadelphia, PA:
Research for Better Schools Inc.
Ma, L. (1999). Knowing and teaching elementary mathematics: Teachers’ understanding of fundamental
mathematics in China and the United States. Mahwah, NJ: Lawrence Erlbaum Associates.
Mason, J., & Pimm, D. (1984). Generic examples: Seeing the general in the particular. Educational
Studies in Mathematics, 15(3), 277–289.
McNamara, D. (1991). Subject knowledge and its application: Problems and possibilities for teacher
educators. Journal of Education for Teaching, 17(2), 113–128.
McNamara, O., Jaworski, B., Rowland, T., Hodgen, J., & Prestage, S. (2002). Developing mathematics
teaching and teachers: A research monograph. Unknown Publisher. Retrieved January 15, 2008, from
http://www.maths-ed.org.uk/mathsteachdev/

39

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MIKE ASKEW & HAMSA VENKAT

Millett, A., Askew, M., & Simon, S. (2004). Reponses of teachers to a course of intensive training. In A.
Millett, M. Brown, & M. Askew (Eds.), Primary mathematics and the developing professional (Vol.
1, pp. 127–154). Dordrecht: Kluwer Academic Publishers.
Millett, A., Brown, M., & Askew, M. (Eds.). (2004). Primary mathematics and the developing professional
(Vol. 1). Dordrecht: Kluwer Academic Publishers.
Monk, D. H. (1994). Subject area preparation of secondary mathematics and science teachers and student
achievement. Economics of Education Review, 13(2), 125–145.
Mosvold, R., & Hoover, M. (2017). Mathematical knowledge for teaching and the teaching of
mathematics. In T. Dooley & G. Gueudet (Eds.), Proceedings of the tenth congress of the European
Society for Research in Mathematics Education (pp. 3105–3112). Dublin: DCU Institute of Education
and ERME.
National Research Council. (2001). Adding it up: Helping children learn mathematics. Washington, DC:
National Academy Press.
Newman, F., & Holzman, L. (1993). Lev Vygotsky: Revolutionary scientist. London: Routledge.
Noddings, N. (1992). The challenge to care in schools. New York, NY: Teachers College Press.
Novak, J. D., & Gowin, D. B. (1984). Learning how to learn. New York, NY: Cambridge University
Press.
Poulson, L. (2001). Paradigm lost? Subject knowledge, primary teachers and education policy. British
Journal of Educational Studies, 49(1), 40–55.
Rad, S. R., Martingano, A. J., & Ginges, J. (2018). Toward a psychology of Homo sapiens: Making
psychological science more representative of the human population. Proceedings of the National
Academy of Sciences, 115(45), 11401–11405.
Rowland, T., Barber, P., Heal, C., & Martyn, S. (2003). Prospective primary teachers’ mathematics
subject knowledge: Substance and consequence. Proceedings of the British Society for Research into
Learning Mathematics, 23(3), 91–96.
Rowland, T., Martyn, S., Barber, P., & Heal, C. (2000). Primary teacher trainees’ mathematics subject
knowledge and classroom performance. In T Rowland & C Morgan (Eds.), Research in mathematics
education: Papers of the British Society for Research into Learning Mathematics (Vol. 2, pp. 3–18).
London: British Society for Research into Learning Mathematics.
Schmittau, J. (2003). Cultural-historical theory and mathematics education. In A. Kozulin, B. Gindis, V.
S. Ageyev, & S. M. Millar (Eds.), Vygotsky’s educational theory in cultural context (pp. 225–245).
Cambridge: Cambridge University Press.
Shulman, L. S. (1986). Those who understand: Knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Shulman, L. S. (1987). Knowledge and teaching: foundations of the new reforms. Harvard Educational
Review, 57(1), 1–22.
Simmt, E., Davis, B., Gordon, L., & Towers, J. (2003). Teachers’ mathematics: Curious obligations. In
N. A. Pateman, B. J. Dougherty & J. T. Zilliox (Eds.), Proceedings of the 2003 joint meeting of PME
and PME-NA (pp. 175–182). Honolulu, HI: Centre for Research and Development Group, University
of Hawaii.
Simon, S., & Brown, M. (1996). Teacher beliefs and practices in primary mathematics. In L. Puig & Á.
Gutiérrez (Eds.), Proceedings of the 20th Conference of the International group for the Psychology of
Mathematics Education (p. 200). Valencia: PME.
Skemp, R. (1976). Relational understanding and instrumental understanding. Mathematics Teaching, 77,
20–26.
Stacey, K., Helme, S., Steinle, V., Baturo, A., Irwin, K., & Bana, J. (2003). Preservice teachers’ knowledge
of difficulties in decimal numeration. Journal of Mathematics Teacher Education, 4(3), 205–225.
Sullivan, P. (2003). Editorial: Incorporating knowledge of, and beliefs about, mathematics into teacher
education. Journal of Mathematics Teacher Education, 6(4), 293–296.
Sullivan, P., Clarke, D., & Clarke, B. (2009). Converting mathematics tasks to learning opportunities: An
important aspect of knowledge for mathematics teaching. Mathematics Education Research Journal,
21(1), 85–105.

40

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL SUBJECT KNOWLEDGE

Thompson, A. G., Philipp, R. A., Thompson, P. W., & Boyd, B. A. (1994). Calculational and conceptual
orientations in teaching mathematics. In D. B. Aichele & A. F. Coxford (Eds.), Professional
development for teachers of mathematics. 1994 yearbook. Reston, VA: National Council of Teachers
of Mathematics.
Vygotsky, L. S. (1978). Mind in society. Cambridge, MA: Harvard University Press.
Walkerdine, V. (1988). The mastery of reason: Cognitive development and the production of rationality.
London: Routledge.
Whitehead, A. N. (1929, reprinted 1967). The aims of education and other essays. New York, NY: The
Free Press.
Wilson, S. M., Floden, R. E., & Ferrini-Mundy, J. (2001). Teacher preparation research: Current
knowledge, gaps, and recommendations. An executive summary of the research report. Seattle, WA:
University of Washington, Center for the Study of Teaching and Policy.
Wilson, S. M., Shulman, L. S., & Richert, A. E. (1987). ‘150 different ways’ of knowing: Representations
of knowledge in teaching. In J. Calderhead (Ed.), Exploring teachers’ thinking (pp. 84–103). London:
Cassell.
Wittgenstein, L. (1953). Philosophical investigations (G. E. M. Anscombe, Trans.). Oxford: Blackwell.
Wittrock, M. (Ed.). (1986). Handbook of research on teaching. New York, NY: Macmillan.
Wragg, E. C., Bennett, S. N., & Carre, C. G. (1989). Primary teachers and the National Curriculum.
Research Papers in Education, 4(3), 17–45.

Mike Askew
Wits School of Education
University of the Witwatersrand

Hamsa Venkat
Wits School of Education
University of the Witwatersrand

41

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH AND OLIVE CHAPMAN

2. BUILDING TEACHERS’ CAPACITY IN


MATHEMATICS AUTHENTIC ASSESSMENT

Given the central role of tasks in the teaching and learning of mathematics,
supporting teachers’ development of mathematics task knowledge for teaching is
of great importance. In this chapter, we address mathematical tasks from a reform-
oriented perspective of mathematics education with a focus on authentic assessment.
We present a sample of the large body of literature on the nature and role of tasks
in teaching and learning mathematics, frameworks to determine the quality of
mathematics tasks and teachers’ knowledge of mathematics tasks. We discuss
authentic assessment and authentic assessment tasks in teaching mathematics
and draw on our research involving the use of an ‘authentic intellectual quality
framework’ to provide examples of promising ways to support prospective and
practising teachers’ development of capacity in the design, selection, and use of
mathematical authentic assessment tasks.

INTRODUCTION

Mathematical tasks are central to learning and teaching mathematics. Thus, they have
received extensive attention in the research literature. This includes consideration of
their nature/attributes and role in teaching and learning mathematics (e.g., Boesen,
Lithner, & Palm, 2010; Hsu & Silver, 2014; Henningsen & Stein, 1997; Mason, 2016;
Mason & Johnston-Wilder, 2006; National Council of Teachers of Mathematics
[NCTM], 1991, 2000; Shimizu, Kaur, Huang, & Clarke, 2010; Sullivan, Clarke, &
Clarke, 2013; Swan, 2008; Watson & Ohtani, 2015) and consideration of their use in
mathematics teacher education and professional development (e.g., Koh, 2014; Koh
& Chapman, 2018; Lee, 2017; Leikin & Levav-Waynberg, 2009; Levenson, 2015;
Pepin & Jones, 2016; Swan, 2007; Watson & Mason, 2007; Watson & Sullivan,
2008; Wilhelm, 2014; Zaslavsky & Sullivan, 2010). In this chapter, we address
mathematical tasks from a reform-oriented perspective of mathematics education that
focuses on helping students to learn important mathematics with deep understanding
and develop mathematical thinking/competencies/processes (e.g., problem solving,
communication, reasoning, connections, and representation; NCTM, 2000). Our
intent is not to discuss the large body of research literature on tasks but to draw
on specific aspects being linked to mathematics teacher knowledge and education.
We first consider the nature and role of tasks in teaching and learning mathematics,

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_003

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

followed by frameworks of quality of mathematics tasks and teachers’ knowledge of


mathematics tasks. We then focus on our work on mathematics authentic assessment
and authentic assessment tasks and teachers’ development of capacity in the design,
selection, and use of mathematical authentic tasks for the purposes of both learning
and assessment in grades K–12 classrooms. We end the chapter by making some
recommendations for initial teacher preparation and practising teacher professional
learning in mathematics authentic assessment.

TASKS IN THE TEACHING AND LEARNING OF MATHEMATICS

The Nature and Role/Purpose of Mathematical Tasks

A mathematical task could be a set of problems or a single complex problem


that focuses students’ attention on a particular mathematical idea (Stein et al.,
1996) and/or a way of thinking (Boaler, 2016; Mason, Burton, & Stacey, 2010;
Schoenfeld, 1994). Mathematical tasks have been labelled in a variety of ways
in the literature to represent tasks that offer experiences to develop students’
mathematical understandings and mathematical thinking; for example: worthwhile
(NCTM, 1991, 2010), authentic (Koh & Chapman, 2018), rich (Flewelling &
Higginson, 2001), inquiry-based (Jaworski, 2015), dynamic mathematical
activity (Henningsen & Stein, 1997), high cognitive demand (Stein, Grover,
& Henningsen, 1996), competence oriented (Pettersen & Nortvedt, 2018); and
learner-focused (Chapman, 2013a). We consider all of these to be “worthwhile”
tasks. The following ten criteria are outlined by NCTM (2010) to define
worthwhile tasks: (1) focuses on important mathematical ideas; (2) provides
students with the opportunity to engage in higher-order thinking and problem
solving; (3) contributes to students’ development of conceptual understanding
of mathematics; (4) enables the teacher to assess students’ learning, identify
misconceptions, and diagnose individual areas of difficulty in order to improve
instruction and student learning; (5) should allow students to solve the problem
using multiple solution strategies; (6) has more than one solution and supports
students’ use of alternative strategies and perspectives to defend their solutions;
(7) promotes students’ engagement and discourse; (8) enables students to make
meaningful connections to other important mathematical ideas; (9) creates
opportunities for students to apply mathematical knowledge in a skilful way;
and (10) supports students to practice important mathematical skills. However,
NCTM points out that not every task or problem is expected to meet all ten
criteria. Rather, it depends on the teacher’s instructional goals.
As discussed in Chapman (2013a), the purposes of mathematical tasks could
impact their attributes and determine their uses in the classroom, that is, for
instruction, learning, review, practice, and assessment. Other influences on the
nature and effectiveness of the tasks includes: the teacher, learner, curriculum and
social context/expectations. Thus, all of these aspects of the task landscape play

44

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

an important role in the full nature of the task, that is, the form of the task, the
category/attributes, the purpose, and classroom influences. Collectively, they are
necessary to understand tasks not as lifeless objects but as lived experiences in the
classroom shaped by the teacher, students, and social context of the classroom. For
example, Watson and Mason (2007) identified influences of the effectiveness of a
task to include ethos and atmosphere; established practices and ways of working;
students’ expectations of themselves and of each other, influenced by the system
and their pasts; and learners’ sense of self-confidence, agency (mathematically and
socially) and identity. Shimizu et al. (2010) also explained that the social context
could take many forms that leads to different enactment of the task including:
an exploratory instructional activity in small collaborative groups; a whole class
discussion, orchestrated by the teacher to draw out existing student understandings;
or, an assessment task to be undertaken individually.
While these factors influence the role/purpose of a mathematics task, the literature
suggests specific roles/purposes of worthwhile tasks. For example, they “have the
potential to provide intellectual challenges for enhancing students’ mathematical
understanding and development” and “direct students to investigate important
mathematical ideas and ways of thinking toward the learning goals” (NCTM, 2010,
p. 1). They play an important role to develop students’ capacity for mathematical
thinking and reasoning (Boaler, 2016; Mason et al., 2010; Stein et al., 1996). They
“focus students’ attention on a particular mathematical idea” (Stein et al., 1996, p.
460). They have “implications for ideas about what students need to learn and the
kinds of activities in which students and teachers should engage during classroom
interactions” (Henningsen & Stein, 1997, p. 525).
In general, mathematics tasks that are “worthwhile” focus on promoting students’
conceptual understanding, thinking and reasoning ability, and communication
skills and a conative dimension that is about enhancing students’ interests and
curiosity in mathematics. They are truly complex and intellectually challenging,
which can be meaningful and intriguing to students so as to promote their interests
in and persistence on tasks. Thus, the nature and role of mathematics tasks are
multidimensional and complex that add to the challenge for teachers to develop
appropriate task-knowledge for teaching without meaningful support and learning
experiences to do so.

Authentic Assessment Tasks


Since our work focuses specifically on authentic assessment tasks, we address them
separately in this section to present our perspective of them. As discussed above,
‘worthwhile tasks’ are important to promote meaningful learning of mathematics,
however, they can also be used to meaningfully assess that learning. Authentic tasks
have unique features that make them relevant and important not only to support
meaningful learning but also to achieve meaningful assessment of students’ learning
of mathematics from a reform perspective.

45

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

Reform perspectives of school mathematics require students to make meaningful


connections among mathematical concepts, between mathematical concepts and
procedures, and between mathematics and other disciplines and the ‘real-world’; to
solve worthwhile problems; and to demonstrate mathematical thinking, reasoning,
and communication. Such “mathematical power” or mathematical proficiency
can be best captured by well-designed authentic assessment tasks that are of high
intellectual demands (Koh & Chapman, 2018; Newmann, Marks, & Gamoran, 1996;
Romberg, 1995).
Authentic assessment tasks replicate real-world challenges and standards of
performance that experts or professionals typically face in the field (Koh, 2017;
Wiggins, 1989). According to Wiggins, they are designed to be truly representative
of performance in the field, that is, they replicate the intellectual challenges
and performance standards typically faced by professionals in the field (e.g.,
mathematicians, scientists, writers, designers). As such, the tasks must include
contextualized, complex intellectual challenges involving students’ application of
knowledge in messy, ill-structured contexts. Specific to mathematics, Lesh and
Lamon (1992) define mathematics authentic assessments as authentic mathematical
activities that involve the following elements: real mathematics, realistic situations,
questions or issues that might actually occur in a real-life situation and use of
realistic tools and resources. In addition to real-world challenges, Lajoie (1995) has
pointed out that authentic tasks need to give students the opportunities to “reflect,
organize, model, represent, and argue within and across mathematical domains”
(p. 30). Thus, in attempting mathematics authentic tasks, students are required to think
and act like real mathematicians who are doing mathematics or like professionals who
use mathematical concepts and procedures to solve real-world problems. Table 2.1
provides an example of a mathematics authentic assessment task.
Authentic assessment tasks fall in the category of “good assessment tasks,”
which, according to Shepard (2000) are open-ended performance tasks that require
students to solve complex problems, to think and reason critically, and to apply their
knowledge in real-world contexts. But they can also include instructional goals that
help students develop conative skills (e.g., interest in and persistence on tasks, self-
efficacy, self-directed learning) and socialize them into the discourse and practices
of academic disciplines. For example, non-algorithmic problems embedded in a
mathematics authentic task not only help students think and reason about an important
mathematical idea (e.g., visualizing part-whole relationships in fractions), but also
develop their confidence in mathematical argumentation using appropriate symbols,
representations, and language. Thus, while we focus on mathematics authentic
tasks in relation to assessment in this chapter, we acknowledge that they are equally
valuable for use in instruction to support students’ learning. As Shepard (2000)
notes, “Good assessment tasks are interchangeable with good instructional tasks”
(p. 8). This double role is also connected to the relationship between assessment
and instruction, that is, assessment should be an integral part of instruction. This
indicates a formative assessment or assessment for learning perspective in which

46

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Table 2.1. An example of mathematics authentic assessment task


(adapted from Lim, 2011)

In the 2016 Summer Olympics, the top three countries are USA, UK, and China. The
medal table below presents each type of medal won by each of the countries. As a sports
reporter for the Calgary Herald, you are asked by the chief editor to write a report based
on your analysis of the medals. In your report, you need to include three different sources
of information: mean, median, and mode, as well as to provide your explanation for each
of them. You also have to decide on the most appropriate average type for reporting your
analysis of the medals. Your report will need to be in a format which is suitable for the
sports column of the Calgary Herald. Prior to submitting your report to the newspaper, you
will present your work to your peers to receive feedback.

Top 3 Countries Gold Silver Bronze

United States (USA) 46 37 38


United Kingdom 27 23 17
China 26 18 26

Guidelines for students:

‡ Your written report should include evidence for each of the average types (mean,
median, and mode).
‡ Calculate the mean, median, and mode for all Gold, Silver and Bronze medals.
‡ When do you use or not use mean, median, and mode? Explain your thinking.
‡ Justify the most appropriate average type in your report.
‡ Your work will be graded based on the rubric discussed during the first lesson.

assessment tasks or activities are used primarily to obtain information to improve


instruction and student learning. This perspective allows for assessment to be
integrated into instructional activities to promote student learning (Shepard et al.,
2005). Thus, as Shepard indicates, if properly designed, authentic assessment tasks
can serve as classroom instructional activities to promote teachers’ and students’
engagement in formative assessment or assessment for learning during the teaching
and learning process prior to summative assessment of student learning.
To illustrate the relationship between authentic assessment task and formative
assessment, we consider Black and Wiliam’s (1998) definition of formative
assessment “as encompassing all those activities undertaken by teachers, and/
or their students, which provide information to be used as feedback to modify the
teaching and learning activities in which they are engaged” (p. 7). They offer five
key formative assessment strategies: explicit sharing of learning goals, performance
standards, and success criteria; effective questioning; descriptive feedback; self-
assessment; and peer assessment (Black & Wiliam, 2007; Wiliam, 2007). These

47

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

strategies can be incorporated into the use of mathematics authentic assessment task.
For example, considering the task in Table 2.1, a mathematics teacher can share
the identified learning goals (i.e., understanding of the concepts of mean, median,
and mode) with his or her students during the first lesson. Students are then invited
to engage in establishing or negotiating a set of success criteria and performance
standards (what count as evidence of understanding) with the teacher for developing
an analytic rubric, which will later be used for teacher, self, and peer assessments.
In the process of attempting the task, the teacher uses the rubric to provide timely
descriptive feedback that aims to move students’ learning forward and to engage
students in monitoring and regulating their own learning (self-assessment), as well
as to promote collaboration among students by asking them to read each other’s draft
report and to provide constructive feedback (peer assessment).The teacher can also
ask open-ended, rich questions to elicit evidence of students’ progress toward the
learning goals.
We thus consider mathematics authentic assessment tasks as having the potential
of being worthwhile mathematics tasks and formative assessment tasks or being
about authentic assessment depending on how they are used by the teacher.
While this section focused on the nature of authentic tasks, in a later section we
focus on the nature of authentic assessment, which frames our work. But first, we
discuss ways of considering the quality of mathematics tasks and of helping teachers
to learn from them.

FRAMEWORKS OF QUALITY OF TASKS

Researchers have developed frameworks to study the quality of mathematical


tasks, teachers’ knowledge of mathematical tasks, and ways to support teachers’
development of mathematics task knowledge to use in the teaching. Through these
frameworks, teachers could learn how to recognize and select appropriate tasks to
be used in their teaching to support students’ development of deep mathematical
knowledge and mathematical competence and to assess this knowledge and
competence. Three of these frameworks are presented here as examples. The
framework of levels of competencies identifies different competencies associated
with learning or doing mathematics, which are used as a basis to analyze the level
of demand for each competency for a task. In contrast, the framework of cognitive
demands identifies characteristics of a task associated with the levels of thinking or
challenge to obtain a solution. The authentic intellectual quality framework contains
some aspects of both of these in addressing cognitive demands of a task and some
key competencies. We next elaborate on each of these frameworks.

Framework of Level of Competencies

Pettersen and Braeken (2019) cites Niss and Højgaard (2011, p. 49) as describing
mathematical competence as “having the knowledge of, understanding, doing, using

48

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

and having an opinion about mathematics and mathematical activity in a variety


of contexts where mathematics plays or can play a role.” Growing emphasis on
mathematical competence in research on mathematics education has resulted in
competency frameworks which Kilpatrick (2014, p. 85) defined as a “structural
plan for organizing the cognitive skills and abilities used in learning and doing
mathematics.” Examples of these competency frameworks are the Danish KOM
framework and the related framework of the Program for International Student
Assessment (PISA) mathematics expert group (Niss, 2015). The KOM framework
consists of eight mathematical competencies regarding what it means to master
mathematics that overarches mathematical topics and content areas (Kilpatrick, 2014).
The PISA mathematics expert group framework, which is based on the perspective
of mathematical competence underlying the PISA mathematics frameworks,
consists of six mathematical competencies (see Table 2.2, as presented in Pettersen
and Braeken). Pettersen and Braeken explained that this framework is “a modified
version of the competencies in the KOM framework where the mathematical thinking
competency and the reasoning competency have been merged into Reasoning and
argument, and the Aids and tools competency has been omitted (Turner et al., 2015)”
(p. 407). These six competencies are the basis of a task analysis scheme connected
to the PISA mathematics expert group framework that “includes descriptions of four

Table 2.2. Definitions of the six mathematical competencies in the item analysis scheme
(Turner et al., 2015) [as cited in Pettersen & Braeken, 2019, p. 408]

Communication. Reading and interpreting statements, questions, instructions, tasks,


images and objects; imagining and understanding the situation presented and making
sense of the information provided including the mathematical terms referred to; presenting
and explaining one’s mathematical work or reasoning.
Devising strategies. Selecting or devising a mathematical strategy to solve a problem as
well as monitoring and controlling implementation of the strategy.
Mathematising. Translating an extra-mathematical situation into a mathematical model,
interpreting outcomes from using a model in relation to the problem situation or validating
the adequacy of the model in relation to the problem situation.
Representation. Decoding, translating between and making use of given mathematical
representations in pursuit of a solution; selecting or devising representations to capture the
situation or to present one’s work.
Symbols and formalism. Understanding and implementing mathematical procedures and
language (including symbolic expressions, arithmetic and algebraic operations), using the
mathematical conventions and rules that govern them; activating and using knowledge of
definitions, results, rules and formal systems.
Reasoning and argument. Drawing inferences by using logically rooted thought processes
that explore and connect problem elements to form, scrutinise or justify arguments and
conclusions.

49

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

different levels of demand for each competency ranging from 0 (lowest demand) to
3 (highest demand) (for the full item analysis scheme, see Turner, Blum, and Niss
(2015))” (Pettersen & Braeken, 2019, p. 407). The use of this analysis scheme with
mathematics teachers in studies by Pettersen and Braeken (2019) and Petersen and
Nortvedt (2018) is presented in the next section of this chapter.

Framework of Cognitive Demands

Smith, Henningsen, and Silver’s (2000) framework of cognitive demands of tasks


has received much attention in mathematics education, particularly in studies
involving task analysis and teachers’ knowledge or learning. The level of cognitive
demand refers to the kind and level of mathematical thinking and reasoning that is
required of students to successfully engage with a task. For example, as Stein and
Smith (1998) explained:
Tasks that ask students to perform a memorized procedure in a routine manner
lead to one type of opportunity for student thinking; tasks that require students
to think conceptually and that stimulate students to make connections lead to a
different set of opportunities for student thinking. (p. 269)
The framework has four levels of cognitive demands that rank a task as low level
or high level. The four levels are: memorization, procedures without connections,
procedures with connections, and doing mathematics. These are summarized
in Table 2.3 based on Stein et al.’s (2000) Task Analysis Guide. Smith and Stein
(1998) note that to evaluate a task effectively, it is necessary to look at the extent to
which it provides students with opportunities to engage in important mathematical

Table 2.3. Cognitive demands or complexity of mathematical tasks

Type of tasks Descriptions Levels of cognitive


demands/complexity

Memorization Tasks require students to reproduce Low demand


previously learned facts, rules, formulas,
or definitions without using procedures
Procedures without Tasks require students to perform Low demand
connections algorithms that underlie the procedures
being used but have no connection to
understandings, meaning, or concepts
Procedures with Tasks require students to focus on High demand
connections meanings or concepts underlying the
procedures being used to solve problem
Doing mathematics Tasks require students to engage in High demand
complex and nonalgorithmic thinking

50

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

processes and practices, such as problem solving, reasoning, communication, making


connections, and representation.
Since the cognitive demands of mathematical instructional tasks can influence
students’ learning and doing of mathematics in the classroom (Henningsen & Stein,
1997), the framework has relevance and importance to teacher education. Examples
of this are presented in the next section of the chapter.

Framework of Level of Authentic Intellectual Quality

The authentic intellectual quality framework is a central aspect of our work in


authentic assessment. It was initially developed and used by the first author with
teachers in Singapore (see Koh, 2011a; 2011b for more details). Koh developed the
criteria for the framework based on Newmann and Associates’ (1996) standards of
intellectual quality, Anderson and Krathwohl’s (2001) revised Bloom’s taxonomy
of educational objectives, and Marzano’s (1992) learner-centered instruction. For
example, she drew on Newmann and associates “authentic intellectual work,” which
consists of three broad criteria: construction of knowledge, disciplined inquiry,
and value beyond the school. Specific standards are embedded within each of the
criteria, which “provide a benchmark for teachers to judge whether particular forms
of instruction and assessment are likely to help students produce authentic work”
(Scheurman & Newmann, 1998, p. 25). Koh used the revised Bloom’s taxonomy to
further unpack the construction of knowledge and disciplined inquiry criteria. This
approach resulted in a comprehensive framework consisting of seven criteria and
10 sub-criteria. Table 2.4 provides the key elements of this framework that form
the Authentic Intellectual Quality criteria for judging the quality of mathematics
assessment tasks or the authenticity of mathematics performance tasks. These
criteria are more fine-grained than Stein et al.’s (2000) levels of cognitive demand
and extend beyond them to include items related to competencies and assessment.
The first criterion, depth of knowledge, in Koh’s authentic intellectual quality
framework is similar to Stein et al.’s levels of cognitive demands. The criterion
is comprised of three sub-criteria: factual knowledge, procedural knowledge,
and advanced concepts. The second criterion, knowledge criticism, and the third
criterion, knowledge manipulation, are focused on students’ higher-order thinking
and reasoning, complex problem-solving, and knowledge production. Knowledge
criticism constitutes presentation of knowledge as a given, comparing and
contrasting information/knowledge, and critiquing information/knowledge while
knowledge manipulation is unpacked by reproduction; application/problem-solving;
organization, interpretation, analysis, synthesis, or evaluation of information; and
generation or construction of knowledge new to students. The fourth criterion,
extended communication, refers to students’ ability to make mathematical arguments
in both oral and written formats. The fifth criterion, making connections to the real
world beyond the classroom, emphasizes the authenticity of assessment tasks in terms
of their realistic values to meet the needs of students’ real lives. One of the greatest

51

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
Table 2.4. Authentic intellectual quality criteria for mathematics authentic assessment tasks

52
Criteria Descriptions Indicators

Depth of Knowledge The task calls for students to demonstrate an exemplary


understanding of the different types of knowledge.
‡ Factual knowledge A necessary knowledge base for the learning and mastery of Recognize mathematical terms; state
more advanced concepts. concepts, facts, or principles; identify objects,
patterns, or list properties
‡ Procedural knowledge Use of domain-specific skills, rules, techniques, tools, and Know how to carry out a set of steps;
methods. perform arithmetic operations
KIM KOH & OLIVE CHAPMAN

‡ Conceptual understanding Advanced concepts that are central to the task. Explain one or more mathematical relations;
understand how one major mathematical
topic/idea relates to another
Knowledge Criticism The task requires students to judge the value, credibility, and
soundness of different sources of information or knowledge
through comparison and critique.
‡ Presentation of knowledge No opportunity for questioning the validity of information or Accept or present ideas and information as
as a given knowledge. truth or a fixed body of facts; follow a set of
preordained procedures
‡ Comparing and contrasting Comparing and contrasting different sources of information/ Identify the similarities and differences in
information or knowledge knowledge. observations, data, and theories; classify,
organize, and compare data
‡ Critiquing information Opportunity for problematizing knowledge. Make mathematical arguments; pose and
or knowledge formulate mathematical problems
Knowledge Manipulation The task calls for students to apply their higher-order
thinking and reasoning skills.
‡ Reproduction Reproducing information expounded by “authoritative Reproduce facts or procedures; recall familiar
sources,” i.e., the teacher or textbooks. mathematical objects and properties

(cont.)

via University College London


Downloaded from Brill.com 04/03/2024 05:15:13PM
- 978-90-04-41887-5
Table 2.4. Authentic intellectual quality criteria for mathematics authentic assessment tasks (cont.)

Criteria Descriptions Indicators

‡ Organization, interpretation, Organizing, interpreting, analyzing, synthesizing, or Interpret tables, graphs, and charts; predict
analysis, synthesis, or evaluating different sources of information/knowledge. mathematical outcomes from the trends in
evaluation of information the data
‡ Application/Problem- Applying mathematical concepts or procedures to solve Use problem-solving heuristics for non-
Solving non-routine or new problems. routine problems
‡ Generation or Construction Making of own hypotheses/ conjectures and generalizations Come up with new proofs or solutions to a
of Knowledge New to in order to arrive at conclusions, solve problems, or discover mathematical problem; apply modeling to
Students new meanings new contexts
Extended Communication The task asks students to elaborate on their understanding, Generate diagrams, sketches, drawings, or
explanations, arguments, or conclusions through extended symbolic representations (e.g., graphs, tables,
communication. equations, number sentences) with elaboration
Making Connections to The task asks students to address a question, issue, concept,
the Real World beyond the or problem that resembles one that they have encountered
Classroom or are likely to encounter in daily life beyond the classroom.
Students are expected to apply their knowledge and skills or
to share their work with audiences beyond the classroom.
Student Control Students determine the parameters of a task such as topics or
questions to answer, alternative procedures, tools and resources
to use (e.g., textbook, internet, or newspaper), length of writing
or response, or success criteria and performance standards.
Explicit Performance The task is provided with the teacher’s clear expectations
Standards or Success Criteria for students’ performance and the success criteria are made
explicitly clear to the students. Reference to only technical
or procedural requirements (e.g., the number of examples,
length of an essay or response) is not evidence of explicit
performance standards or success criteria.
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

53

via University College London


Downloaded from Brill.com 04/03/2024 05:15:13PM
- 978-90-04-41887-5
KIM KOH & OLIVE CHAPMAN

advantages of the criterion is that it allows for ensuring that authentic tasks are not
camouflage. According to Cumming and Maxwell (1999), “camouflage occurs
when a traditional form of assessment is ‘dressed up’ to appear authentic, often by
the introduction of ‘real world’ elements or tokenism” (p. 188). An example of a task
that is camouflage: Mary and Tom each bought 20 candies. If Mary eats 4 candies
per day and Tom eats 2 candies per day, who will finish eating their candies first?
Such word problems are commonly found in elementary mathematics textbooks
and supplementary materials. The sixth criterion, student control, and seventh
criterion, explicit performance standards/success criterion, enable an examination
of the extent to which the authentic assessment supports formative assessment or
assessment for learning.
Similar to the framework of competencies, the authentic intellectual quality
framework has a scoring scheme associated with it (see Koh, 2011b). The scheme
consists of four different levels of intellectual quality for each item of the framework
ranging from 1 (lowest level) to 4 (highest level). Thus, it can be used to analyze,
select or design mathematics authentic assessment tasks. We share application of
this framework with teachers in a later section of the chapter.

MATHEMATICS TEACHERS’ KNOWLEDGE OF TASKS

The preceding frameworks suggest that for teachers to be able to recognize, select
or create and use appropriate tasks in their teaching to effectively support students’
learning and doing of mathematics from a reform perspective, they should understand
levels of cognitive demand or competence demand of the tasks and the authentic
intellectual quality of the tasks. Given this level of complexity in thinking about
tasks, in addition to other task characteristics and influences on tasks previously
discussed, it is understandable that studies, as in the examples that follow, have
indicated limitations in mathematics teachers’ selection, analysis, and enactment of
tasks for instructional and assessment purposes and ability to identify elements of the
frameworks in tasks. However, as also addressed in this section, some studies have
indicated possible ways of helping the teachers to enhance their task knowledge.
Selection of tasks. Studies have identified the level of mathematics teachers’
knowledge of selection of tasks based on the tasks they use in their teaching. Thus,
limitations in the tasks can be related to limitations in their knowledge. Some
studies highlighted the level of demand of the tasks. For example, Silver, Mesa,
Morris, Star, and Benken (2009) found that teachers in the United States used
tasks in instruction which included a range of mathematics topics but were not
consistently intellectually challenging and tasks that involved hands-on activities or
real-world contexts and technology rarely required students to provide explanations
or demonstrate mathematical reasoning. Baumert et al. (2010) showed that tasks
provided by German grade 10 teachers had a low overall level of cognitive challenge.
In Swedish mathematics classrooms, the complex nature of the competencies is not
fully captured as aspects such as the ability to evaluate and reflect on mathematics

54

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

and to draw conclusions are not involved in the tasks on tests or instruction (Boesen
et al., 2010). Studies in South America also indicate use of tasks with very low level
of cognitive demand (e.g., Cueto, Ramírez, & León, 2006). However, in the case of
a grade 8 teacher in Taiwan, Hsu and Silver (2014) found that the geometry tasks she
used from non-textbook sources (e.g., materials designed by her or collaboratively
with her colleagues) were more cognitively demanding than those taken directly
from textbooks. Other studies highlighted other aspects of task selection. For
example, Hiebert et al. (1997) found that teachers in the United States selected
instructional tasks largely based on the concepts and skills that they need to cover
in their teaching. Also, Lingard et al. (2001) and Koh and Luke (2009) found that,
in Australia and Singapore, respectively, teachers’ assessment tasks tended to mirror
the content and format of high-stakes examinations.
Analysis/appraisal of tasks. Another aspect of teacher task knowledge is reflected
in their analysis/appraisal of tasks as in the following studies. In the United States,
Arbaugh and Brown (2005) found that when analyzing tasks, teachers tended to focus
on surface features such mathematical content, context, and use of manipulatives,
while Boston (2006) and Osana, Lacroix, Tucker, and Desrosiers (2006) found that
practising teachers and prospective teachers, respectively, experienced difficulties to
recognize and understand the cognitive demands of problems involving high levels
of complexity. Focusing on mathematics assessment tasks, Pettersen and Nortvedt
(2018) found that the sample of Norwegian practicing and prospective teachers,
who used a competency analysis tool based on Table 2.1 to identify the level of
competency demand of the tasks, demonstrated high consistency in carrying out
the analysis, but they utilized a restricted range of the scale, rarely judging a task to
demand a high level of competence. Thus, while they could use tool with minimal
training to identify which of the competencies were involved in solving a task, they
could only differentiate to a limited extent between tasks that demand a low level of
competence and those that demand a high level. Focusing on labels for tasks, Foster
and Inglis (2017) investigated United Kingdom teachers’ appraisals of adjectives used
to describe mathematics tasks and found that their task appraisals varied on seven
dimensions: engagement, demand, routineness, strangeness, inquiry, context and
interactivity. They also found that among teachers there was some agreement about
inquiry and context, some disagreement about routineness and clear disagreement
about engagement and demand. Collectively, these studies suggest underlying issues
with teachers’ task knowledge.
Enactment of task. Some studies focused on understanding teachers’ enactment
of mathematical instructional tasks and how the cognitive demands of the tasks were
maintained or declined between task set up (i.e., when the task is announced by the
teacher) and task implementation (i.e., when students begin to work on the task). For
example, when the tasks were enacted in the United States the cognitive demand
lessened (Resnick, 2006). Researchers found that teachers in the United States have
difficulty implementing tasks of high cognitive demand, even when they are planned
as such (Boston & Smith, 2009). Stein et al. (1996, 2000) found that it was challenging

55

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

for teachers to maintain a high demand tasks from the phase in which the teacher sets
up the task in the classroom through the phase in which the task is “implemented
by students during the lesson” (p. 460). They found that in classrooms where tasks
with the potential for high levels of cognitive demand were used, teachers and/or
students often decreased the cognitive demand during implementation of the tasks.
In Estrella, Zakaryan, Olfos, and Espinoza (2016) investigated the factors affecting
the maintenance and decline of cognitive demand in the Chilean primary school
teachers’ implementation. In terms of the decline in the level of cognitive demand,
the greatest changes were found in factors corresponding to the routinization
of problematic aspects of the task. Wilhelm’s (2014) study sought to understand
how aspects of United States middle school mathematics teachers’ knowledge and
conceptions are related to their enactment of cognitively demanding tasks (i.e.,
selection and maintenance of the cognitive demand of high-level tasks). Findings
indicated that the teachers’ mathematical knowledge for teaching and conceptions of
teaching and learning mathematics were contingent on one another and significantly
related to teachers’ enactment of cognitively demanding tasks. Baumert et al. (2010)
found that the pedagogical content knowledge of German mathematics teachers
determined the level of cognitive activation in the tasks they assigned to students
and the level of cognitive challenge of the observed tasks was low.
The preceding sample of studies regarding teachers’ task knowledge suggests that
if teachers are not able to select, analyze, and use worthwhile tasks in instruction they
are also unlikely to use such tasks for assessment particularly authentic assessment.
However, as discussed next, appropriate intervention could make a difference to
their learning and use of worthwhile tasks in instruction and, in the case of our work,
in assessment.
Enhancement of task knowledge. The following are examples of studies that
indicate the level of success in helping teachers to enhance their task knowledge
for teaching, in particular, regarding cognitive demands of tasks. Boston (2006) and
Boston and Smith (2009) reported on a study that analyzed mathematics teachers’
selection and implementation of instructional tasks in their own classrooms before,
during, and after their participation in a professional development workshop focused
on the cognitive demands of mathematical tasks. The 18 United States secondary
mathematics teachers who participated in a six-session professional development
workshop enhanced their knowledge of the cognitive demands of mathematical tasks.
In particular, they improved their ability to identify and describe the characteristics
of tasks that influence students’ opportunities for learning. They improved their
ability to maintain high-level cognitive demands during implementation and the
frequency in selecting high-level tasks as the main instructional tasks in their own
classrooms. There were significant differences between them and the contrast group
that did not participate in the workshops in task selection and implementation during
lesson observations. They also outperformed the contrast group on the post-measure
of the knowledge of cognitive demands of mathematical tasks. Using a different
professional development approach, Arbaugh and Brown (2005) sought to understand

56

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

how focusing the teachers on critically examining mathematical tasks influenced


their thinking about the nature of mathematical tasks as well as their choice of
tasks to use in their classrooms. They found that the teachers showed growth in the
ways that they consider tasks, and that some of the teachers changed their patterns
of task choice. In another intervention study, Polly, Neale, and Pugalee (2014)
examined how a task-focused, year-long mathematics professional development
SURJUDPLQÀXHQFHG8QLWHG6WDWHVHOHPHQWDU\VFKRROWHDFKHUV¶NQRZOHGJHEHOLHIV
and practices. The observed teachers enacted some high-level mathematical tasks
and questions, but these were more visible at the end of the study compared to the
beginning of the study. In one study outside the United States, Estrella, Zakaryan,
Olfos, and Espinoza (2016) reported on two Chilean primary school teachers who
participated in a Lesson Study group to design and implement statistical tasks of
high cognitive demand. They found notable changes in the maintenance factors
associated with the teachers’ questions and feedback, the scaffolding provided for
the students’ reasoning and the modelling of high-level performance. But there was
decline in the level of cognitive demand, with the greatest changes found in factors
corresponding to the routinization of problematic aspects of the task.
Conclusions. In general, then, research shows that while there are different
limitations in the teachers’ task knowledge for teaching, providing them with
relevant experiences could make a difference to this knowledge and their practice.
For example, engaging them in learning experiences involving the cognitive demand
and competency frameworks as tools to analyze tasks as demonstrated in the studies
of Arbaugh and Brown (2005) and Boston and Smith (2011) and suggested in
the study of Pettersen and Nortvedt (2018) could be useful to them as a means of
selecting or creating and using worthwhile tasks to enhance their students’ learning of
mathematics and development of mathematical competencies. As Stein et al. (2000)
argued, the cognitive demand framework can support teachers in differentiating
between different levels of demand and raise their awareness of the demand of tasks
and how they relate to goals for student learning. Our work discussed next also
indicates the promising potential of the authentic assessment task framework to
support teachers’ learning of authentic mathematics assessment tasks.

BUILDING TEACHERS’ CAPACITY IN MATHEMATICS


AUTHENTIC ASSESSMENT

For teachers to successfully implement reform-based mathematics curriculum,


they need to know how to purposefully plan for and enact tasks of high cognitive
demand, not only for instruction but also for assessment. As Lingard et al.’s
(2001) and Koh and Luke’s (2009) studies showed, when teachers assigned
realistic and intellectually demanding assessment tasks, most students were able
to rise to the challenge by producing authentic intellectual work. Thus, building
teachers’ capacity in mathematics authentic assessment needs its own attention
in mathematics teacher education or professional development. In this section,

57

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

we contribute to this by placing the focus on teachers’ learning of authentic


assessment. We first discuss the nature of authentic assessment, followed by a
discussion of teachers’ knowledge of assessment, and then offer examples from
our recent and ongoing work with prospective and practising teachers to support
their development of knowledge of authentic assessment and authentic assessment
tasks in teaching mathematics.

Theory of Authentic Assessment

The term authentic as used in authentic assessment was first coined by Archbald
and Newmann (1988) to define the qualities of student achievement that are deemed
meaningful and worthwhile in the context of K–12 schooling. Newmann and
Archbald (1992) stated that authentic achievement can be measured by three broad
criteria: construction of knowledge, disciplined inquiry, and value beyond school.
Although Archbald and Newmann introduced the term authentic, Wiggins (1989)
was the first to use it to describe the desirable features of grades K–12 assessments
that help prepare students for success in their future lives and workplace. According
to him, there are four key principles of authentic assessment consisting of: (1)
Authentic assessment tasks (as previously defined). (2) Success criteria and
performance standards presented in well-developed rubrics. These are explicitly
shared with students and others in the learning community, that is, the assessment
needs to be transparent. (3) Students’ involvement in self-assessment. This requires
students’ active engagement in reflecting, revising, modifying, and redirecting their
efforts to improve the quality of their work or performance. Students take initiative
in monitoring their own progress and evaluating their own work against the success
criteria and performance standards in a rubric. (4) Students presenting and defending
their work or performance to real audiences.
Specific to mathematics education, the term authentic in relation to authentic
assessment was first adopted by Romberg (1995) who equated it to “trustworthy”
(p. vii), noting that “assessment of student performance should be trustworthy
indicators of mathematical power” (p. vii), for example, it should indicate how
well a student can solve the authentic assessment tasks, which need to serve as a
reliable and valid indicator of students’ mathematical understanding (Romberg &
Wilson, 1995). Romberg and Wilson associated authentic assessment with teaching
that engage students in “doing mathematics,” that is, actively engaging students in
a set of “dynamic and integrative activities” that include “discovering, exploring,
conjecturing, sense making, and proving” (p. 4). This view of learning mathematics
led them to underscore the importance of designing authentic assessments that
enable “students to engage in rich activities that include problem solving, reasoning,
communications, and making connections” (p. 4). They also advocated for the design
and use of authentic assessment tasks, based on Archbald and Newmann’s (1988)
criteria (i.e., construction of knowledge, disciplined inquiry, and value beyond
school) to capture these important mathematics reform-based curriculum outcomes.

58

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Lajoie (1995), another mathematics educator, also proposed principles for


operationalizing authentic assessment in the context of teaching and learning of
mathematics that include: (1) Authentic assessment should be multidimensional
to serve as multiple indicators of student learning and performance. For example,
it should attend to cognitive dimensions as well as conative or socioemotional
dimensions of learning. (2) Authentic assessment needs to be “relevant, meaningful,
and realistic” (p. 30). For example, it should attend to desirable learning outcomes,
students’ ability to apply their knowledge and skills, and meaningfulness to students’
real lives. (3) Authentic assessment must be accompanied by reliable and valid
scoring and rating procedures as in the form of well-designed rubrics. (4) Authentic
assessment must yield information that help improve teaching and learning. (5)
Authentic assessment should take into account gender, racial or ethnic, cultural, and
aptitude biases. For example, the content and format of assessment tasks must be
responsive to the learning, emotional, and social needs of culturally and linguistically
diverse students. (6) Authentic assessment must be an integral part of the instruction.
For example, it should be treated as formative assessment or assessment for learning
practice to support instruction and student learning.
Some researchers consider authentic assessment to be the best assessment to
use in the classroom because it is both valid and relates to the real world and it
can be used to examine student performance on intellectual tasks that require prior
knowledge and reasoning skills (e.g., Frey & Allen, 2012; Wiggins, 1990) and
crafting of thorough and justifiable answers (Wiggins, 1990).
The preceding views of authentic assessment suggest that it could be interpreted as
formative assessment, which is now strongly recommended for use in the mathematics
classroom with suggestions to help teachers to do so (e.g., Fennell, Kobett, & Wray,
2017; Silver & Mills, 2018). However, authentic assessment requires special types
of mathematical tasks (i.e., authentic mathematics tasks as previously discussed), the
use of rubrics and connection to real-world or real audiences to make it authentic. In
our work with mathematics teachers, we focus on these elements of assessment that
have not received much attention in the mathematics education literature.

Teachers’ Knowledge of Authentic Assessment

Practising teachers’ assessment literacy, especially their competence in the design,


selection, and use of alternative forms of assessment, has been an area of concern in
many education systems around the world. Practising teachers have reported a low
level of assessment literacy due to their inadequate assessment preparation at the
prospective teacher level (e.g., Mertler, 2009; Popham, 2011). Additionally, many
beginning teachers were found to lack confidence in assessing their students.
In Watt’s (2005) study, teachers indicated that they were satisfied with traditional
mathematics assessments and used occasional alternative assessment methods such
as oral tasks and observations. There were six main responses to the question of
why alternative assessments were not used: insufficient time for implementation,

59

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

unstructured nature, unsuitable, unreliable/subjective, and insufficient resources at


hand to permit implementation.
Limitations in practising teachers’ knowledge of alternative assessment is likely
to be transfer to prospective teachers, who, as research indicate, tend to learn from
and follow their mentor teachers’ assessment practices during their field experiences
(DeLuca & Klinger, 2010). Research on prospective mathematics teachers have
indicated that, without appropriate intervention, their knowledge and beliefs of
assessment tended to be more about traditional or summative assessment (e.g., “paper
and pencil tests”) and much less about alternative or formative assessment (Coffey,
2000; Santos & Teixeira, 2018; Wallace, 2012). Their perspective of assessment
included procedural tasks that were used for grading purposes or to identify topics
to review. They were more familiar with the written test, felt insecure with other
instruments, and in general, had narrow views of assessment. In some cases, even
with exposure to alternative assessment, many still showed superficial changes in
their thinking of non-traditional assessment.
Mathematics curriculum reforms require corresponding reforms in how students’
learning and performance are assessed in the grades K–12 mathematics classroom
(Lesh & Lamon, 1993; NCTM, 2000; Romberg, 1995). This means that teachers
implementing such reforms must change their assessment practices to include
alternative forms of assessment such as authentic assessments that enable richer
demonstration and more holistic representation of what students know and can
perform in mathematical and real-world contexts. However, as Webb (2009) pointed
out, “Teachers’ limited conceptions of and confidence in assessment often restrict
implementation of [reform] mathematics curricula … therefore, there is an urgent
need to provide mathematics teachers opportunities to develop their expertise in
classroom assessment” (p. 1).
While there has been a growing body of literature to help mathematics teachers
to implement formative assessment (e.g., Fennell et al., 2017; Silver & Mills, 2018)
and some studies on their development of task knowledge to support alternative
assessment, there has been little focus on their knowledge of authentic assessment
or authentic assessment tasks as defined in this chapter. One related study is Borko,
Mayfield, Marion, Flexer, and Cumbo (1997). In line with mathematics curriculum
reform, they examined the process of change in a group of third-grade teachers’
mathematics assessment practice. They conducted year-long series of weekly
workshops to help the participating teachers design and implement classroom-based
performance assessments (i.e., authentic assessments), which were aligned with the
teachers’ instructional goals. Borko et al. (1997) found that although most teachers
developed their ideas about mathematics performance assessment, their beliefs were
incompatible with the reformed mathematics agenda. They further suggested that
teachers need to work together with university researchers and their colleagues in
school-based professional learning community. This provide them with not only the
opportunity to see themselves as an equal partner of the researchers, but also ample
time to confer with their colleagues, experiment with new forms of assessment, and

60

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

reflect on their beliefs and practices. Taken together, these would lead to changes in
their classroom assessment practices.
The preceding discussion suggests that both practising teachers and prospective
teachers (who learn assessment mostly from practising teachers) have limitations
in their assessment literacy in general, and alternative innovative assessments such
as authentic assessment, in particular. Thus, they need special help to develop
knowledge of, and implement, assessment methods to improve teaching and learning
of mathematics consistent with reform-based curriculum and pedagogy. In the next
section we present examples of our work that are intended to achieve this. Taken
together, these suggest the importance of building mathematics teachers’ capacity in
the design, selection, and use of mathematics authentic assessment.

Applying the Authentic Intellectual Quality Framework to Practising


Mathematics Teachers’ Learning

Koh’s (2011a) authentic intellectual quality criteria (Table 2.3) were developed
to support teachers in the design, adaptation, selection, and use of mathematics
authentic assessment tasks. They have been proven useful for helping practicing
teachers to gain a deeper insight into the desirable features of authentic assessment
tasks in several core subjects including mathematics. For example, in the following
two studies with practising elementary school mathematics teachers in Singapore
and Canada, the authentic intellectual quality criteria were found to improve the
teachers’ assessment literacy pertaining to the aspect of task design
Singapore teachers. In this intervention study, Koh (2011a, 2014) demonstrated
the importance of developing elementary practising teachers’ (i.e., grades 4 and 5)
assessment literacy, especially in the aspect of authentic assessment task design
through the provision of ongoing, sustained professional development. In the
study, practicing teachers from two public schools in Singapore were engaged in
workshops and school-based professional learning community pertaining to the
design of mathematics authentic assessments over two school years. Koh’s (2011a,
2011b) criteria for authentic intellectual quality (see Table 2.3) were used as the
guideposts for the teachers in their design of mathematics authentic assessments.
Koh (2011a, 2014) found that the teachers’ knowledge about the principles of
authentic assessment had significantly improved at the end of the intervention.
More important, the teachers had adopted the criteria for authentic intellectual
quality in their design of mathematics authentic assessments for use with their
students in mathematics lessons. The improved assessment tasks yielded positive
effects on the quality of students’ work. This finding indicates the importance of
building mathematics teachers’ capacity in the design, selection, and use of authentic
assessment tasks.
Canadian teachers. This case took place in an elementary school where a group of
practicing teachers were actively involved in designing and implementing authentic
assessment for a grade 6 geometry unit of instruction. A critical inquiry approach was

61

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

employed to support the teachers to develop their capacity in mathematics authentic


assessment and assessment for learning during their school-based professional
learning community. The teachers participated in eight biweekly professional
learning community sessions over a school year. The teachers were asked to
reflect on the various elements of Wyatt-Smith and Gunn’s (2009) assessment as
critical inquiry framework; namely, their conceptions of mathematical knowledge,
the principles governing their professional judgement, the relationship between
assessment, learning and teaching, and the curricular literacies in mathematics
education. Throughout the remaining professional learning community sessions, the
teachers were introduced to and guided through the design principles of authentic
assessment, the Koh criteria for authentic intellectual quality, the patchwork text
approach, and the Structure of the Observed Learning Outcome (SOLO) taxonomy.
The SOLO taxonomy, originated from Biggs and Collis (1982), was used to
support teachers’ identification and framing of learning outcomes from the grade 6
mathematics curriculum. It also enabled them to ensure a close alignment between
assessment tasks, instructional activities, and learning outcomes while planning
and designing the geometry unit of instruction. The taxonomy describes levels
of increasing complexity in a student’s understanding of a subject (i.e., learning
progression), through five stages: pre-structural, uni-structural, multi-structural,
relational, and extended abstract. At both pre-structural and unistructural levels,
students have minimal or no knowledge about a topic being investigated and hence
assessment tasks and instructional activities need to develop students’ fundamental
knowledge and skills. Progressing from unistructural to multistructural, students
are expected to acquire the relevant knowledge and skills. At the relational level,
students’ higher-order thinking manifests in their ability to make and explain the
connections between different ideas around a related topic. Extended abstract is the
highest level of the taxonomy and a culminating assessment (e.g., a project) is used
to enable students to apply what they learned in the previous tasks.
To help the teachers internalize the authentic intellectual quality criteria, they
were also asked to analyze existing assessment tasks using the criteria. At the end of
the professional learning community, the teachers decided to select a mathematics
unit of instruction for their collaborative design of an authentic assessment. A
series of assessment tasks were developed to be closely aligned with the intended
learning outcomes across the various complexity levels of the SOLO taxonomy.
The patchwork text approach enables the mapping of the assessment tasks in the
geometry unit to the specific learning outcomes across five different levels of the
SOLO taxonomy: pre-structural, uni-structural, multi-structural, relational, and
extended abstract. All the tasks from the unistructural to the relational levels were
used as formative assessment or assessment for learning. The teacher incorporated
informal observations, formative feedback, and student self-assessment and
reflections through the use of analytic rubrics and checklists. These formative tasks
aimed to scaffold students learning and progress toward the culminating project (i.e.,
a garden design along with an oral presentation). Table 2.5 provides an example

62

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Table 2.5. An example of patchwork text assessment approach (Koh et al., 2015)

SOLO taxonomy Task

Pre-structural Student reflections on prior knowledge


Uni-structural Classifying shapes & angles
Multi-structural Measuring angles and drawing garden base map
Relational Problem-solving – applying shapes and space to design garden
Extended abstract Final garden design and oral presentation

of the use of the Structure of the Observed Learning Outcome Taxonomy and the
various assessment tasks that lead toward the final design/project (i.e., Patchwork
Text Assessment Approach).
In short, the mathematics authentic assessment was designed by the teachers
to provide students with opportunities to engage in mathematical reasoning
and critical thinking, application of mathematical concepts to solve real-world
problems, extended communication, collaboration, generation of new knowledge,
and making connections to other subject areas. Our findings have shown that
teachers’ rich professional conversations over the features of authentic assessment
tasks, the authentic intellectual quality criteria, and the identification of specific
mathematics learning outcomes using the SOLO taxonomy have not only improved
their assessment literacy (i.e., understanding of authentic assessment and assessment
for learning), but also increased their mathematical knowledge for teaching (i.e.,
identifying students’ errors and misconceptions, understanding students’ thinking
and reasoning, Ball, Thames, & Phelps, 2008). Teachers’ critical inquiries into
the alignment between assessment, curriculum, and pedagogy along with their
application of the authentic intellectual quality criteria in their collaborative design
of mathematics authentic assessment enabled them to develop ‘designers’ eyes’
(Webb, 2009). They have become competent in the design of assessment tasks that
are well aligned with the desirable mathematics learning outcomes, with an eye
toward promoting students’ learning of mathematics with understanding.

Applying the Authentic Intellectual Quality Framework to Prospective Mathematics


Teachers’ Learning

As noted above, the authentic intellectual quality framework was developed to


use with practising teachers and was found to be effective in helping them to build
capacity in mathematics authentic assessment. However, in extending the use of
it, we are in the process of exploring its use with prospective teachers. We are
not aware of this framework being studied to understand its use with prospective
teachers. Thus, despite the early stage of our work with them, share aspects of it
consisting of a completed pilot study (Chapman & Koh, 2017) and the initial stage

63

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

of an in-progress project1 (Chapman, Koh, & Piñeiro, 2019) to draw attention to the
potential and importance of integrating this framework in an intervention aimed at
building elementary school prospective teachers’ capacity in mathematics authentic
assessment.
Pilot study. Our description of this study is based on our presentation at the
2017 – North American Chapter of the International Group of the Psychology of
Mathematics Education conference and reproduces selected sections of the paper
(Chapman & Koh, 2017). Participants of this pilot study were three prospective
teachers specializing in mathematics education; one at the elementary and two at
the secondary school levels. The secondary school prospective teachers were at
the end of their two-year Bachelor of Education program and had completed an
assessment course, while the elementary school prospective teacher was at the end
of the first year of the same program and had not completed the assessment course.
They participated in the Authentic Assessment Learning Activities (the intervention)
during four two-hour sessions led by the researchers at the university.
The Authentic Assessment Learning Activities include: use of the authentic
intellectual quality criteria (Table 2.4) to analyze tasks and guide the creation of
tasks; use of design principles of authentic assessment and associated rubrics to
guide development of assessment strategies and analysis and creation of rubrics;
and use of the SOLO taxonomy to guide the identification and framing of learning
outcomes from the mathematics curriculum and to ensure a close alignment
between assessment tasks, instructional activities, and learning outcomes. The
participants worked individually and as a group in carrying out the activities that
included analyzing examples of different mathematics assessment tasks, revising
a mathematics assessment task selected from their field experience to make it
authentic, and creating an authentic assessment task with rubric for a mathematics
topic and grade of their choice with explanation of how they would use it to assess
and guide students’ learning.
Findings of this pilot study (Chapman & Koh, 2017) indicated that participants’
engagement in the Authentic Assessment Learning Activities helped them enhance
their understanding of authentic assessment and authentic assessment tasks in
teaching mathematics. At the beginning of the intervention, there was little difference
between the two participants who had completed the assessment course in their
teacher education program and the one who had not regarding their conceptions of
authentic assessment and tasks. While the course enabled them to develop initial
understanding of forms of assessment of and for learning, it did not allow them
to conceptualize authentic assessment and related tasks adequately, in general
or specific to mathematics. The intervention helped them start to develop more
meaningful understanding of authentic assessment tasks to assess what students
know and can perform in mathematical and real-world contexts.
The two secondary school prospective teachers (PT-A and PT-B), who were at the
end of the teacher education program, submitted final authentic tasks they developed
for units of work of their choice, which they expressed strong interest to implement

64

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

in their teaching as beginning teachers. The third participant, the elementary school
prospective teacher, was unable to complete and submit a final task because of
personal issues occurring in her life. Table 2.6 contains PT-A’s authentic task for use
in a grade 7 algebra unit that includes concepts of equality, expressions and linear
equations based on the provincial mathematics curriculum. PT-A also explained
in detail how the task would be integrated into instruction and used in formative
assessment and for authentic assessment. The rubric for authentic assessment
consisted of four columns ranging from insufficient planning to exemplary field trip
plan and three rows consisting of model and solve problems and equations, logistics,
and representation of equality.

Table 2.6. PT-A’s grade 7 authentic mathematics task

Task 1: Equality and marshmallows


In groups of 2, discuss and write down your answers to the following 4 questions.
Remember, there is no right or wrong answer! We will regroup and share our answers
after __ minutes.
What is equality?
What does it mean to “preserve” something? To “preserve” equality?
What are examples of equality in your everyday life?
What are examples of equality in math?
Marshmallow challenge:
Utilizing the marshmallows and toothpicks as connectors handed out, build a structure
that is “different, yet equal” at the same time. you have to use at least 10 marshmallows.
How are you and your partner’s structure different?
how are you and your partner’s structure equal?

Task 2: Planning a field trip


**This is an ongoing task and lessons will constantly refer back to this task.
You will work in groups to plan a field trip. Your group will decide on a possible field trip
destination and plan the logistics of the entire trip (all the nitty-gritty stuff as well).
Things that you will have to keep in mind to make the field trip idea work:
‡ Cost
‡ Cost per person
‡ Who is going? The whole class, the whole grade, the whole school, etc.
‡ How many school buses are needed? Volunteers?
‡ Will lunches be provided? How much is needed?
‡ How long does it take to get there by bus and back? When would you leave?
‡ What other things should you keep in mind?
*Hint: simple linear equations will hep with a lot of theses calculations (cost, cost per
person, etc.!
The best planned field trip will be chosen and made an actual field trip!
**Please check rubric for how you will be marked.

65

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

PT-B’s authentic task for use in a grade 10 algebra unit focused on linear relations
and functions based on the provincial mathematics curriculum. The following are
some features of the task, which is too long to reproduce here.
Part1: Students will pair up and flashcards with different figures will be
distributed (i.e., labeled graphs, images of everyday objects (e.g., basket
of fruit, ski hill, thermometer, etc.), table of values, bank statements, and
equations. Students will discuss in their groups:
What relationships exist within the photo(s)?
How could you represent these relationships?
Other parts implemented at different points in the unit included: (a) Students working
in small groups to devise a way to relate and graph their collected data in part 1.
They were expected to interpret and explain the relationships among their data,
graphs, and situations be able to state a reasonable domain and range and explain
any restrictions they set. (b) Students working in their groups to analyze their graphs
and collected data, identify if their relation is linear and confirm if their relation is
a function.
These examples of the prospective teachers’ tasks show what they were able to
accomplish in the short period of working with the Authentic Assessment Learning
Activities. There were clear shifts in their knowledge of worthwhile mathematics
tasks and how to develop and use of them in formative or authentic assessment
compared to what they demonstrated at the beginning of the intervention.
The participants indicated that the intervention provided them with a practical and
systematic way of making sense of selecting, unpacking, adapting, and designing
authentic tasks for assessment and instruction in mathematics. However, in using the
authentic intellectual quality criteria and scoring scheme, it was challenging for them
to analyze the level to which the tasks required deep understanding and promoted
knowledge criticism, higher-order thinking, reasoning skills, and connections to the
real world beyond the classroom. But, with guidance, this created opportunities for
them not only to understand the strengths and limitations of the tasks and to make
meaningful suggestions to modify them, but also to understand the meaning of these
mathematics task features from a reform-oriented perspective. Applying the authentic
intellectual quality criteria also challenged their understanding of mathematics
concepts involved in the assessment tasks being analyzed or being created. Again,
with guidance, this created opportunities for them to think about these concepts in
alternative ways. As one participant explained: “I think it’s really interesting that our
perception of how well we understand a concept is certainly pushed and tested when
you’re trying to develop a task.” Another noted, regarding designing the rubric: “I
found it challenging to narrow down the task and imagine what expectations I had
for the final project.” The intervention, then, also has potential to contribute to their
development of mathematics knowledge for teaching other than the task knowledge.
The main concern about the intervention was the need for more time and practice
with analyzing and creating tasks to allow for deeper engagement with the Authentic

66

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Assessment Learning Activities and group discussions. There were also issues
regarding their mathematics knowledge for teaching that seemed to limit the ability
to interpret the authentic intellectual quality and design authentic tasks. This situation
was more evident in a follow-up pilot with two elementary prospective teachers who
were constrained by their instrumental understanding of the mathematics concepts
for which they chose to design authentic tasks. The intervention was modified to
take these issues into consideration for the larger project.
The modified intervention consists of two 5-day summer institutes at the end of the
first and second year, respectively, of their education program. The focus of the first
institute is on helping them to develop their mathematical content knowledge and
mathematical knowledge for teaching, as well as the basic principles of mathematics
authentic assessment and the authentic intellectual quality criteria. The second
institute will focus on developing their expertise in using the criteria for authentic
intellectual quality to select, adapt, and design authentic intellectual assessment
tasks for teaching mathematics. Since holding knowledge does not mean it will be
implemented or implemented appropriately, they will also be tracked as beginning
teachers to determine their assessment practices in mathematics classrooms to
further establish the impact of the intervention.
Participants’ pre-intervention thinking. Participants for the larger in-progress
project are 16 elementary school prospective teachers who volunteered to participate
in the study. They were at the end of the first year of their two-year Bachelor of
Education program and had completed the required assessment course in the
program. They all hold university degrees but not in mathematics or mathematics-
related disciplines. At the beginning of the project the participants were interviewed
and wrote reflective journals on prompts to capture their initial thinking prior to the
intervention. Findings are based on analysis of this data.
Findings reported in Chapman, Koh, and Piñeiro (2019) indicate that most of
the prospective teachers held a surface understanding of authentic assessment
and authentic assessment mathematics tasks that reflected more features of
traditional assessment than formative assessment. As one participant explained
“authentic means real … real evaluation of where students are at with their math
understanding and knowledge.” This is representative of how participants made
sense of ‘authentic’ in conceptualizing authentic assessment. However, three
themes emerged regarding their conceptions of authentic assessment: use to teacher
(e.g., allowing the teacher to see, gauge, or understand what students actually
know), use to student (e.g., to really show what they learned during a lesson,
to engage with the material in their way, to apply their mathematical learning),
and the task (a real, real-world, or real-life task). While all of the participants
associated “real world” (e.g., real-world example, application, concept, task, or
problem) as the central feature of authentic assessment tasks, most of them had
a traditional word problem interpretation of real world. This was reflected in the
examples and explanations they provided to illustrate worthwhile mathematics
and assessment tasks. For example:

67

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

Task: If a baker bakes 3 cakes each day and the baker works for 5 days, how
many cakes would they have? Show two ways to solve this problem.
Explanation: Gives insight to not only if they know the answer but that they
know and understand the steps to getting the answer.
Task: Sally had 10 cookies. She gave 2 to Sam and ate 2 more. Steve had two
more cookies than Sally, but he ate 4 cookies in total. How many cookies do
Sally and Steve now have?
Explanation: It would identify areas of understanding, and more importantly,
misunderstanding that students have of the concept.
Most of their examples were routine-oriented problems that mimicked real-world
situations as a context for the task and lacked depth in terms of the Authentic
Intellectual Quality criteria. Their conceptions of what made their examples good
assessment tasks for students were also traditional-assessment oriented and included
that they allow them to show: their thinking, if they understand the concept, if they
know and understand the steps to get the answer, areas of their understanding and
misunderstanding, if they can explain or demonstrate their thinking, and multiple
answers.
Compared to the authentic intellectual quality criteria, on the surface, the
prospective teachers’ conceptions seemed to include two key elements of authentic
tasks: (1) Real-world connection, but for the most part their interpretation of it did
not attend to the authenticity of an out-of-school real-world task. In most cases, it
meant instrumental use of the real-world context or situation as a way for students
to see the use of the mathematics concept being learnt. (2) Application and problem
solving, but for the most part these involved a focus on numerical computations
and procedural knowledge. Outside of these two criteria, there was little or no
consideration of levels of cognitive demand or levels of competencies associated
with the tasks or the other authentic intellectual quality criteria.
The study provided further evidence to support the need for special attention
in teacher education research to determine effective ways of helping prospective
elementary school mathematics teachers to develop meaningful and useful know-
ledge of authentic assessment and mathematics tasks. Improving their knowledge
of authentic assessment will require clarifying, extending, and, for some aspects,
reconstructing their conceptions of assessment, curriculum, teaching and learning
and coherence among them. Our intervention in the larger project is intended to help
them to achieve this.

DISCUSSION AND CONCLUSION

Teachers need to hold mathematics task knowledge for teaching that is aligned with
reform-perspective of mathematics teaching and learning. This includes several
factors as highlighted in Chapman (2013b). In this chapter we focused on the need

68

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

for teachers to understand worthwhile tasks for instruction and worthwhile tasks for
assessment. While these tasks are directly related, their purposes are different with the
former focusing on learning and the latter on evaluating that learning. However, given
the relationship between them, our earlier discussion in this chapter of the nature of
worthwhile tasks provides a basis for the nature of assessment tasks. In particular, we
address these tasks from the perspective of reform-based mathematics curriculum that
includes 21st century competencies. We also draw attention to mathematics teachers’
inadequate level of task knowledge and examples of limitations in this knowledge
regarding selecting and analysing tasks and enacting tasks in the classroom. These
limitations are related to challenges the teachers encounter in implementing reform-
based mathematics pedagogy; challenges associated with lack of appropriate
mathematics knowledge for teaching. However, there is evidence that with appropriate
intervention their task knowledge can be enhanced to make a meaningful difference
to their teaching. These interventions could include frameworks to assess quality of
tasks, such as the three we described, that is, level of cognitive demand, level of
competence, and authentic intellectual quality of tasks.
Based on our work, we highlighted authentic assessment tasks, which shares a
lot of common principles with reform-based mathematical instructional tasks. One
of the most important principle is the focus on developing students’ mathematical
conceptual understanding, mathematical thinking and reasoning, and communication.
The richness, high intellectual demands and real-world elements of authentic
assessment tasks, as well as their use in authentic or formative assessment make them
instructionally responsive to students irrespective of their sociocultural backgrounds.
These approaches to assessment help students to develop positive habits of mind,
self-directed learning dispositions, and growth mindsets (Boaler, 2016). In addition,
the real-world elements of authentic tasks can motivate students as they perceive the
relevance of learning and assessment, which in turn contributes to improved quality
of work or performance. Stein and Lane (1996) found that the use of tasks that
involved high levels of cognitive demand led to greater student gains on a performance
assessment involving high levels of mathematical thinking and reasoning.
While authentic assessment tasks are highly beneficial to students, they are
unlikely to be implemented in mathematics classrooms without support to teachers
through teacher education or professional development programs. Not only
developing, selecting, and implementing authentic tasks is challenging for both
teachers, but also maintaining the cognitive demand of tasks during implementation
can be challenging for teachers (Brodie, Jina, & Modau, 2009; Stein et al., 2000;
Stigler & Hiebert, 2004). Given the limitations in both their task knowledge and
assessment knowledge, on their own, teachers could unintentionally convert these
tasks into something less desirable during implementation, for example, high-
demanding tasks could be changed into routine exercises or other cognitively low
demanding activities (Stigler & Hiebert, 2004). They could also view these tasks
being problematic for their schedule because of the length of time required for task
planning, selection or design, set up, and implementation. For example, in the case of

69

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

using the authentic intellectual quality framework, analyzing tasks require working
individually and collaboratively with colleagues to arrive at consensus of each
element of the framework. Without appropriate experience in using this framework,
it could be viewed as overwhelming and time consuming.
While recognition and selection of appropriate tasks is crucial, by itself, it is
not enough to ensure meaningful, successful mathematics teaching. As Chapman
(2013b) suggest, mathematical-task knowledge for teaching includes knowledge of
how to orchestrate and organise students work and support their process of thinking
without reducing or eliminating the cognitive challenge. Stein and Smith’s (1998)
Mathematics Task Framework offers a perspective of the complex journey of a task
in the classroom. The framework outlines three different phases that a task passes
through from set up to implementation. First, it appears in curriculum, whether it
be a textbook or elsewhere. Next, it is set up by how the teacher presents it to the
class. Finally, it is implemented by the students. Every step is important to student
learning. This framework based on a mathematical instructional task, “defines a
mathematical task as a classroom activity, the purpose of which is to focus students’
attention on a particular mathematical concept, idea, or skill” (Henningsen & Stein,
1997, p. 582). While we acknowledge the appropriateness of this framework for this
purpose, based on our experience in working with teachers on authentic assessment,
we suggest that it needs to highlight some additional elements when dealing with
authentic assessment tasks.
Figure 2.1 shows the modified Stein and Smith’s (1998) mathematical tasks
framework consisting of three phases and relationships among various task-related
variables and students’ learning outcomes and the elements we added highlighted. The
elements we added make explicit factors regarding the teacher, authentic assessment
tasks, and learning outcomes. For example, in adding authentic intellectual quality in
the first phase and conative processes in the third phase, we argue that mathematics
authentic assessment extends beyond capturing students’ cognitive processes. Its
“aesthetic, utilitarian, and personal value” (Newmann et al., 1996) addresses the
conative dimensions of learning. In adding task design by teacher in the initial phase,
we acknowledge the need for teachers’ collaborative design of mathematics authentic
assessment with their colleagues, which can increase their mathematical knowledge
for teaching, and in turn leads to their effective enactment of assessment tasks that can
serve both learning and accountability purposes. We also consider teacher assessment
literacy and related beliefs/conceptions to be influential factors in setting up the tasks.
The framework recognizes that the characteristics and demands of tasks can
change when moving from instructional material to classroom implementation,
however, it is the quality of the change that is the issue. Teachers with appropriate
mathematical-task knowledge of teaching are more likely to make necessary changes
to maintain the richness of the tasks for students. Thus, approaches to support
teachers development of this knowledge need to take into consideration not only
the teacher’s thinking/knowledge and analysis of features and demands of a task to
determine its appropriateness, but also the students who are to engage in the task

70

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Figure 2.1. Adapted mathematical tasks framework

(such as their abilities, interests, and motivation) and the social context in which
the task is implemented. Thus, the successful journey of the task in the classroom
for instruction or authentic assessment requires knowledge of tasks, students, and
the student-task interaction as well as the ability to implement and adjust tasks in
accordance with the sociocultural context of the classroom.
The preceding discussion points to the importance of supporting teachers’
learning in developing appropriate task knowledge with specific focus on tasks
for authentic assessment. Our work in supporting prospective teachers’ learning of
authentic assessment is at an early stage but suggest the need for helping them to
improve their understanding of authentic assessment task knowledge. Our review of
the mathematics education literature also reveals a scant body of empirical studies
pertaining to building prospective teachers’ capacity in formative assessment, but not
explicitly authentic assessment. There is thus a need for future research to focus on
the design and development of high-quality professional development or intervention
programs to build teachers’ capacity in the design and use/set up of mathematics
authentic assessment in reform-oriented grades K–12 classrooms. Jones and Pepin

71

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

(2016) have underscored the need for engaging teachers as partners in task design,
which not only benefits their professional learning, but also increases the quality of
mathematical tasks. In sum, as Chapman (2013) aptly pointed out “ongoing research
is necessary to determine the nature of mathematical-task knowledge for teaching
in relation to practice and other effective ways to help teachers to develop it” (p. 6).

NOTE
1
This project is supported by an Insight Research Grant from the Social Sciences and Humanities
Research Council of Canada.

REFERENCES
Anderson, L. W., & Krathwohl, D. R. (2001). A taxonomy for learning, teaching, and assessing: A
revision of Bloom’s taxonomy of educational objectives. New York, NY: Longman.
Arbaugh, F., & Brown, C. A. (2005). Analyzing mathematical tasks: A catalyst for change? Journal of
Mathematics Teacher Education, 8(6), 499–536.
Archbald, D., & Newmann, F. M. (1988). Beyond standardized testing: Assessing authentic academic
achievement in the secondary school. Reston, VA: National Association of Secondary School Principals.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Baumert, J., Kunter, M., Blum, W., Brunner, M., Voss, T., Jordan, A., … Tsai, Y.-M. (2010). Teachers’
mathematical knowledge, cognitive activation in the classroom, and student progress. American
Educational Research Journal, 47, 133–180.
Biggs, J. B., & Collis, K. F. (1982). Evaluating the quality of learning – The SOLO taxonomy. New York,
NY: Academic Press.
Black, P. J., & Wiliam, D. (1998). Assessment and classroom learning. Assessment in Education:
Principles, Policy & Practice, 5(1), 7–73.
Boaler, J. (2016). Mathematical mindsets. San Francisco, CA: Jossey-Bass.
Boesen, J., Lithner, J., & Palm, T. (2010). The mathematical reasoning required by national tests and the
reasoning used by students. Educational Studies in Mathematics, 75, 89–105.
Borko, H., Mayfield, V., Marion, S., Flexer, R., & Cumbo, K. (1997). Teachers’ developing ideas and
practices about mathematics performance assessment: Successes, stumbling blocks, and implications
for professional development. Teaching & Teacher Education, 13(3), 259–278.
Boston, M. D. (2006). Developing secondary mathematics teachers’ knowledge of and capacity
to implement instructional tasks with high level cognitive demands. Ann Arbor, MI: ProQuest
Information and Learning Company.
Boston, M. D., & Smith, M. S. (2009). Transforming secondary mathematics teaching: Increasing
the cognitive demands of instructional tasks used in teachers’ classrooms. Journal for Research in
Mathematics Education, 40(2), 119–156.
Brodie, K., Jina, Z., & Modau, S. (2009). Challenges in implementing the new mathematics curriculum
in Grade 10: A case study. African Journal for Research in Mathematics, Science and Technology
Education, 13(1), 19–32.
Chapman, O. (2013a). Engaging children in learner-focused mathematical tasks. In J. Novotná &
H. Moraová (Eds.), Proceedings of the 13th biennial International Symposium for elementary
mathematics teaching (pp. 9–20). Prague, Czech Republic: Charles University.
Chapman, O. (2013b). Mathematical-task knowledge for teaching. Journal of Mathematics Teacher
Education, 16(1), 1–6.
Chapman, O., & Koh, K. (2017). Preservice teachers’ development of knowledge of authentic assessment
mathematics tasks. In E. Galindo & J. Newton (Eds.), Proceedings of the 39th Annual Meeting of the

72

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

North American Chapter of the International Group for the Psychology of mathematics education.
Indianapolis, IN.
Chapman, O., Koh, K., & Piñeiro, J. P. (2019). Prospective elementary teachers’ conceptions of authentic
assessment mathematics tasks. In J. Novotná & H. Moraová (Eds.), Proceedings of the 15th biennial
International Symposium for elementary mathematics teaching. Prague, Czech Republic: Charles
University.
Coffey, D. C. (2000). An investigation into relationships between alternative assessment and pre-service
elementary teachers’ beliefs about mathematics. Ann Arbor, Ml: Bell & Howell Information and
Learning Company.
Cueto, S., Ramírez, C., & León, J. (2006). Opportunities to learn and achievement in mathematics in
a sample of sixth grade students in Lima, Peru. Educational Studies in Mathematics, 62(1), 25–55.
Cumming, J. J., & Maxwell, G. S. (1999). Contextualising authentic assessment. Assessment in Education:
3ULQFLSOHV3ROLF\ 3UDFWLFH  ௅
DeLuca, C., & Klinger, D. A. (2010). Assessment literacy development: Identifying gaps in teacher
candidates’ learning. Assessment in Education: Principles, Policy & Practice, 17  ௅
Estrella, S., Zakaryan, D., Olfos, R., & Espinoza, G. (2019). How teachers learn to maintain the cognitive
demand of tasks through lesson study. Journal of Mathematics Teacher Education, 1–18. (Online
first) https://doi.org/10.1007/s10857-018-09423-y
Fennell, F., Kobett, B. M., & Wray, J. A. (2017). The formative 5: Everyday assessment techniques for
every math classroom. Thousand Oaks, CA: Corwin.
Flewelling, G., & Higginson, W. (2001). A handbook on rich learning tasks. Queen’s University,
Kingston: Queen’s University Centre for Mathematics, Science, and Technology Education.
Foster, C., & Inglis, M. (2017). Teachers’ appraisals of adjectives relating to mathematics tasks.
Educational Study in Mathematics, 95, 283–301. doi:10.1007/s10649-017-9750-y
Frey, B., & Allen, J. (2012). Defining authentic classroom assessment. Practical Assessment, Research
& Evaluation, 17(2), 1–15.
Henningsen, M., & Stein, M. K. (1997). Mathematical tasks and student cognition: Classroom-base
factors that support and inhibit high-level mathematical thinking and reasoning. Journal for Research
in Mathematics Education, 28(5), 524–549.
Hiebert, J., Carpenter, T. P., Fennema, E., Fuson, K., Wearne, D., Murray, H., … Human, P. (1997).
Making sense: Teaching and learning mathematics with understanding. Portsmouth, NH: Heinemann.
Hsu, H.-Y., & Silver, E. A. (2014). Cognitive complexity of mathematical instructional tasks in a Taiwanese
classroom: An examination of task sources. Journal for Research in Mathematics Education, 45(4),
460–496.
Jaworski, B. (2015). Teaching for mathematical thinking: Inquiry in mathematics learning and teaching.
Mathematics Teaching, 248, 28–34.
Kilpatrick, J. (2014). Competency frameworks in mathematics education. In S. Lerman (Ed.),
Encyclopedia of mathematics education (pp. 85–87). Dordrecht: Springer.
Koh, K. (2011a). Improving teachers’ assessment literacy through professional development. Teaching
Education, 22  ௅
Koh, K. (2011b). Improving teachers’ assessment literacy. Singapore: Pearson.
Koh, K. (2014). Teachers’ assessment literacy and student learning in Singapore mathematics classrooms.
In B. Sriraman, J. Cai, K. Lee, L. Fan, Y. Shimizu, C. Lim, & K. Subramaniam (Eds.), The first
sourcebook on Asian research in mathematics education: China, Korea, Singapore, Japan, Malaysia
and India (pp. 981–1010). Charlotte, NC: Information Age Publishing.
Koh, K. (2017). Authentic assessment. In G. W. Noblit (Ed.), Oxford research encyclopedia of education.
Oxford: Oxford University Press.
Koh, K., & Chapman, O. (2018). Developing authentic assessment tasks in mathematics. In K. Philipp, T.
Leuders, & J. Leuders (Eds.), Diagnostic competence in mathematics teacher education. New York,
NY: Springer.
Koh, K., Hadden, J., Parks, C., Monaghan, L., Sanden, L., Gallant, A., & LaFrance, M. (2015). Building
teachers’ capacity in authentic assessment and assessment for learning: A critical inquiry approach.

73

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

In P. Preciado Babb, M. Takeuchi, & J. Lock (Eds.), Proceedings of the IDEAS designing responsive
pedagogy (pp. 43–52), Calgary: Werklund School of Education, University of Calgary.
Koh, K., & Luke, A. (2009). Authentic and conventional assessment in Singapore schools: An empirical
study of teacher assignments and student work. Assessment in Education: Principles, Policy &
Practice, 16  ௅
Lajoie, S. P. (1995). A framework for authentic assessment in mathematics. In T. A. Romberg (Ed.),
Reform in school mathematics and authentic assessment (pp. 19–37). Albany, NY: SUNY Press.
Lee, K. H. (2017). Convergent and divergent thinking in task modification: A case of Korean prospective
mathematics teachers’ exploration. ZDM-The International Journal on Mathematics Education,
49(7), 995–1008.
Leikin, R., & Levav-Waynberg, A. (2009). Development of teachers’ conceptions through learning
and teaching: The meaning and potential of multiple-solution tasks. Canadian Journal of Science,
Mathematics and Technology Education, 9(4), 203–223.
Lesh, R., & Lamon, S. J. (Eds.). (1993). Assessment of authentic performance in school mathematics.
Washington, DC: AAAS.
Levenson, E. (2015). Exploring Ava’s developing sense for tasks that may occasion mathematical
creativity. Journal of Mathematics Teacher Education, 18, 1–25.
Lim, L. (2011). Authentic assessment of lower secondary mathematics. In K. Koh (Ed.), Mastering the
art of authentic assessment: From challenges to champions (Vol. 1, pp. 75–93). Singapore: Pearson.
Lingard, B., Ladwig, J., Mills, M., Bahr, M., Chant, D., & Warry, M. (2001). The Queensland school
reform longitudinal study. Brisbane: Education Queensland.
Marzano, R. J. 1992. A different kind of classroom: Teaching with dimensions of learning. Alexandria,
VA: The Association for Supervision and Curriculum Development.
Mason, J., Burton, L., & Stacey, K. (2010). Thinking mathematically (2nd ed.). Essex: Pearson.
Mason, J., & Johnston-Wilder, S. (2006). Designing and using mathematical tasks. St Albans: Tarquin
Publications.
Mertler, C. A. (2009). Teachers’ assessment knowledge and their perceptions of the impact of classroom
assessment professional development. Improving Schools, 12(2), 101–113.
National Council of Teachers of Mathematics. (1989). Curriculum and evaluation standards for school
mathematics. Reston, VA: National Council of Teachers of Mathematics.
National Council of Teachers of Mathematics. (1991). Professional teaching standards for school
mathematics. Reston, VA: National Council of Teachers of Mathematics.
National Council of Teachers of Mathematics. (2000). Principles and standards for school mathematics.
Reston, VA: National Council of Teachers of Mathematics.
National Council of Teachers of Mathematics. (2010). Why is teaching with problem solving important
for student learning? Research Brief. Reston, VA: National Council of Teachers of Mathematics.
Newmann, F. M., & Archbald, D. A. (1992). The nature of authentic academic achievement. In H. Berlak,
F. M. Newmann, E. Adams, D. A. Archbald, T. Burgess, J. Raven, & T. A. Romberg (Eds.), Towards
a new science of educational testing and assessment (pp. 71–84). Albany, NY: State University of
New York Press.
Newmann, F. M., & Associates. (1996). Authentic achievement: Restructuring schools for intellectual
quality. San Francisco, CA: Jossey-Bass.
Newmann, F. M., Marks, H. M., & Gamoran, A. (1996). Authentic pedagogy and student performance.
American Journal of Education, 104(4), 280–312.
Niss, M. (2015). Mathematical literacy. In S. J. Cho (Ed.), The Proceedings of the 12th International
Congress on mathematical education. Cham: Springer.
Osana, H. P., Lacroix, G. L., Tucker, B. J., & Desrosiers, C. (2006). The role of content knowledge
and problem features on preservice teachers’ appraisal of elementary mathematics tasks. Journal of
Mathematics Teacher Education, 9(4), 347–380.
Pepin, J., & Jones, K. (2016). Research on mathematics as partners in task design. Journal of Mathematics
Teacher Education, 19(2), 105–121.
Pettersen, A., & Braeken, J. (2019). Mathematical competency demands of assessment items: A search
for empirical evidence. International Journal of Science and Mathematics Education, 17(2), 405–442.

74

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BUILDING TEACHERS’ CAPACITY IN MATHEMATICS AUTHENTIC ASSESSMENT

Pettersen, A., & Nortvedt, G. A. (2018). Identifying competency demands in mathematical tasks: Recognising
what matters. International Journal of Science and Mathematics Education, 16(5), 949–965.
Polly, D., Neale, H., & Pugalee, D. K. (2014). How does ongoing task-focused mathematics professional
GHYHORSPHQWLQÀXHQFHHOHPHQWDU\VFKRROWHDFKHUV¶NQRZOHGJHEHOLHIVDQGHQDFWHGSHGDJRJLHV"Early
Childhood Education Journal, 42, 1–10. doi:10.1007/s10643-013-0585-6
Popham, W. J. (2011). Assessment literacy overlooked: A teacher educator’s confession. The Teacher
Educator, 46(4), 265–273.
Resnick, L. B. (2006). Do the math: Cognitive demand makes a difference. Research Points Essential
Information for Education Policy, 4(2), 1–4.
Romberg, T. A. (Ed.). (1995). Reform in school mathematics and authentic assessment. Albany, NY:
SUNY Press.
Romberg, T. A., & Wilson, L. D. (1995). Issues related to the development of an authentic assessment
system for school mathematics. In T. A. Romberg (Ed.), Reform in school mathematics and authentic
assessment (pp. 1–18). Albany, NY: SUNY Press.
Santos, E. R., & Teixeira, B. R. (2018). Learning assessment in preservice teacher who teaches
mathematics education. Research, Society and Development, 7(3), 1–17.
Scheurman, G., & Newman, F. M. (1998). Authentic intellectual work in social studies: Putting
performance before pedagogy. Social Education, 62(1), 23–25.
Schoenfeld, A. H. (1994). Reflections on doing and teaching mathematics. In A. H. Schoenfeld (Ed.),
Mathematical thinking and problem solving (pp. 53–70). Hillsdale, NJ: Erlbaum.
Shepard, L. A. (2000). The role of assessment in a learning culture. Educational Researcher, 29  í
Shepard, L., Hammerness, K., Darling-Hammond, L., Rust, F., Baratz Snowden, J., Gordon, E., &
Pacheco, A. (2005). In L. Darling-Hammond & J. Bransford (Eds.), Preparing teachers for a changing
world (pp. 275–326). San Francisco, CA: John Wiley.
Shimizu, Y., Kaur, B., Huang, R., & Clarke, D. (Eds.). (2010). Mathematical tasks in classrooms around
the world. Rotterdam, The Netherlands: Sense Publishers.
Silver, E. A., Mesa, V. M., Morris, K. A., Star, J. R., & Benken, B. M. (2009). Teaching mathematics for
understanding: An analysis of lessons submitted by teachers seeking NBPTS certification. American
Educational Research Journal, 46(2), 501–531.
Silver, E. A., & Mills, L. M. (2018). A fresh look at formative assessment in mathematics teaching.
Reston, VA: National Council of Teachers of Mathematics.
Smith, M. S., & Stein, M. K. (1998). Reflections on practice: Selecting and creating mathematical tasks:
From research to practice. Mathematics Teaching in the Middle School, 3(5), 344–350.
Stein, M., Smith, M., Henningsen, M., & Silver, E. (2000). Implementing standards-based mathematics
instruction: A casebook for professional development. New York, NY: Teacher College Press.
Stein, M. K., Grover, B. W., & Henningsen, M. (1996). Building student capacity for mathematical
thinking and reasoning: An analysis of mathematical tasks used in reform classrooms. American
Educational Research Journal, 33(2), 455–488.
Stein, M. K., & Lane, S. (1996). Instructional tasks and the development of student capacity to think and
reason: An analysis of the relationship between teaching and learning in a reform mathematics project.
Educational Research and Evaluation, 2(1), 50–80.
Stein, M. K., & Smith, M. S. (1998). Mathematical tasks as a framework for reflection: From research to
practice. Mathematics Teaching in the Middle School, 3(4), 268–275.
Stein, M. K., Smith, M. S., Henningsen, M., & Silver, E. A. (2000). Implementing standards-based mathematics
instruction: A casebook for professional development. New York, NY: Teachers College Press.
Stigler, J. W., & Hiebert, J. (2004). Improving mathematics teaching. Educational Leadership, 61(5), 12–17.
Sullivan, P., Clarke, D., & Clarke, B. (2013). Teaching with tasks for effective mathematics learning. New
York, NY: Springer.
Swan, M. (2007). The impact of task-based professional development on teachers’ practices and beliefs:
A design research study. Journal of Mathematics Teacher Education, 10(4–6), 217–237.
Swan, M. (2008). Designing a multiple representation learning experience in secondary algebra. Educational
Designer, 1(1). Retrieved from http://www.educationaldesigner.org/ed/volume1/ issue1/article3

75

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM KOH & OLIVE CHAPMAN

Turner, R., Blum, W., & Niss, M. (2015). Using competencies to explain mathematical item demand:
A work in progress. In K. Stacey & R. Turner (Eds.), Assessing mathematical literacy: The PISA
experience (pp. 85–115). New York, NY: Springer.
Wallace, M. (2012). Developing assessment practices: A study of the experiences of preservice
mathematics teachers as learners and the evolution of their assessment practices as educators. Ann
Arbor, MI: ProQuest LLC.
Watson, A., & Mason, J. (2007). Taken-as-shared: A review of common assumptions about mathematical
tasks in teacher education. Journal of Mathematics Teacher Education, 10(4–6), 205–215.
Watson, A., & Ohtani, M. (Eds.). (2015). Task design in mathematics education: An ICMI study 22.
Heidelberg: Springer.
Watson, A., & Sullivan, P. (2008). Teachers learning about tasks and lessons. In D. Tirosh & T. Wood
(Eds.), Tools and resources in mathematics teacher education (pp. 109–135). Rotterdam, The
Netherlands: Sense Publishers.
Watt, H. (2005). Attitudes to the use of alternative assessment methods in mathematics: A study with
secondary mathematics teachers in Sydney, Australia. Educational Studies in Mathematics, 58, 21–44.
Webb, D. C. (2009). Designing professional development for assessment. Educational Designer, 1(2),
௅5HWULHYHGIURPKWWSZZZHGXFDWLRQDOGHVLJQHURUJHGYROXPHLVVXHDUWLFOH
Wiggins, G. (1989). A true test: Toward more authentic and equitable assessment. Phi Delta Kappan,
  ௅
Wiggins, G. (1990). The case for authentic assessment. Practical Assessment, Research & Evaluation,
2(2), 1–3.
Wilhelm, A. G. (2014). Mathematics teachers’ enactment of cognitively demanding tasks: Investigating
links to teachers’ knowledge and conceptions. Journal for Research in Mathematics Education, 45(5),
636–674.
Wiliam, D. (2007). Five “key strategies” for effective formative assessment. Research brief. Reston, VA:
The National Council of Teachers of Mathematics.
Wyatt-Smith, C., & Gunn, S. (2009). Towards theorising assessment as critical inquiry. In J. J. Cumming
& C. Wyatt-Smith (Eds.), Educational assessment in the 21st century: Connecting theory and practice
SS௅ /RQGRQ6SULQJHU
Zaslavsky, O., & Sullivan, P. (Eds.). (2010). Constructing knowledge for teaching secondary mathematics:
Tasks to enhance prospective and practicing teacher learning. New York, NY: Springer.

Kim Koh
Werklund School of Education
University of Calgary

Olive Chapman
Werklund School of Education
University of Calgary

76

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI, JEONGSUK PANG, HUIRONG ZHANG AND
NAIQING SONG

3. MATHEMATICS CONCEPTUAL KNOWLEDGE


FOR TEACHING
Helping Prospective Teachers Know Mathematics
Well Enough for Teaching

What should prospective teachers know and be able to do to be ready for their
professional career in mathematics teaching? This is not a trivial question, but it
is a crucial one for all of those who are responsible for teachers’ preparation in
mathematics. In this chapter, we conceptualize Mathematics Conceptual Knowledge
for Teaching as the core of prospective teachers’ professional competency that
can and should be developed in teacher preparation programs. Specifications of
mathematics conceptual knowledge for teaching are discussed, and examples on the
content topic of fraction division are provided to illustrate different components of
mathematics conceptual knowledge for teaching. Prospective elementary teachers’
performance variations across different components of mathematics conceptual
knowledge for teaching in fraction division are examined in the cases of China and
South Korea, respectively.

INTRODUCTION

What teachers need to know in mathematics for teaching is not a trivial question.
On one hand, it seems that students who just finished eighth grade can ‘teach’ (or
talk about) eighth-grade mathematics. If taking this assumption for granted, the
value of teacher preparation and offering advanced content courses in mathematics
becomes questionable. In reality, we know this is not the case. Teachers are expected
not just to present and state content that needs to be taught, but also to be able to
help students understand what needs to be learned, which includes helping students
make mathematical connections across different content topics and answering
students’ various questions. However, acknowledging that teachers need to know
more than their students does not specify what mathematics teachers need to know
more than students for teaching. If taking teacher preparation curriculum as an
indicator of what teachers need to know, on the other hand, existing studies already
documented great variations in mathematics teacher preparation programs both
across and within education systems (e.g., Li, Huang, & Shin, 2008; Li, Ma, &

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_004

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

Pang, 2008; Schmidt et al., 2007). Although it remains unclear what curriculum
structure or models may best prepare mathematics teachers, it is important for us
to examine and understand how teacher knowledge of mathematics differs from
student knowledge.
Efforts to examine the nature and structure of teachers’ mathematics knowledge
needed for teaching have been taking different approaches. For example, Ball,
Thames and Phelps (2008) built on Shulman’s (1986) notion of pedagogical content
knowledge to develop a practice-based theory of mathematical knowledge for teaching
through examining teachers’ classroom instruction. The analysis led the researchers
to discuss and specify teachers’ pedagogical content knowledge as containing
knowledge of content and students and knowledge of content and teaching, and an
important specific content knowledge, specialized content knowledge. A group of
researchers at the University of Cambridge developed a theoretical framework, the
Knowledge Quartet (Rowland, 2013), through analyzing videotaped mathematics
lessons taught by prospective elementary teachers.
Likewise, by focusing on prospective teachers and building upon existing studies
on teachers’ mathematics knowledge, Li and Kulm (2008) built upon Shulman’s
work (1986) and Ball and her colleagues’ work (Ball et al., 2008) to put forward
a framework to outline five knowledge components of prospective teachers’
mathematics knowledge needed for teaching, consisting of common content
knowledge, specialized content knowledge, knowledge of content and students,
knowledge of content and teaching, and knowledge of mathematics curriculum. Such
work has paid close attention to classifying different knowledge components into
types, which provides important support to different course offerings in a program
that can help prospective teachers to develop different knowledge components.
However, it remains unclear how different knowledge components might actually
work together and how teachers’ learning of different (and separate) knowledge
components might help them turn knowledge acquisition into teaching competence.
In fact, making possible connections among different knowledge components
becomes a great challenge when different knowledge components are acquired
through courses often offered by different departments. For example, when
prospective teachers took content courses from mathematics department, they likely
learned least common denominator and different number base systems (e.g., base-
2, base-5) in addition to base-10 number system. When they come to education
department or college, they likely learn about curriculum standards, lesson plan
development, and assessment, but not connect specifically to the content topics
of least common denominator and different number base systems learned in the
mathematics department. Thus, prospective teachers are left to merge together those
different knowledge components by themselves when they go to teach specific
content topics in mathematics classrooms. Because mathematics teaching and
learning in classrooms are content-topic based, current course offerings and structure
in many teacher preparation programs present a gap between teacher knowledge
preparation and what teachers actually need to know and be able to do in classrooms.

78

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

In this chapter, we aim to propose a different conception about what teachers need
to know and how their knowledge of mathematics differs from student knowledge,
that is, a topic-based knowledge package called Mathematics Conceptual Knowledge
for Teaching (MCKT). The conception and specific components of mathematics
conceptual knowledge for teaching will be described and discussed in the follow-
up sections. To illustrate how prospective teachers’ knowledge of mathematics
conceptual knowledge for teaching can be differentiated, an instrument will also be
developed and used to collect and analyze data from prospective teachers.
This chapter builds upon a previous study on mathematical preparation for
prospective elementary school teachers in several selected education systems in
East Asia, including Hong Kong, Japan, Mainland China, Singapore, South Korea,
and Taiwan (Li et al., 2008). It was documented that mathematics education for
prospective elementary school teachers was emphasized in the selected education
systems in East Asia. In particular, extensive data were collected from prospective
elementary school teachers and teacher educators sampled in both Mainland China
and South Korea. The results from sampled prospective teachers indicated that they
had strong education in mathematics, which echoed our belief that mathematics
education is fundamental to what teachers need to know more than students for
teaching. However, it remains to be understood whether strong preparation in
mathematics is good enough for what we believe teachers need to know more than
students for teaching. Thus, we plan to again take the cases of Mainland China and
South Korea for examining their mathematics conceptual knowledge for teaching in
this study.
The following sections are organized in four parts. In the first part, we provide an
overview of related research, the conception of mathematics conceptual knowledge
for teaching and the survey developed for this study. Secondly, we report and discuss
sampled prospective elementary school teachers’ knowledge in school mathematics,
structured as mathematics conceptual knowledge for teaching, using the data
collected from Mainland China. Thirdly, we report and discuss sampled prospective
elementary school teachers’ knowledge in school mathematics, structured as
mathematics conceptual knowledge for teaching, using the data collected from
South Korea. In the last part, we summarize the results obtained from Mainland
China and South Korea and discuss possible research in the future and implications
for mathematical preparation of elementary school teachers in other educational
systems.

RESEARCH BACKGROUND AND THE CONCEPTION OF MATHEMATICS


CONCEPTUAL KNOWLEDGE FOR TEACHING

Existing research has generally documented the importance of knowledge in


expertise acquisition and development in knowledge-rich and complex domains,
including mathematics instruction (Li & Kaiser, 2011). The importance of
knowledge in teachers’ expertise also goes beyond a quantity measure to include

79

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

knowledge structure with certain depth. Such a knowledge-based characterization


of teachers’ expertise is commonly used in large-scale international studies,
for example, the Teacher Education and Development Study in Mathematics
(TEDS-M, see Blömeke, Hsieh, Kaiser, & Schmidt, 2014; Döhrmann, Kaiser, &
Blömeke, 2018), which measures three types of knowledge: mathematics content
knowledge, mathematics pedagogical content knowledge, and general pedagogical
knowledge.
Knowing the existence of different knowledge components is important in teacher
preparation, as the program needs to be structured in certain ways for teachers’
knowledge acquisition. However, if salient connections among different knowledge
components are left unspecified in their knowledge preparation, prospective teachers
would be left to make such connections by themselves after learning separate pieces
of knowledge. In fact, this is often the case with preparation in mathematics, since
mathematics content courses are typically offered by a mathematics department and
pedagogy courses are delivered in an education department. These two departments
may communicate little, if at all, about the content and instruction of these courses for
the same group of prospective teachers. Checking whether prospective teachers are
ready or not for their professional career in teaching often results in course counting
rather than examining what is offered in these courses and how various topics being
covered can and should be connected. To be ready for their professional career
in mathematics teaching, as we indicated at the beginning, prospective teachers
should be expected not just to present and state what (content) needs to be taught in
classrooms, but to be able to help students develop conceptual understanding and
make content connections across different topics and representations. Such features
of teacher expertise require packages of topic-based integrated knowledge (e.g., Ma,
1999), rather than a collection of separate knowledge pieces.
The notion of knowledge package is well illustrated and discussed in Ma’s (1999)
book. Teacher Chen in Ma’s study discussed about his perception of knowledge
package that teachers have:
There is not a firm, rigid, or single right way to “pack” knowledge. It is all up
to one’s own viewpoint. Different teachers, in different contexts, or the same
teacher with different students, may “pack” knowledge in different ways. But
the point is that you should see a knowledge “package” when you are teaching
a piece of knowledge. And you should know the role of the present knowledge
in that package. You have to know that the knowledge you are teaching is
supported by which ideas or procedures, so your teaching is going to rely on,
reinforce, and elaborate the learning of these ideas. (p. 18)
To Teacher Chen, what teachers know about mathematics is not knowledge in
pieces but in groups. He argued that specific knowledge package can have some
variations across different teachers and specific groups of students they teach. Such a
notion suggests that a knowledge package, in Teacher Chen’s view, can have certain
flexibility in the action of teaching. Such a knowledge package is only available

80

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

in Chinese teachers’ mind, but not in the American teachers’ in Ma’s study. Across
different Chinese teachers, Ma’s (1999) analyses highlight that these teachers’
knowledge packages share similarity in terms of the principles of how to ‘pack’ the
knowledge and what are the ‘key’ pieces. In particular, Ma presents three knowledge
package models derived from her data analyses: subtraction with regrouping (p. 19),
multidigit multiplication (p. 47), and division by fraction (p. 77).
To be able to help students learn mathematics with understanding, we believe
that teachers need to have structured and coherent knowledge construct called:
Mathematics Conceptual Knowledge for Teaching (MCKT). The notion of
mathematics conceptual knowledge for teaching is consistent with Ma’s notion
of knowledge packages. Specifically, by mathematics conceptual knowledge for
teaching, we mean topic-based conceptual knowledge packages that are needed for
understanding, explaining, as well as teaching specific mathematics content topics
with connections. Building upon Ma’s notion of knowledge package that focuses on
mathematics content aspect, we propose that mathematics conceptual knowledge for
teaching can be specified as containing the following three topic-based knowledge
components that can and should be offered in the same course:
‡ Having knowledge and skills directly associated with a specific content topic;
‡ Being able to connect and justify the main points of a content topic, and to place
it in wider contexts;
‡ Knowing and being able to use various representations for teaching the content
topic and being able to teach the relations between them.
Specifically, the knowledge component (1) refers to common content knowledge
that students are also expected to learn. Knowledge component (2) goes beyond the
content topic itself to place the content topic in a broader knowledge structure. The
combination of these two knowledge components is similar to topic-based knowledge
packages as specified in Ma’s book. For example, prospective teachers should learn
well about division of whole numbers both procedurally and conceptually. They
should be able to do and explain 12 ÷ 2, as dividing 12 into 2 pieces and finding the
value of one piece or multiplying 12 by ½. If prospective teachers can’t explain why
12 ÷ 0 is meaningless, then their knowledge component (2) is limited for this topic of
division of whole numbers. More often than not, (2) is presented as great challenges
to prospective and practicing teachers as it often requires conceptual understanding
across different content topics that are connected. Knowledge component (3) refers
to mainly pedagogical aspects of teaching this topic as presented in knowledge
components (1) and (2). Across these three knowledge components of mathematics
conceptual knowledge for teaching, a likely sequence of familiarity order for
prospective teachers is (1), (3) and (2).
With mathematics conceptual knowledge for teaching, we emphasize the
depth and systematic view of mathematics knowledge with associated pedagogy
that can empower teachers for further expertise development in the future. After
prospective teachers complete their teacher education program, they should have

81

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

many more opportunities for developing knowledge about students’ learning of


different mathematics content topics through working with and learning from their
own students, but opportunities to develop mathematics conceptual knowledge for
teaching will not appear automatically.

How Mathematics Conceptual Knowledge for Teaching May Differ from or


Connect with Other Knowledge Constructs in Literature

In order to clarify possible differences and connections with other knowledge


constructs that have been discussed in literature (e.g., Petrou & Goulding, 2011;
Döhrmann et al., 2018), we agree with Ruthven (2011) in that such comparison of
categories from different models is no less problem than the distinction between
categories within any one. Nevertheless, we hope such discussions can help further
clarify the concept of mathematics conceptual knowledge for teaching itself.
First of all, mathematics conceptual knowledge for teaching relates to the notions
of subject matter knowledge and pedagogical content knowledge as discussed by
Shulman (1986) but emphasizes the connections between these two knowledge
components in the same topic-based package. Specifically, the first and second
components of mathematics conceptual knowledge for teaching are mainly subject
matter knowledge, and the third component of mathematics conceptual knowledge
for teaching is mainly pedagogical content knowledge. At the same time, mathematics
conceptual knowledge for teaching differs from knowledge specifications provided
by Shulman. We emphasize that teachers’ knowledge is not by pieces that can be
treated separately as subject matter knowledge or pedagogical content knowledge
but topic-based packages. The combination of such subject matter knowledge and
pedagogical content knowledge in the topic-based knowledge package makes it
more practical for teachers if they acquire and use such knowledge in teaching.
At the same time, the notion of mathematics knowledge for teaching is coined
and used by Ball and her colleagues (2008), as further developed from Shulman’s
knowledge constructs through analyzing teachers’ classroom instruction.
Specifically, Ball et al. specify Shuman’s pedagogical content knowledge as
containing knowledge of content and students, knowledge of content and teaching,
and knowledge of curriculum; Shulman’s subject matter knowledge as including
common content knowledge, specialized content knowledge and horizon knowledge.
With these knowledge constructs, it is clear that mathematics conceptual knowledge
for teaching is very different from Ball et al.’s knowledge constructs. The notion
of mathematics conceptual knowledge for teaching tends to simplify knowledge
constructs, other than to expand knowledge constructs as Ball and her colleagues
try to do. Moreover, mathematics conceptual knowledge for teaching emphasizes
a structure of topic-based knowledge package that contains inter-connected subject
matter knowledge and pedagogical content knowledge, and Ball, et al. aim to
differentiate and specify the differences across these knowledge sub-constructs
under subject matter knowledge and pedagogical content knowledge, respectively.

82

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

Likewise, the notion of mathematics conceptual knowledge for teaching


differs from teachers’ knowledge constructs used in large-scale assessments like
TEDS-M (Döhrmann et al., 2018). The three types of knowledge assessed in
TEDS-M (mathematics content knowledge), mathematics pedagogical content
knowledge, and general pedagogical knowledge highlight three self-contained
knowledge constructs, but not necessarily connections across different knowledge
constructs.
The Knowledge Quartet is a theoretical framework developed by members of the
Faculty of Education at the University of Cambridge (Rowland, 2013) discussed
also in this volume of the Handbook. Building upon Shulman’s subject matter
knowledge and pedagogical content knowledge constructs, the team was interested
in investigating the relationship between subject matter knowledge and pedagogical
content knowledge in action through observing mathematics lessons taught by
prospective teachers. The Knowledge Quartet contains four dimensions: Foundation,
Transformation, Connection and Contingency. Each dimension is composed of a
number of key components and contributory codes. As examples, the key components
of the Foundation dimension include: knowledge and understanding of mathematics
per se and of mathematics-specific pedagogy, beliefs concerning the nature of
mathematics, the purposes of mathematics education, and the conditions under
which students will best learn mathematics. The key components of the Connection
dimension include: the sequencing of material for instruction, and an awareness of
the relative cognitive demands of different topics and tasks (p. 25). It is clear that
the Knowledge Quartet has been developed as containing enough components and
codes, so that teachers’ different actions in mathematics classrooms can be well
covered for the purposes of teaching observation and assessment. Different from
the Knowledge Quartet, the notion and components of mathematics conceptual
knowledge for teaching are not developed for assessing mathematics classroom
teaching, but instead for conceptualizing what teachers’ subject matter knowledge-
oriented and pedagogical content knowledge-oriented knowledge components are
needed when teaching specific content topics.
It should be pointed out that subject matter knowledge and pedagogical content
knowledge constructs developed by Shulman (1986) and used in large-scale
assessment (e.g., Döhrmann et al., 2018) are consistent with the structure of course
offerings in many teacher preparation programs, that is, subject matter knowledge
and pedagogical content knowledge can be separated and obtained through different
courses from different departments. In contrast, mathematics conceptual knowledge
for teaching is developed to provide the opposite view, that is, mathematics
conceptual knowledge for teaching envisions the need and importance of combining
and offering subject matter knowledge-oriented and pedagogical content knowledge-
oriented knowledge components in the same course rather than through different
courses. Mathematics conceptual knowledge for teaching aims to promote topic-
based mathematics content and related pedagogical approaches for prospective
teacher preparation and teacher professional development.

83

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

Clearly, specific mathematics conceptual knowledge for teaching varies


from one content topic to another, as illustrated in Ma’s (1999) three topic-based
knowledge packages. The task of specifying the mathematics conceptual knowledge
for teaching is needed but enormous for different content topics. Nevertheless,
teachers’ acquisition of the mathematics conceptual knowledge for teaching would
provide them a base to further develop a profound understanding of fundamental
mathematics they teach as termed by Ma (1999). Given the dramatic variations
across mathematical content topics, we will focus on the mathematics conceptual
knowledge for teaching that teachers would need to have (more than students) for
teaching fraction division as an example. In the follow-up sections, we will illustrate
the three components of the mathematics conceptual knowledge for teaching
fraction division, and then develop an instrument for assessing prospective teachers’
mathematics conceptual knowledge for teaching of fraction division.

The Case: Mathematics Conceptual Knowledge for Teaching Fraction Division

The topic of fraction division is difficult in school mathematics not only for students
(Li, 2008), but also for prospective teachers (Borko et al., 1992; Li & Kulm, 2008;
Simon, 1993). Mathematically, fraction division can be presented as an algorithmic
procedure that can be easily taught and learned as “invert and multiply.” However,
the topic is conceptually rich and difficult, as its meaning requires explanation
through connections with other mathematical knowledge, various representations,
or real-world contexts (Greer, 1992; Li, 2008). The selection of the topic of fraction
division, as a special case, can provide a rich context for exploring possible depth
and limitations in prospective teachers’ knowledge in mathematics and pedagogy.
Before discussing and specifying what knowledge teachers would need to have
about fraction division, it is important to know what knowledge teachers may
actually have. In fact, existing studies already examined prospective and practicing
teachers’ mathematics knowledge on fraction division within and across educational
systems (e.g., Li & Huang, 2008; Li & Kulm, 2008; Li et al., 2008).
In a study on United States prospective teachers’ knowledge and confidence
(Li & Kulm, 2008), it was to our surprise how sampled United States prospective
middle school teachers responded to the question whether the division of fractions
(i.e., ) works. Only one out of 46 prospective teachers said this works, the
rest gave all other kinds of answers. What was even more surprising is that several
prospective teachers indicated that the division of fractions should be done with a
procedure of “KFC.” Even if one is not a fast-food goer, you certainly know what
“KFC” (i.e., Kentucky Fried Chicken) stands for. At the same time, however, one
can quickly realize that what everyone knows about “KFC” is different from what
these sampled prospective teachers wanted to convey. It was interesting that few
participants wrote a note next to the term “KFC” (Keep the first, Flip the second, and
Change the computation). After talking to the course instructors for the mathematics

84

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

content and methods courses, Li and Kulm (2008) found that the use of “KFC”
is something that some prospective teachers were taught and thus remembered as
school students. It became clear that the use of “KFC” is effective in helping students
(now prospective teachers) memorize the computational procedure for the division
of fractions. It is also safe to infer that these soon-to-graduate prospective teachers
will very likely teach their future students to learn and remember the division-of-
fraction algorithm with the proven effective use of “KFC.” This raises the question
of what these prospective teachers should learn about fraction division, beyond
procedural knowledge mentioned above, for teaching.
Nowadays, it has become a common understanding that school students’ learning
of computational procedure (e.g., fraction division) is not enough. As the expectation
for students’ learning of mathematics is enhanced, the same goes for teachers. The
expectation of developing students’ conceptual understanding of mathematics, as
emphasized by the National Council of Teachers of Mathematics (NCTM) (e.g.,
NCTM, 2000), requires teachers to know and teach more than computational
procedures (e.g., “invert-and-multiply” for fraction division). Although various
instructional approaches have been created and shared for teaching the difficult
topic of fraction division (Li, 2008), the creation of various instructional approaches
does not entail what knowledge teachers would need to have for teaching fraction
division. Our conception of mathematics conceptual knowledge for teaching would
suggest the following components of topic-based knowledge on fraction division.

Knowledge component (1): Having knowledge and skills about fraction


division. Fraction division can easily be taken as a procedural skill, and students
can be expected to know how to do fraction division. As we discussed in the last
section, students’ learning of fraction division as a computational procedure is not
enough. The same expectation goes for teachers. Teachers should know that the
meaning of fraction division is the same as the meaning of division with whole
numbers, that is, it is an inverse operation of multiplication. In fact, teachers’
understanding of the meaning of fraction division can and shall go even further.
Teachers’ knowledge of fraction division should enable them to understand how
the division-of-fractions computation can be derived and to use fraction division in
solving and posing problems.
For example, Tirosh (2000) showed how the division-of-fractions presented as
the inverse operation of multiplication can be derived:
if , then .

Thus, .

In fact, prospective elementary school teachers in some other systems, like Mainland
China, were once expected to also know that such a value of X = should be the

85

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

one and only solution (Li, 2002). Basically, prospective teachers would need to learn
how to prove its uniqueness as below:
Assuming there are two different fractions (i.e., and ) resulted from , then we
have and .
Therefore,

, thus (1)

Because thus
Therefore, we can have from (1), then . This result is contradictory to
the assumption at the beginning. Therefore, the original assumption cannot be true.
In other words, the quotient of is unique (i.e., ).

Alternatively, the computation of fraction division can also be derived from the
definitions of fraction equivalence and fraction division as follow:
Definition (a): if and only if ad = bc;
Definition (b): if and only if ;

Hence, , thus ayd = bxc based on definition (a).

Therefore, EDVHGRQGH¿QLWLRQ D DJDLQ


Knowledge component (2): Mathematical connections and of main points related
to fraction division. Fraction division is often memorized as “invert and multiply,”
as this is the most commonly taught algorithm. However, division-of-fractions
computation can be carried out in different ways (e.g., Elashhab, 1978; Tirosh, 2000),
as it has close connections with several other important concepts and procedures.
For example, the following are three common procedures:

Invert-and-Multiply. . For example,

Complex fraction. .

For example,

Common denominator. .

For example,

86

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

Mathematically, each algorithm presents a different way for computing division-


of-fractions, as different concepts and procedures are employed in the process of
carrying out the computation. At the same time, they are also connected to each
other. In particular, as the invert-and-multiply algorithm is the only one that is
given without explanation, the procedures shown above for ‘complex fraction’ and
‘common denominator’ can be taken as ways for justifying the invert-and-multiply
algorithm. Moreover, the invert-and-multiply algorithm can also be justified in ways
with some modifications as shown below.

, or

, or

The fraction division computation specified in the above algorithms not only
shows differences in their computation procedures, but also provides a mathematical
foundation for creating and understanding various representation models for dividing
fractions. Teachers’ understanding of different algorithms/concepts and their
connections is essential in the conceptual knowledge for teaching fraction division.

Knowledge component (3): Representational variations and connections for teaching


fraction division. Going beyond basic knowledge and skills associated with fraction
division, teachers also need to be able to create and justify possible explanatory
models for dividing fractions. For example, the multiplication of fractions can be
explained with an area model, where the product can be represented as the area of
a rectangle created with the multiplicand and multiplier as its two sides. Fraction
division (as the inverse operation of multiplication) can also be explained with an area
model (e.g., ). The area model for fraction division computation can be shown
as a process of building a new rectangle where its area (here, ) and the length of one
side (here, ) are known. And the task becomes to find the length of its other side.
Clearly, the area model is developed and explained in terms of the inverse
relationship between fraction multiplication and fraction division. There are several
other models that can be used to explain how fraction division can be carried out
(e.g., Li, 2008), especially when the value of divisor is smaller than the value of
dividend. The selection and use of different models would require teachers to
develop solid understanding of the underlying mathematical ideas for different
models and approaches. Knowing different algorithms discussed above provides
a basis for developing and understanding various representations and models for
dividing fractions.

87

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

The model development for dividing fractions is often associated with the case
for whole number division. Thus, the commonly used measurement and partitive
models for whole number division can also be extended to the case for fractions
(Gregg & Gregg, 2007). As examples, Figure 3.1 shows three different models for
÷ . The model (C) is related to the partitive model for whole number division,
and the other two are related to the measurement model for whole number division.
As pointed out by Gregg and Gregg (2007), the measurement model is tied to
the common denominator algorithm and the partitive model relates to the invert-
and-multiply algorithm. They can all be used for developing and understanding the
invert-and-multiply algorithm.
Likewise, Figure 3.2 shows another possible model for explaining how fraction
division can be done. In fact, this model is basically developed with the ideas from
the complex fraction division algorithm.
The connections between these sample models and different algorithms suggest
the importance for teachers to have in-depth mathematical understanding of these
algorithms. Without an in-depth understanding of different algorithms for fraction
division, teachers can easily get overwhelmed with various explanatory models and
will not be able to identify and discuss their differences and connections. It is the
depth of teachers’ own understanding that will enable them to select and structure
the teaching of fraction division effectively.

A few more remarks. Taken together, mathematics conceptual knowledge for


teaching is not just pieces of knowledge that can be delineated but connected. It is the
connections that teachers and students can build among different pieces of knowledge
that will help them make a better sense of mathematics. No exceptions for teachers, it
also takes time for them to develop and improve mathematics conceptual knowledge
for teaching, here on fraction division. If teachers have a clear sense of what they need
to have in order to facilitate students’ learning with conceptual understanding, teachers
shall be on the right track in learning through teaching and for teaching.
Now, let us revisit the story that was shared at the beginning of this section in
studying a group of United States prospective teachers’ responses to the question
whether the division of fractions (i.e., ) works. On the surface, what has

been discussed so far has no direct connection with . The question then
becomes whether we can assume that prospective teachers in that study simply did
not know this algorithm, as they might never get a chance to learn this very specific
content. Because this is very likely the case, the assumption is reasonable to a certain
degree.
At the same time, their unsatisfactory performance also suggests that those
sampled prospective teachers might not learn mathematics conceptual knowledge
for teaching fraction division. In fact, it is not possible to teach prospective teachers
every bit of detailed knowledge about a specific content topic (here, fraction

88

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

A The three shaded areas in the top bar represents the value of
The one shaded area in the bottom bar represents the value of
Thus, can be measured off with three times, or three
can fit into

B The left half shaded area represents , which has 6 shaded cells.
The top one horizontal shaded area represents , which has 2
shaded cells.
Thus, ÷ = 6 cells ÷ 2 cells = 3
Or, you can use 2 shaded cells (i.e., ) to measure 6 shaded
cells (i.e., ). You will get the answer 3.

C ÷ as a problem to find ‘how many kilograms of honey can you store in a full-size
container, if kilogram of honey can fill size of the container?’

Figure 3.1. Three possible models for ÷

division). It is the fundamental idea to teach prospective teachers the mathematics


conceptual knowledge for teaching, so that they can develop a better understanding
of the topics they will teach. If those prospective teachers learned mathematics
conceptual knowledge for teaching, it is then possible to expect that they may be
able to justify as

89

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

Expanding the divisor ( ) 6 times, so that it becomes 1. At the same time, expanding the
dividend ( ) also 6 times, then it becomes 3 wholes. So the answer is 3.

Figure 3.2. A model that differs from (A) in Figure 3.1 and connects with the complex
fraction division algorithm

Teachers may even discover that this algorithm, once combined with the concept
of complex fraction, can lead to

A Survey of Prospective Elementary Teachers’ Mathematics Conceptual Knowledge


for Teaching Fraction Division

Li et al. (2008) reported that sampled prospective elementary school teachers in


Mainland China and South Korea performed well in mathematics (specifically in
fraction division). It remains unclear if prospective elementary school teachers may
have balanced development across different knowledge components of mathematics
conceptual knowledge for teaching. To gain a better understanding of elementary
teachers’ mathematics conceptual knowledge for teaching in various components,
we thus developed and conducted a survey of prospective elementary school teachers
sampled in Mainland China and South Korea with a focus on fraction division.
The results are expected to provide empirical evidence to help illustrate if the
mathematics conceptual knowledge for teaching construct is feasible to differentiate
such knowledge components.
The whole survey contains two main parts with three items for Part 1 and seven
items for Part 2. Part 1 contains items on elementary school teachers’ knowledge
of elementary mathematics curriculum and their confidence in their readiness for
mathematics instruction. Part 2 has seven main items that assess elementary school
teachers’ three knowledge components of mathematics conceptual knowledge for
teaching on the topic of fraction division. Most items were taken from previous
studies (Li & Smith, 2007; Li et al., 2008), with some items adapted from school

90

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

mathematics textbooks and others’ studies (e.g., Hill, Schilling, & Ball, 2004; Tirosh,
2000). The research team reviewed the selection of these items for this survey. Given
the limited page space we have here, only three items (note: each item containing
two sub-questions) from Part 2 and elementary school teachers’ responses to these
items are included for analyses to provide a glimpse of sampled elementary school
teachers’ mathematics conceptual knowledge for teaching. Further analysis of the
entire data will be needed for detailed reporting of sampled elementary school
teachers’ performance both within and across Mainland China and South Korea.
The three items analyzed and reported in this chapter are:
Item 5 (designed to assess knowledge component (2) of mathematics conceptual
knowledge for teaching). Explain the meaning of fraction division, and how
fraction division relates to other content topics.
Item 6 (designed to assess knowledge component (1) of mathematics conceptual
knowledge for teaching). Solve the following problems (no calculator). Be
sure to show your solution process.

‡ Say whether ÷ is greater than or less than ÷ without solving. Explain your
reasoning.
‡ Johnny’s Pizza Express sells several different flavour large-size pizzas. One day,
it sold 24 pepperoni pizzas. The number of plain cheese pizzas sold on that day
was of the number of pepperoni pizzas sold, and of the number of deluxe pizzas
sold. How many deluxe pizzas did the pizza express sell on that day? (Note: in the
Chinese version, the pizza was changed to mooncake and the person’s name was
also changed to accommodate cultural differences, while all other numerical and
context information stayed the same.)
Item 7 (designed to assess knowledge component (3) of mathematics
conceptual knowledge for teaching). How would you explain to your students
why ÷ 2 = ? Why ÷ = 4? (item adapted from Tirosh, 2000).

ELEMENTARY SCHOOL TEACHERS’ MATHEMATICS CONCEPTUAL


KNOWLEDGE FOR TEACHING: SAMPLE RESPONSES
FROM MAINLAND CHINA

In Mainland China, the survey was given to prospective elementary school teachers
in five institutions located in three provinces. All five institutions are province-level
normal universities or colleges that offer 4-year B.A. or B.Sc. preparation programs,
and the three provinces are diverse in terms of their locations and economic
development. However, the selection of these institutions and provinces resulted
mainly from convenience sampling, with access to prospective elementary school
teachers readily available for conducting the survey. All surveys were conducted
in classrooms to be completed within one hour with instructors’ supervision. 350

91

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

surveys were distributed, and 319 responses (response rate: 91.1%) were collected.
All 319 responses (299 females, 18 males) are used for data reporting, with 243
(76%) responses from prospective teachers in their third year in the program, 40
(13%) responses from prospective teachers in their fourth year in the program, and
36 (11%) responses from master programs.
Due to the sampling difficulties associated with prospective elementary school teachers
in Mainland China, the results reported here should not be taken as reflecting the overall
situation in Mainland China or assuming well-controlled samplings for comparison.
Nevertheless, the results should offer a glimpse of sampled prospective elementary
school teachers’ mathematics conceptual knowledge for teaching fraction division.

Sampled Teachers’ Mathematics Conceptual Knowledge for Teaching in


Elementary School Mathematics: Fraction Division

In general, the results from all sampled teachers’ responses present a consistent
pattern. At the item level, sampled prospective teachers did very well in answering
Item 6 that assesses elementary school teachers’ knowledge and skill associated
directly with fraction division, and were quite successful on an item (i.e., Item 7)
assessing elementary school teachers’ knowledge and ability of teaching fraction
division using various representations or models. But those prospective teachers
were much less successful in answering Item 5 that examines elementary school
teachers’ knowledge of fraction division and their ability to connect and justify
possible association between fraction division and other content topics.
In the following sections, we present the results item-by-item from sampled
prospective (319 respondents) teachers in Mainland China. (Note: the results
presented in the following text may not add to 100% due to rounding errors.)

Item 6. Item 6 was designed to assess elementary school teachers’ knowledge and
skills associated directly with fraction division (i.e., knowledge component (1)).

Elementary school teachers’ responses to the first sub-question of Item 6. For the
group of sampled prospective elementary school teachers, about 96.2% of those
respondents provided the correct answer (i.e., the first numerical expression is
greater than the second one). And the remaining 3.8% did not get the correct answer.
Among those who provided the correct answer, about 78.4% did not use fraction
division computations. The common explanations include (a) “If the dividend is the
same, the smaller the divisor, the larger the quotient.” and (b) “ ( ) is smaller than
( ).” And many respondents provided both reasons. The other 16.9% used the
computation rule for fraction division (i.e., converting division into multiplication,
then followed by comparing and ) to reach the correct answer. A very small
percent of sampled prospective teachers (3 respondents, 0.9%) used both methods.
The remaining 3.8% did not get the correct answer due to either misconceptions

92

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

(e.g., if the dividend is the same, the larger the divisor, the larger the quotient; or
a mistake about which fraction is bigger between and ), or computation errors.
Overall, it is clear that sampled prospective elementary school teachers did very
well on this sub-item (96.2% correct).

Elementary school teachers’ responses to the second sub-question of Item 6. For


the group of prospective elementary school teachers, 94% of respondents answered
correctly. Specifically, 64.3% used a multi-step computation method to get the answer,
about 27% used a combined computation method, 1.6% adopted an algebraic approach
to set up and solve an equation for the solution, and a few (about 1.1%) provided more
than one solution approach. About 6% of these respondents responded incorrectly.
Overall, sampled prospective elementary school teachers did very well (94%
correct). For those who responded incorrectly, data show their errors, resulting
either from computation error (e.g., proving a computation as 24 × ÷ = 36) or
misunderstanding the problem (e.g., proving a computation as 24 ÷ = 36).

Item 5. Item 5 was designed to assess elementary school teachers’ knowledge


of fraction division and their ability to connect and justify possible associations
between fraction division and other content topics (i.e., knowledge component (2)).
This item contains two sub-questions and sampled prospective teachers were asked
to answer both questions.

Elementary school teachers’ responses to the first sub-question of Item 5. For the
first sub-question about the meaning of fraction division, about 73% of the sampled
prospective teachers provided correct explanations. Among those correct answers,
9.4% prospective teachers responded with “the meaning of fraction division is
the same as the division of whole numbers,” 28.2% provided their answers as “if
knowing the product of two factors and the value of one factor, it is an operation to
find the value of the other factor,” 17.6% responded with an answer that combines
the above two, as “the meaning of fraction division is the same as the meaning of
the division of whole numbers, and if knowing the product of two factors and the
value of one factor, it is an operation to find the value of the other factor,” 11.9%
explained the meaning of fraction division as “partitioning a number into several
parts, taking one part or several parts.” The rest of the answers include comparing
fraction division with ratio, or operations involving whole numbers or fractions. Few
of the sampled prospective teachers (1.5%) used two different ways to explain the
meaning of fraction division. Among the sampled prospective teachers, 27% either
provided a wrong explanation (13.5%) or no explanation at all (13.5%).

Elementary school teachers’ responses to the second sub-question of Item 5. For the
second sub-question, sampled teachers were asked to explain how fraction division
relates to other content topics. Only 18.3% of prospective teachers provided correct

93

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

explanations, with 13.2% indicating that “fraction division is an inverse operation of


fraction multiplication” and 3.8% mentioning that “fraction division relates to inverse
numbers, for example, dividing by a number equals to multiplying by its inverse
number.” The vast majority (81.8%) of prospective teachers either stated what content
topics may relate to fraction division but failed to explain how (37.6%), provided some
other wrong explanation (7.8%), or simply did not answer this sub-question (36.4%).

Item 7. Item 7 was designed to assess elementary school teachers’ knowledge


and ability in teaching fraction division using various representations or models
(i.e., knowledge component (3)). It also contains two sub-questions and sampled
prospective teachers were asked to answer both questions.

Elementary school teachers’ responses to the first sub-question of Item 7. For the first
fraction division (i.e., explaining why ÷ 2 = ?), 95.2% of sampled prospective teachers
provided valid explanations for dividing a fraction by a natural number (i.e., ÷ 2 = ).
The dominant explanation (49.5%) used the meaning of fractions such as, “dividing a
whole into three equal parts, each part should be , so means having two such parts.
Dividing into two equal pieces, so each piece should be .” The other 17.9% were
dominated by explanations that were based on the algorithm, “dividing a number equals
to multiplying its reciprocal,” 14.1% explained with a drawing or number line, and
about 4.9% provided correct explanations with two or more different approaches.

Elementary school teachers’ responses to the second sub-question of Item 7. For


the second fraction division (i.e., explaining why ÷ = 4), 84.3% of sampled
prospective teachers provided valid explanations but the dominant explanation was
based on the fraction division algorithm (37.9%). Among the prospective teachers
27.3% provided their explanations mainly as “changing fractions so that they have
the same denominator first, and then using the meaning of fractions for the solution,
for example, changing into its equivalent fraction , and has four .” Some (8.5%)
used drawings or a number line to help explain. About 2.1% provided different
explanations. The rest (7.2%) provided incorrect explanations. The dominant error
was due to the inability to explain why ÷ = 4, although those prospective teachers
tried to provide a problem context for the fraction division.

Summary of elementary school teachers’ responses to Item 7. Overall, sampled


prospective teachers did very well in explaining these two fraction divisions. They
used a broad range of approaches in providing their explanation, with an approach
that relies on either the meaning of fraction or the fraction division algorithm.
Moreover, sampled prospective teachers performed better in explaining a fraction
divided by a whole number than explaining a fraction divided by a fraction. The
results suggest that when both the dividend and divisor are fractions, it likely makes

94

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

the explanation more difficult when the divisor is a fraction than the case when the
divisor is a whole number.

ELEMENTARY SCHOOL TEACHERS’ MATHEMATICS CONCEPTUAL


KNOWLEDGE FOR TEACHING: SAMPLE RESPONSES FROM SOUTH KOREA

In South Korea, the survey was given to prospective elementary school teachers
in four national universities that offer 4-year B.A. or B.Sc. teacher preparation
programs. There are 13 special universities offering elementary school teacher
education programs in South Korea, which are geographically spread out across
the country. Ten of them are national universities that are designed only to prepare
elementary school teachers, and the remaining three universities provide various
programs along with an elementary school teacher education program. In this
study, three were selected from the 10 universities specializing in preparing only
elementary school teachers and the remaining one is a comprehensive university
specializing in teacher education ranging from preschool through elementary to
secondary school education. The selection of these universities represents quite well
of those leading institutions in teacher preparation in South Korea. All surveys were
conducted in classrooms to complete within one hour with instructors’ supervision.
238 surveys were distributed and collected. However, 17 surveys were not used
because there were no answers to three or more items. These unreliable responses
happened mostly in one university in which the survey was administrated in the last
class session along with the final exam for the course. Thus, 221 responses (167
females, 54 males) were used for data analyzing and reporting, with 135 (61%) of
responses from prospective teachers in their third year in the program, 86 (39%)
responses from prospective in their fourth year in the program.
Due to the sampling difficulties associated with prospective elementary school
teachers in South Korea, the results reported here should not be taken as simply
reflecting the overall situation in the country. Nevertheless, with the consideration of
selecting samples across the country, the results should allow us to get a good sense
of prospective elementary school teachers’ mathematics conceptual knowledge for
teaching in fraction division in South Korea.

Sampled Teachers’ Mathematics Conceptual Knowledge for Teaching in


Elementary School Mathematics: Fraction Division

Overall, similar to the case of Mainland China, the results from all sampled teachers’
responses present a quite consistent pattern. At the item level, sampled prospective
teachers did very well in answering Item 6 that assesses elementary school teachers’
knowledge and skill associated directly with fraction division, and had excellent
performance on Item 7 that examines elementary school teachers’ knowledge and
ability of teaching fraction division using various representations or models. But

95

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

these prospective teachers were much less successful in answering Item 5 that tends
to assess elementary school teachers’ knowledge of fraction division and their ability
to connect and justify possible associations between fraction division and other
content topics.
In the following sections, we present the results item-by-item from sampled
prospective (221 respondents) teachers in South Korea. (Note: the results presented
in the following items may not add to 100% due to rounding errors.)

Item 6

Elementary school teachers’ responses to the first sub-question of Item 6. About


96% of sampled prospective teachers answered correctly (i.e., the first numerical
expression is greater than the second one). The most common explanation is that
is smaller than . Some teachers showed why is smaller than by comparing these
fractions with 1 (i.e., 1 – vs 1 – ), converting them to equivalent fractions with

the same denominator (i.e., and ), or drawing a picture to represent and for
comparison, etc. About 26% mentioned, “If the divisor is the smaller, the result of
the division (or quotient) is bigger.” About 5% who got the correct answer changed
the division of the given numerical expressions into multiplication and mentioned
that is greater than , implying “the greater the multiplier, the larger the product.”

Elementary school teachers’ responses to the second sub-question of Item 6. For


the group of prospective elementary teachers, 93% of respondents responded
correctly. Specifically, 70% used a multi-step computation method to get the answer
(i.e., 24 ×   ͒×   ͒ ×  ͒  DERXWXVHGDFRPELQHG
computation method (i.e., 24 ×  ͒×  ͒ × ÷ = 24 × × = 27), 5%
adopted an algebraic approach to set up and solve an equation for solution (e.g.,
Cheese pizzas, x = 18, Deluxe pizzas = y, y =18, 2y = 54, y = 27), 6% provided
a correct answer with no explanation, 5% provided a simple explanation without
variables (e.g., of 24 is 18, of Deluxe is 18, so (the answer is) 27), and about
1% (2 respondents) provided something else. The remaining prospective elementary
school teachers either did incorrectly (15 respondents, 7%) or provided no answer at
all (one respondent). Incorrect responses are mainly due to computation (e.g., errors
in reducing fractions) or misunderstanding the problem.
Overall, sampled prospective elementary school teachers did very well (93%
correct). Furthermore, those prospective teachers tended to use either a multi-step
or combined arithmetic approach. For those respondents who answered incorrectly,
data show their errors, resulting either from computation errors or misunderstanding
of the problem.

96

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

Item 5

Elementary school teachers’ responses to the first sub-question of Item 5. As


explained above, this item includes two sub-questions and sampled prospective
and practicing teachers responded to both questions. For the first sub-question
about the meaning of fraction division, about 41% sampled prospective teachers
provided correct explanations, and as much as 46% (i.e., 101 prospective teachers)
did not get the correct answer. If we look closely at the correct answers, the most
common explanation was the measurement interpretation of fraction division
(39%), followed by the partitive interpretation (29%). In addition, more than 10%
of sampled prospective teachers were able to provide other meanings of fraction
division such as the inverse of multiplication (10%) or determination of a unit
rate (15%). Note that 28% of the prospective teachers were able to explain the
meanings of fraction division in two ways or more. Among the incorrect answers
(46%), the most common explanation (32%) was to describe the meaning of
fraction division as division with fractions (i.e., division with the divisor and/
or the dividend as fractions). About 13% provided no answer or simply stated “I
don’t know.”

Elementary school teachers’ responses to the second sub-question of Item 5. For


the second sub-question, as much as 69% were able to relate fraction division to other
content topics, although only about 41% of sampled prospective teachers provided
correct explanations of the meaning of fraction division (i.e., the first sub-question).
The most common content topic related to was fraction multiplication mainly
because the multiplicative inverse of the divisor is used in fraction division. Note
that both measurement interpretation and partitive interpretations used in answering
the first sub-question are related to the meaning of division and more than 29% of
the prospective teachers were able to provide these interpretations. In contrast, only
15% of the prospective teachers related fraction division to whole number division
and 13% related it to the division of decimal numbers. Note also that ratio, rate,
and proportion were mentioned mainly because of the meaning of determination of
a unit rate in fraction division. Another noticeable aspect is that 21% were able to
provide two or more related content topics. About 8% of prospective teachers failed
to provide a correct explanation, and 22% provided no answer or simply stated “I
don’t know.”

Item 7

Elementary school teachers’ responses to the first sub-question of Item 7. For the first
fraction division, almost all prospective teachers (99%) provided valid explanations
for dividing a fraction by a whole number. The majority of respondents (more than

97

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

54% respondents) used drawings to show that if you equally divide into 2 pieces,
you get . Other respondents (11%) used the meaning of division or fraction without
drawing. Some respondents (5%) used the common denominator and others (5%) used
an algorithmic approach (i.e., dividing a number equals to multiplying its reciprocal).
About 11% of these prospective teachers provided valid explanations in two or more
ways. In doing so, drawing was often used as a basic approach.

Elementary school teachers’ responses to the second sub-question of Item 7. With


regard to the second fraction division (i.e., dividing a fraction by a fraction), about
97% of sampled prospective teachers provided valid explanations. Again, the most
dominant explanation was based on drawings to show the meaning of measurement
division. Even though the drawings were different from explaining the first fraction
division, the main idea was to display how many s are included in (or ). Additionally,
19% of these prospective teachers explained the meaning of the measurement division
in words or numerical expressions without drawing. 15% of these respondents used an
algorithmic approach of using the inverse number. Note that the rate for the category
of finding the common denominator was increased from 5% (for explaining the first
fraction division) to 14% (for explaining the second fraction division), and that the
rate for the category of using an algorithmic approach was also increased from 5% to
15%. Among the prospective teachers 10% were able to provide two or more kinds of
explanations. Again, drawing was the most prevalent approach included.

Summary of elementary school teachers’ responses to Item 7. Putting together


sampled prospective teachers’ performance in explaining these two sub-questions
about fraction divisions, it is clear that these prospective teachers had almost
perfect performance in answering this item. The results show that these teachers
not only did very well in explaining these fraction divisions, but also used a range
of approaches in their explanations. The results likely suggest that these sampled
teachers, although diverse in terms of when and where they received their education,
shared quite similar understanding in how to explain these fraction divisions (note:
with the dominant approach of drawings).

SAMPLED ELEMENTARY SCHOOL TEACHERS’ MATHEMATICS


CONCEPTUAL KNOWLEDGE FOR TEACHING IN MAINLAND CHINA AND
SOUTH KOREA: WHAT CAN WE LEARN?

The results obtained from prospective teachers sampled in Mainland China and
South Korea provide rich information not only about their mathematics conceptual
knowledge for teaching, but also about possible patterns that can allow us to gain
initial understanding of teacher knowledge differentiated into different components.
Before we summarize the results below, we should emphasize again that the sampling
differences within and across these two countries do not allow us to generalize the results

98

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

for representing the entire education systems or to make a cross-system comparison,


such as using a statistical analysis. Instead, we put the main results reported in the
above two parts together to illustrate what we can learn from these sampled teachers’
responses. We believe such illustrations can help inspire further investigations about
teachers’ mathematics conceptual knowledge for teaching, and the role of culture and
education in shaping teachers’ mathematics conceptual knowledge for teaching.
Table 3.1 summarizes the percentages of sampled prospective teachers in
Mainland China and South Korea who provided correct answers or explanations to
items 5, 6, and 7. In general, sampled teachers had better performance on items 6
and 7 than their performance on item 5 in both Mainland China and South Korea.
For item 6, sampled teachers had similar great performance across these two
education systems. The results suggest that prospective teachers sampled in both
education systems had solid knowledge and skill directly associated with fraction
division, a knowledge component that is also typically required for school students.
It is noticeable that sampled prospective teachers in South Korea did even better on
item 7 that aims to assess their knowledge and ability of teaching fraction division.
The results likely suggest that teacher preparation programs in South Korea placed
great emphases on mathematics and mathematical pedagogy education. For the case
of Mainland China, the results present a pattern different from the case of South
Korea. Sampled prospective teachers in Mainland China also did well on item
7 but not better than their performance on item 6. The results likely suggest that
prospective teachers in Mainland China also received good education on mathematics
and mathematics pedagogy through teacher preparation program, but they would be
expected to gain even more on how to teach fraction division through teaching. The
similarities and differences in prospective teachers’ performance on items 6 and 7

Table 3.1. Summary results of sampled elementary school teachers’


mathematics conceptual knowledge for teaching in fraction division

Mainland China South Korea


Correct Correct

Item 6
Sub-question 1 96.2% 96%
Sub-question 2 94% 93%
Item 5
Sub-question 1 73% 91%
Sub-question 2 18.3% 69%
Item 7
Sub-question 1 95.3% 99%
Sub-question 2 84.3% 97%

99

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

between Mainland China and South Korea are consistent with what we learned last
time (see Li et al., 2008).
Noticeable results were revealed from sampled prospective teachers’ performance
in answering item 5 in Mainland China and South Korea. While sampled teachers
in both education systems did better on the first sub-question than the second sub-
question, it is clear that there are overall dramatic differences across these two
education systems. First of all, the group of prospective teachers sampled in both
Mainland China and South Korea did much less successfully in answering sub-
question 2 than sub-question 1. Second, sampled prospective teachers in South Korea
did better in both sub-questions than sampled prospective teachers’ performance
in Mainland China. The results likely suggest that sampled prospective teachers in
South Korea received good education through their teacher preparation program on
such knowledge component in mathematics conceptual knowledge for teaching,
whereas it may be less so for the sampled prospective teachers in Mainland China.
To better understand these prospective elementary school teachers’ performance
it is also important to know the system and cultural contexts of Mainland China
and South Korea. Readers can find detailed information from a previous study
(Li et al., 2008). In South Korea, elementary school teachers need to teach every
school subject, prospective elementary school teachers are required to take the same
courses with only the exception of their focus areas. In general, about 85% of the
course requirements are the same for any prospective elementary school teacher.
With regard to requirements common in mathematics, elementary school teacher
education programs consist of the liberal arts courses and the major courses. Among
the liberal art courses, prospective teachers take a compulsory course (2-3 credit
hours) dealing with the foundations or basics of mathematics. Among the major
courses, prospective teachers take two to four compulsory courses (4–7 credit hours)
in mathematics. The most common courses are (Elementary) Mathematics Education
I with 2 credit hours and (Elementary) Mathematics Education II with 3 credit hours.
The former mainly deals with overall theories (including the national mathematics
curriculum) related directly to teaching elementary school mathematics, whereas
the latter covers how to teach elementary school mathematics tailored to multiple
content areas such as number and operations.
The situation in Mainland China shows great variations across different teacher
preparation programs that are still in co-existence in the system (Xie, Ma, & Chen,
2018). The comprehensive-type teacher preparation program offers only two or three
mathematics courses such as, Advanced Mathematics, Foundations of Mathematics,
and Elementary Number Theory. In contrast, the discipline-based or stream model
program offers many more compulsory and elective courses in mathematics to
help prospective teachers to build a strong subject matter knowledge. Such courses
can be classified into three categories: basic courses (e.g., calculus), professional
courses (e.g., elementary number theory), and advanced courses (e.g., mathematical
modelling). In addition, there are some courses that are closely related to curriculum
and instruction in elementary school mathematics.

100

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

The mathematics course requirements have more consistence across different


programs in South Korea than the case in Mainland China. The credit hour
requirements are in line with what is proposed in the United States (CBMS,
2012). Although there are dramatic variations in the credit hour requirements in
mathematics in Mainland China, the minimum requirements in mathematics in the
comprehensive model programs is also aligned with CBMS’s recommendations
(2012). Sampled prospective elementary school teachers’ performance in South
Korea and Mainland China in this study suggest that these sampled prospective
teachers benefited from strong mathematics preparation provided through their
programs. Even more important, although programs in South Korea do not seem to
require more courses in mathematics than programs in Mainland China, the good
education provided in teacher preparation programs in South Korea is likely due to
the nature of courses provided. Specifically, their most common courses such as,
(Elementary) Mathematics Education I and (Elementary) Mathematics Education
II, likely help provide important preparation in mathematics as measured in this
study. Further research would be needed to understand the nature of mathematics
courses offered in different system contexts, and how the content in those courses
may provide mathematical preparation that is needed and important for teaching.

CONCLUSION

The results and discussion presented previously show some important findings. It is
quite clear that, across these two education systems, sampled prospective teachers’
performance shared much similarities in terms of their mathematics conceptual
knowledge for teaching on fraction division. However, there are subtle differences
in details across these two education systems. For example, prospective teachers
sampled in Mainland China and South Korea showed quite different tendency in
approaches when answering the three items targeted on teachers’ mathematics
conceptual knowledge for teaching, especially for item 7 on mathematical pedagogy
and item 5 on connections of mathematical ideas. Sampled teachers in Mainland
China tended to use numerical, algorithmic or verbal explanation when answering
these questions, whereas sampled teachers in South Korea showed more use of visual
representations such as drawings for explanation. Such differences are likely beyond
sampling differences and can possibly relate to the different education provided for
prospective elementary school teachers and cultural practice embedded in teaching
in these two education systems.
At the beginning of the chapter, we mentioned that it remains to be understood
whether a strong education in mathematics is good enough for what we believe that
teachers need to know more than students for teaching. Specifically, we chose the
topic of fraction division. Fraction division is a difficult topic in elementary school
mathematics (e.g., Ma, 1999), not only for school students but also for teachers
(e.g., Li, 2008; Li & Smith, 2007). By focusing on this content topic, we developed
a survey with test items that aims to assess prospective teachers’ three different

101

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

knowledge components of Mathematics Conceptual Knowledge for Teaching on this


topic. Sampled prospective teachers’ performance from Mainland China and South
Korea helped reveal that teachers’ knowledge on fraction division can and should be
differentiated, and the structure of mathematics conceptual knowledge for teaching
with three knowledge components provides an important and feasible lens for us to
know the strength and weakness of teachers’ knowledge. In particular, for the case
of Mainland China, the results suggest that prospective teachers likely gain much
more on mathematics, somehow less on mathematical pedagogy, but very limited
on connections of mathematical ideas through their program studies. For the case
of South Korea, great performance demonstrated by sampled prospective teachers
suggests that they received very good education on mathematics and mathematical
pedagogy, but less so on connections of mathematical ideas through their program
studies. It is the education they received through teacher preparation programs that
likely lays a great foundation for their professional career in the future. It is our
hope that the conception and structure of Mathematics Conceptual Knowledge
for Teaching (MCKT) can and shall further our efforts in helping to guide our
prospective and practicing teachers to know mathematics well enough, beyond what
students need to know, for teaching.

ACKNOWLEDGEMENTS

We would like to thank many graduate students and colleagues in Mainland China
and South Korea for their assistances in the process of collecting and analyzing the
data. We are also grateful to all the survey participants in Mainland China and South
Korea for sharing their thoughts and contributions.

REFERENCES
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Blömeke, S., Hsieh, F.-J., Kaiser, G., & Schmidt, W. H. (Eds.). (2014). International perspectives on
teacher knowledge, beliefs and opportunities to learn. Dordrecht: Springer.
Borko, H., Eisenhart, M., Brown, C. A., Underhill, R. G., Jones, D., & Agard, P. C. (1992). Learning
to teach hard mathematics: Do novice teachers and their instructors give up too easily? Journal for
Research in Mathematics Education, 23(3), 194–222.
Conference Board of the Mathematical Sciences (CBMS). (2012). The mathematical education of teachers
II, CBMS issues in mathematics education (Vol. 17). Washington, DC: American Mathematical
Society and Mathematical Association of America.
Döhrmann, M., Kaiser, G., & Blömeke, S. (2018). The conception of mathematics knowledge for
teaching from an international perspective: The case of the TEDS-M study. In Y. Li & R. Huang
(Eds.), How Chinese acquire and improve mathematics knowledge for teaching (pp. 57–81). Leiden,
The Netherlands: Brill-Sense.
Elashhab, G. A. (1978). Division of fractions – Discovery and verification. School Science and
Mathematics, 78, 159–162.
Greer, B. (1992). Multiplication and division as models of situations. In D. A. Grouws (Ed.), Handbook of
research on mathematics teaching and learning (pp. 276–295). New York, NY: Macmillan.

102

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS CONCEPTUAL KNOWLEDGE FOR TEACHING

Gregg, J., & Gregg, D. U. (2007). Measurement and fair-sharing models for dividing fractions.
Mathematics Teaching in the Middle School, 12, 490–496.
Hill, H. C., Schilling, S. G., & Ball, D. L. (2004). Developing measures of teachers’ mathematics
knowledge for teaching. The Elementary School Journal, 105(1), 11–30.
Li, S., Huang, R., & Shin, Y. (2008). Mathematical discipline knowledge requirements for prospective
secondary teachers from an East Asian perspective. In P. Sullivan & T. Wood (Eds.), International
handbook of mathematics teacher education: Knowledge and beliefs in mathematics teaching and
teaching development (pp. 63–86). Rotterdam, The Netherlands: Sense Publishers.
Li, Y. (2002). Knowing, understanding and exploring the content and formation of curriculum materials:
A Chinese approach to empower prospective elementary school teachers pedagogically. International
Journal of Educational Research, 37(2), 179–193.
Li, Y. (2008). What do students need to learn about division of fractions? Mathematics Teaching in the
Middle School, 13, 546–552.
Li, Y., & Huang, R. (2008). Chinese elementary mathematics teachers’ knowledge in mathematics and
pedagogy for teaching: The case of fraction division. ZDM-The International Journal on Mathematics
Education, 40(5), 845–859.
Li, Y., & Kaiser, G. (Eds.). (2011). Expertise in mathematics instruction. New York, NY: Springer.
Li, Y., & Kulm, G. (2008). Knowledge and confidence of pre-service mathematics teacher: The case of
fraction division. ZDM-The International Journal on Mathematics Education, 40(5), 833–843.
Li, Y., Ma, Y., & Pang, J. (2008). Mathematical preparation of prospective elementary teachers. In
P. Sullivan & T. Wood (Eds.), International handbook of mathematics teacher education: Knowledge
and beliefs in mathematics teaching and teaching development (pp. 37–62). Rotterdam, The
Netherlands: Sense Publishers.
Li, Y., & Smith, D. (2007). Prospective middle school teachers’ knowledge in mathematics and pedagogy
for teaching – The case of fraction division. In J. H. Woo, H. C. Lew, K. S. Park, & D. Y. Seo (Eds.),
Proceedings of the 31st Conference of the International Group for the Psychology of mathematics
education (Vol. 3, pp. 185–192). Seoul, The Republic of Korea: PME.
Ma, L. (1999). Knowing and teaching elementary mathematics: Teachers’ understanding of fundamental
mathematics in China and the United States. Mahwah, NJ: Lawrence Erlbaum Associates.
National Council of Teachers of Mathematics. (2000). Principles and standards of school mathematics.
Reston, VA: National Council of Teachers of Mathematics.
Petrou, M., & Goulding, M. (2011). Conceptualizing teachers’ mathematical knowledge in teaching. In
T. Rowland & K. Ruthven (Eds.), Mathematical knowledge in teaching (pp. 9–25). New York, NY:
Springer.
Rowland, T. (2013). The Knowledge Quartet: The genesis and application of a framework for analyzing
mathematics teaching and deepening teachers’ mathematics knowledge. SISYPHUS Journal of
Education, 1(3), 15–43.
Ruthven, K. (2011). Conceptualizing mathematical knowledge in teaching. In T. Rowland & K. Ruthven
(Eds.), Mathematical knowledge in teaching (pp. 83–96). New York, NY: Springer.
Schmidt, W., Tatto, M. T., Bankov, K., Blömeke, S., Cedillo, T., Cogan, L., … Schwille, J. (2007). The
preparation gap: Teacher education for middle school mathematics in six countries. East Lansing,
MI: Michigan State University.
Shulman, L. S. (1986). Those who understand: Knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Simon, M. A. (1993). Prospective elementary teachers’ knowledge of division. Journal for Research in
Mathematics Education, 24(3), 233–254.
Tirosh, D. (2000). Enhancing prospective teachers’ knowledge of children’s conceptions: The case of
division of fractions. Journal for Research in Mathematics Education, 31(1), 5–25.
Xie, S., Ma, Y., & Chen, W. (2018). Elementary mathematics teacher preparation in China. In Y. Li &
R. Huang (Eds.), How Chinese acquire and improve mathematics knowledge for teaching (pp. 85–108).
Leiden, The Netherlands: Brill Sense.

103

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
YEPING LI ET AL.

Yeping Li
College of Education and Human Development
Texas A&M University

JeongSuk Pang
Department of Elementary Education
Korea National University of Education

Huirong Zhang
Faculty of Education,
Southwest University

Naiqing Song
School of Mathematics and Statistics
Southwest University

104

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

4. RESEARCHING MATHEMATICAL KNOWLEDGE


IN TEACHING

A programme of research which began at the University of Cambridge in 2002 has


investigated the mathematics content knowledge of teachers, and the ways that
this knowledge ‘plays out’, both in teachers’ planning and in their teaching in the
classroom. From this research, a framework – the Knowledge Quartet – for the
observation, analysis and development of mathematics teaching was developed,
with a focus on the contribution of the teacher’s mathematical content knowledge.
The Knowledge Quartet identifies situations in which content-related knowledge
can be seen in the act of teaching. The origins of the Knowledge Quartet were in
observations of primary mathematics teaching, and grounded theory methodology,
in the context of a one-year graduate primary teacher preparation. Since 2002,
there has been a continuous process of refinement of the conceptualisation of the
Knowledge Quartet, and enhancement of the constituent codes, both in response to
additional classroom data and in the process of application. This chapter describes
the process by which the Knowledge Quartet evolved, the chronology of that process,
and some of the associated theoretical influences and pragmatic constraints.

INTRODUCTION

All professions (e.g., lawyers, architects and bakers) have knowledge bases that
are specific to their realm of operation, and teaching is no exception. In the 1980s,
Lee Shulman, a former president of the American Education Research Association,
gave a seminal account of the knowledge bases needed by teachers (Shulman,
1986). Three of these elements are specific to the subject matter being taught.
They are: subject matter knowledge, pedagogical content knowledge and curricular
knowledge. Shulman (1986) notes that the ways of conceptualising subject matter
knowledge will be different for different subject matter (discipline) areas, but in
his generic account he includes Schwab’s (1978) notions of substantive knowledge
(the key facts, concepts, principles and explanatory frameworks in a discipline) and
syntactic knowledge (the nature of enquiry in the field, and how new knowledge
is introduced and accepted in that community). For Shulman, pedagogical content
knowledge (PCK) consists of “the ways of representing the subject which make it
comprehensible to others […] [it] also includes an understanding of what makes
the learning of specific topics easy or difficult […]” (Shulman, 1986, p. 9). The

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_005

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

notion of pedagogical content knowledge as a distinct domain has been disputed


(e.g., McNamara, 1991), but Shulman’s work has fed a cascade of research into
mathematics teacher knowledge from then until the present.
This chapter begins in the context of United Kingdom research on prospective
teachers’ mathematics disciplinary knowledge. It charts research undertaken since
1998 by teams in London and Cambridge. The majority of the chapter concerns
a research project which investigated how mathematics teachers’ disciplinary
knowledge is made visible in teaching itself, and follow-up to that project.

UNITED KINGDOM RESEARCH ON MATHEMATICS TEACHER KNOWLEDGE

The process of “audit and remediation” of mathematics subject knowledge within


primary Initial Teacher Education became a high-profile issue in England following
the introduction of various United Kingdom government requirements at the close
of the last millennium (Department for Education and Employment, 1998). The
introduction of this regime provoked a body of research in the United Kingdom
on prospective elementary teachers’ mathematics subject-matter knowledge (e.g.,
Goulding, Rowland, & Barber, 2002; Goulding & Suggate, 2001; Morris, 2001;
Rowland, Martyn, Barber, & Heal, 2000; Sanders & Morris, 2000). In every case,
the participants were prospective (so-called ‘trainee’) elementary school teachers,
on three- or four-year undergraduate or one-year graduate teacher preparation
courses. The methodology at the heart of these studies involved a questionnaire –
a test – taken at some time before or during the course. The subject matter being
assessed was typically determined by that specified by the Department for Education
and Employment (1999), that is, mainly related directly to the elementary school
curriculum, but also including some topics in the school curriculum up to about
Year 9 (pupil age 14). The logic of the situation pointed to auditing the prospective
teachers’ mathematics knowledge at the beginning of their course, to establish where
the ‘gaps’ were at the outset, and this was the case in most of the studies.
An exception was that of Rowland, Martyn, Barber, and Heal (2000), in which
the audit was administered four months into a one-year course, after the prospective
teachers had encountered the mathematics content within the teaching methods course,
“giving them maximum opportunity and professional motivation to recall those topics
they had forgotten (for lack of use) since they did mathematics at school” (p. 4). The
research of this London-based team included an investigation of the relation between
the prospective teachers’ subject matter knowledge, as assessed by a 16-item written
audit instrument, and their teaching competence. The audit was administered by the
research team; the assessment of teaching, on a four-point scale based on criteria in
regular use, was made jointly by their university-based tutor and a practising teacher-
mentor in the participant’s placement school. A chi-square test showed a significant
association between audit score and an assessment of teaching competence. This
finding turned out to be robust when replicated with a different cohort of prospective
teachers (Rowland et al., 2001). Participants obtaining high (or even middle) scores

106

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

on the audit were more likely to be assessed as strong mathematics teachers than
those with low scores; whereas those with low audit scores were more likely than
other participants to be assessed as weak mathematics teachers.

INTO THE MATHEMATICS CLASSROOM

Although this was interesting in itself, and attracted some media attention, we wanted
to find out more about what was ‘going on’. By now, several United Kingdom-
based researchers had formed a consortium named SKIMA (‘subject knowledge in
mathematics’) and a five-person Cambridge-based SKIMA subgroup took forward
this new line of enquiry.
The reasoning behind this move was, if superior subject matter knowledge really
does make a difference when teaching elementary mathematics, it ought somehow to
be observable in the practice of the knowledgeable teacher. Conversely, the teacher
with more limited subject matter knowledge might be expected to misinform their
pupils, or somehow to miss opportunities to teach mathematics ‘well’. In a nutshell,
we (the Cambridge team) wanted to identify, and to understand better, the ways in
which elementary teachers’ mathematics content1 knowledge, or the lack of it, is
evident in their teaching. Certain parallels can be drawn with the work of Deborah
Ball and her colleagues that provided a “practice-based theory of knowledge for
teaching” (Ball & Bass, 2003), but the two theories are very different. In particular,
Ball’s Mathematical Knowledge for Teaching (MKT) theory unravels and clarifies
the formerly somewhat elusive and theoretically-undeveloped notions of subject
matter knowledge and pedagogical content knowledge. Shulman’s subject matter
knowledge is separated into ‘common content knowledge’ and ‘specialized content
knowledge’, while his pedagogical content knowledge is divided into ‘knowledge
of content and students’ and ‘knowledge of content and teaching’ (Ball, Thames, &
Phelps, 2008). In our theory – Mathematical Knowledge in Teaching – the distinction
between different kinds of mathematical knowledge is of lesser significance than
the classification of the situations in which mathematical knowledge for teaching
surfaces in the classroom. In this sense, the two theories each have useful perspectives
to offer to the other.
From the outset, we envisaged this research as a classroom observation study. We
were genuinely curious to know what knowledgeable teachers do in the classroom
that might enhance their pupils’ experience of learning mathematics, and what others
did not, or could not, do.

Participants and Ethical Desiderata

In the United Kingdom, most university-based prospective teachers follow a one-


year, full-time course leading to a Postgraduate Certificate in Education (PGCE),
some two-thirds of the year being spent working in a school under the guidance of a
school-based mentor. All prospective primary school teachers prepare to be generalist

107

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

teachers of the whole primary school curriculum. Most of the data collection for
this study took place in the context of the one-year Postgraduate Certificate in
Education course in which each of the 149 prospective teachers followed a route
focusing either on the ‘lower primary (LP)’ years (ages 3–8) or the ‘upper primary
(UP)’ (age 7-11). In the first instance, we obtained mathematics content knowledge
data for the whole cohort of prospective teachers, using the 16-item audit discussed
above, administered four months into the course. The next phase, the heart of our
study, entailed observations of lessons taught by some of the 149 participants. Since
we were interested in the relationship between mathematics content knowledge and
classroom teaching, it seemed to us that the participants observed ought to represent
a range of subject knowledge competence, as measured by the audit. For the purpose
of this research, the total scores for each paper (maximum 64) were used to identify
groups with ‘high’, ‘medium’ and ‘low’ scores. In order to compare and contrast
the teaching of the prospective teachers with different levels of audited content
knowledge, we decided to observe equal numbers of participants from each of the
three categories. Finally, because the curriculum would appear to make different
demands on content knowledge in the lower primary and upper primary school
years, we also wanted these two phases to be equally represented. Our resources
made it possible to devote the equivalent of one person full-time for two weeks
to the observation; each classroom visit would take about half a day to observe
and videotape2 one lesson. These factors and constraints eventually influenced the
decision to identify 12 prospective teachers for observation, and to observe each
of them teaching a whole mathematics lesson on two separate occasions. The 12
prospective teachers represented the intersection of each subject knowledge category
with each of the two LP/UP age-phase groups, with two participants in each cell of
the 3 x 2 grid of possibilities. Three were male, reflecting reasonably well the 1 in
6 proportion in this Postgraduate Certificate in Education cohort as a whole. These
12 were invited to participate in the video study phase of our project, and all agreed.
It was made clear to them, in writing, that our observations would play no part in
the university’s summative assessment of their teaching competence, or any other
aspect of their certification. We sought and gained their permission to use data
from the videotapes of their lessons in research papers, and to use short extracts
from some of the tapes for research presentations and in teacher education. Later,
it was also necessary to obtain similar permissions from mentors and headteachers
in these prospective teachers’ placement schools, and from the parents or carers of
the children in their classes. In a few cases this parental permission was withheld,
and practical arrangements were made to respect this choice whilst maintaining the
children’s participation.

Procedures

Data collection. The two lesson observations took place within an 8-week final
teaching placement. By then, the participants had completed the mathematics

108

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

methods course at the university and had taught in two schools for about 15 weeks
in total. The prospective teacher-participants were asked to provide a copy of their
planning for the observed lesson. The focus and content of the lesson were chosen
by the prospective teacher. The camera tracked the prospective teacher but would
occasionally capture some artefact in the classroom (such as a set of exercises
displayed on a wall) which had some immediate relevance to the mathematics
teaching being observed. When and if necessary, the observer made handwritten
field notes about any relevant aspects of the lesson that might not have been captured
on the video recording, such as off-camera pupil comments or interactions. Pupils’
spoken contributions were audible on the video-recording if they were picked up
by the radio microphone: for the most part, this included pupils’ remarks during the
whole-class teaching portions of the lesson and during seatwork portions when the
prospective teacher was working closely with an individual pupil or a group. Our
assessment of the obtrusiveness of the observer and the technology can be anecdotal
at best, but in the video recordings it is rare to see any evidence of the pupils taking
any interest in their ‘visitor’. At the conclusion of the lesson, the observer took care
not to give any feedback or evaluative comments on the lesson to the prospective
teacher in order to emphasise that they were not there in their role as teacher educator.

Data preparation. As soon as possible after the lesson (usually the same day) the
observer/researcher wrote what we call a Descriptive Synopsis of the lesson. This
was a brief (around 500 words) account of what happened in the lesson, so that
a reader might immediately be able to contextualise subsequent discussion of any
events within it. These descriptive synopses were typically written from memory
and with the use of the field notes, with occasional reference to the videotape if
necessary.
The Cambridge research team subscribed to the view that no ‘objective’ account
of a lesson can be written, and none can be read. However, we guarded against
‘smuggling’ interpretive and inferential passages into these descriptive synopses.
With this in mind, in addition to what the observer believed to be their best efforts at
straightforward description, different text styles were used to identify in the synopses
(a) anything that the observer/researcher thought might turn out to be significant, or
critical, moments or episodes with respect to the prospective teacher’s mathematics
content knowledge, for consideration later by the research team; and (b) any
evaluative comment within the descriptive synopsis; this was to allow occasional
(in fact, quite rare) comments of the kind that one might write, as a tutor, on a
lesson observation report (acclaim or criticism), yet which went beyond description.
A pilot lesson was videotaped and analysed in this way, as a kind of rehearsal of our
intended means of ‘capturing’ the lesson. We also discussed aspects of the content
of the lesson that caught our immediate interest with respect to our intended focus
on teacher knowledge.

109

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

Coding. After all the video tapes of the lessons were collected, the hard work of
analysing these 24 lessons began. We took a grounded approach to the data for the
purpose of generating theory (Glaser & Strauss, 1987). At the outset we did not
know what kind of ‘theory’ might emerge from our close scrutiny of the lesson
videotapes. It might have been an explanatory theory of the kind “Because this
teacher knew x, he or she did (or did not do) y in the lesson.” Alternately, it might
have been a ‘lens’ type of theory – a new way of seeing classroom events from the
perspective of teacher knowledge. In the event, the theory that materialised was
more of the second kind.
In our grounded theory approach to the videotapes of the lessons, we watched
the tapes of all the lessons, usually in two and threes, sometimes as a whole team.
We articulated and compared our interpretations of episodes from each of the 24
videotaped lessons. We identified specific actions in the classroom that seemed
to provide significant information about the prospective teachers’ mathematics
content knowledge or their mathematical pedagogical knowledge. We reflected
later that most of these actions related to various `choices made by the prospective
teachers, in their planning or more spontaneously. In this way, we homed in on
particular moments or episodes in the tapes. Each such moment or episode was
assigned a preliminary code (or more than one if appropriate). We developed these
codes as we went along (examples will be given later), and most of them recurred
as we saw what looked like the same kind of phenomenon in different episodes
within the same, or another, lesson. It soon became apparent that the majority of
the salient moments and episodes, and the corresponding codes, related to issues
of mathematics pedagogy3 rather than knowledge of mathematics per se. Perhaps
this is not surprising, since the subject-matter was elementary mathematics,
and half of the lessons were with children aged 4–7. For example, the issue in
introducing subtraction to young children is not whether an educated adult teacher
can himself, or herself, subtract one small integer from another, but whether they
know the fundamental subtraction structures or models, appropriate contexts for
these structures and ways of representing them, and a range of relevant student
mental strategies.4 Nevertheless, pupils’ spontaneous remarks and questions did,
on occasion, tax the prospective teachers’ overt knowledge and understanding of
mathematics, in unexpected ways (see Jason below).
At first the identification of such moments, and accounts of their significance for
our research, was in the form of proposals, or conjectures, for consideration by the
team. They could be challenged or supported and retained or rejected by consensus.
The grounds for such a challenge included relevance and significance. Relevance
is subjective to a degree, but the coding was expected to be relevant to the focus of
our research: the role of teachers’ mathematics-related knowledge in mathematics
teaching. For this reason, ‘child demonstration’ (CD), which was at one time one
of 18 agreed codes in use, was challenged and discarded, leaving 17 codes. Whilst
on several occasions the teachers did invite a child to demonstrate something to
the class, it was agreed that this could happen irrespective of the teacher’s subject

110

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

matter knowledge or pedagogical content knowledge. The notion of significance is


also subjective: while we did not attempt to define what made an event sufficiently
‘significant’ to merit being coded, decisions by individuals or subgroups in the
team were often challenged on the grounds of significance. For example, one of our
codes was called ‘overt subject knowledge’. It related to moments when the teacher
displayed some aspect of their subject matter knowledge. One such example was
when Nathalie compared the probabilities of two particular outcomes when two dice
are thrown, by identifying and listing the sample space and isolating the events. The
label itself (overt subject knowledge) could have been improved, but it stuck. On one
occasion, another prospective teacher had been teaching how to add 10 by adjusting
WKH WHQV GLJLW LQ WKH QXPHUDO HJ  ĺ   2QH RI WKH WHDP KDG SURYLVLRQDOO\
coded overt subject knowledge the prospective teacher’s fluency with the strategy.
This was successfully challenged, on the grounds that there was nothing special or
unusual about that. It is the kind of knowledge that the ‘average citizen’ would be
expected to have, whereas overt subject knowledge was intended to mark knowledge
that somehow went beyond the ordinary and everyday. This same significance
criterion applied equally to codes that related to pedagogical content knowledge
more than subject matter knowledge. In fact, there would be circumstances where
the “how to add 10” might be coded, but with a code more related to pedagogical
content knowledge.
It was important for us to keep in mind that our research focus was mathematics
content knowledge, and not other more general kinds of pedagogical awareness or
expertise. An initial ‘long list’ of codes was rationalised and reduced by negotiation
and agreement in the research team. Typically this came about through identifying
and unifying duplicate codes, and by eliminating the codes associated with events
that were agreed to be weak in significance.

Responding to children’s ideas. By way of illustration of this coding process, we


give here brief accounts of two episodes that we labelled with the code ‘Responding
to children’s ideas’5 (RCI). It will be seen that the contribution of a child, in each
case, was unexpected. Within the research team, this code name was understood to
be potentially ironic, since the observed response of the teacher to a child’s insight
or suggestion was often to put it to one side rather than to deviate from the planned
lesson ‘script’, even when the child offered further insight on the topic at hand.
Illustrative episode 1. Jason was teaching elementary fraction concepts to a Year
3 (pupil age 7–8) class. The pupils each had a small oblong whiteboard and a dry-
wipe pen. Jason asked them to “split” their individual whiteboards into two. Most
of the children predictably drew a line through the centre of the oblong, parallel to
one of the sides, but one boy, Elliott, drew a diagonal line. Jason praised him for his
originality, and then asked the class to split their boards “into four.” Again, most
children drew two lines parallel to the sides, but Elliott drew the two diagonals.
Jason’s response was to bring Elliott’s solution to the attention of the class, but to
leave them to decide whether it is correct. He asks:

111

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

Jason: What has Elliott done that is different to what Rebecca has done?
Sophie: Because he’s done the lines diagonally.
Jason: Which one of these two has been split equally? […] Sam, has Elliott
split his board into quarters?
Sam: Um … yes … no …
Jason: Your challenge for this lesson is to think about what Elliott’s done,
and think if Elliott has split this into equal quarters. There you go
Elliott.
At that point, Jason returned the whiteboard to Elliott, and the question of whether
it had been partitioned into quarters was not mentioned again. What makes this
interesting mathematically is the fact that (i) the four parts of Elliott’s board are not
congruent, (ii) they have equal areas, but (iii) this is not at all obvious. Furthermore,
(iv) an elementary demonstration of (ii) is arguably even less obvious. This seemed
to us a situation that posed very direct demands on Jason’s subject matter knowledge
and arguably pedagogical content knowledge too. It is not possible to infer whether
Jason’s “challenge” is motivated by a strategic decision to give the children some
thinking time, or because he needs some himself.
Illustrative episode 2. Naomi was introducing the subtraction ‘comparison’
structure (e.g., Carpenter & Moser, 1983) to a Year 1 class (pupil age 5–6). She
set up various comparison (or “difference”: see Rowland, 2006) problems, in the
context of frogs in two ponds. Magnetic ‘frogs’ are lined up on a board, in two
neat rows. In the first problem, Naomi says that her pond has four frogs, and her
neighbour’s pond has two. The class agreed that she had two more frogs than her
neighbour.6 Then Hugh offered the following thought:
Hugh: You could both have three, if you give one to your neighbour.
Like Jason, Naomi acknowledged the child’s idea, but in this case she dismissed
any further consideration of the alternative avenues that it could lead down.
Naomi: I could, that’s a very good point, Hugh. I’m not going to do that today
though. I’m just going to talk about the difference. Madeleine, if you
had a pond, how many frogs would you like in it?
One can readily sympathise with Naomi’s response to Hugh’s insight, which
seems to deviate too far from the agenda that she had set for the lesson. Naomi
acknowledges Hugh’s observation, but refuses to be diverted from her course.
The identification of opportunities to respond to children’s ideas (code –
responding to children’s ideas), whether taken or sidestepped in the videotaped
lessons, was not intended to suggest – and certainly not at the data analysis stage
being described here – what the teacher (prospective teacher in this case) should
or should not have done. I am not suggesting that every potential diversion
should be pursued. Before deviating from their plan, the teacher must make a
more-or-less instantaneous cost-benefit assessment of the outcome of doing so,

112

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

whether they feel sufficiently confident to depart from their ‘script’, and whether
the time available is sufficient to see the new venture through to a meaningful
conclusion.

THE EMERGENCE OF THE KNOWLEDGE QUARTET

In September, we presented some findings at a conference (Huckstep et al., 2002).


Our presentation focused on three of the 17 codes: namely choice of examples,
making connections and responding to children’s ideas. Perhaps as a consequence of
feedback received at the conference presentation, we began to realise the practical
potential of what we had ‘found’ in our research study for our work, and for that
of our colleagues, in teacher education. In the United Kingdom, a large part of
the graduate initial education (Postgraduate Certificate in Education) year is spent
teaching in schools under the guidance of a school-based mentor. The proportion of
time differs in other countries, but practicum placement is more-or-less universal.
Placement lesson observation is normally followed by a review meeting between a
school-based teacher-mentor and the prospective teacher. On occasion, a university-
based tutor will participate in the observation and the review. Research shows that
such meetings typically focus heavily on organisational features of the lesson, with
little attention to mathematical aspects of mathematics lessons (Brown, McNamara,
Jones, & Hanley, 1999; Strong & Baron, 2004). Our ‘pure’ research clearly offered
a basis for us to develop an empirically-based conceptual framework for lesson
reviews with a clearer focus on the mathematics content of the lesson and the role
of the prospective teacher’s mathematics subject matter knowledge and pedagogical
content knowledge. Such a framework would need to capture a number of important
ideas and factors about content knowledge within a small number of conceptual
categories, with a set of easily remembered labels for those categories.
The identification of the 17 categories could be a steppingstone in the development
of a framework for observing and reviewing mathematics teaching with prospective
teachers, and potentially not only these novices. We (the Cambridge team) did not
want a 17-point tick-list, like an annual car safety check, but a readily understood
scheme which would serve to frame an in-depth discussion between teacher and
observer. Our codes were useful to the extent that we had a set of concepts and
an associated vocabulary sufficient to identify and describe various ways in which
mathematics content knowledge plays out in elementary mathematics teaching. In
the autumn of 2003, we began to think about how we could group the codes into a
smaller set of ‘big ideas’ for mathematics teaching.
Eventually, we made a large paper label for each of the codes, spread these labels
out on the floor, and began to separate them into sets. Each suggestion for putting
two or more codes in the same subset had to be backed up with a reason of some sort.
For example, someone put choice of examples and choice of representation together,
saying that these were both ways that teachers use to make an abstract idea accessible
when they are teaching it. In fact, this person said, these two codes are characteristic

113

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

examples of the ways that teachers ‘transform’ their subject knowledge, as Shulman
(1987) put it, in order to help others to learn it. Another grouping included the three
codes: decisions about sequencing, anticipation of complexity, and recognition of
conceptual appropriateness. The last two of these capture instances of planning, or
of in-the-moment decision-making, that appeared to be informed by the teacher’s
awareness of the level of challenge and conceptual complexity entailed in the
mathematical subject matter in hand. This is a well-documented topic in the research
and professional literature, and much can be learned from experience too. Indeed,
as remarked earlier, the original conception of pedagogical content knowledge
“includes an understanding of what makes the learning of specific topics easy or
difficult …” (Shulman, 1986, p. 9). These kinds of teacher knowledge contribute
to decisions about the sequencing of instruction and student activity and the
sequencing of exercises – considerations encompassed in the first of the three codes
in this grouping. The crucial awareness and consensus, for us, was that these three
codes contribute to students’ sense of coherence within a lesson, and from lesson to
lesson: that the shifts of focus and activity were by design, and not by chance. As we
discussed and tried to capture this unifying quality with respect to these three codes,
the words, ‘coherence’, ‘cohesion’ and ‘connection’ came to mind.
By an extended process of argument, debate and negotiation, we eventually
agreed on grouping the 17 codes into four superordinate categories which,
together, we later called the Knowledge Quartet. Each of the four categories is a
unit, or dimension of the Knowledge Quartet. We have named the four dimensions:
Foundation; Transformation; Connection; Contingency. These four units represent
more comprehensive, higher-order concepts, in keeping with standard practice in
grounded theory research (Strauss & Corbin, 1998, pp. 113–114), which involves
first categorising the data (‘open’ coding), then connecting categories (‘axial’
coding’) before finally proposing a core category (selective coding). “First analyse,
then synthesise, and finally prioritise” (Dey, 1999, p. 98). This certainly describes
what we had done in grouping the 17 codes into four categories and conceiving the
whole (the Quartet) as a tool for focused mathematics lesson observation.
Not only the constituents, but also the names of these four Knowledge Quartet
categories were in flux for more than a year. In time, the components of four categories
had more or less stabilised but their names were still not settled. ‘Coherence’ was
originally chosen in preference to ‘Connection’, a term which had gained popularity
in the United Kingdom due to the Effective Teachers of Numeracy study (Askew et al.,
1997). Our conceptualisation of the corresponding unit of the Knowledge Quartet was
specific in its inclusion of codes related to the sequencing of instruction, and we wanted
to distinguish our notion from the ‘connectionist’ beliefs orientation which headlines
the Effective Teachers study. In the end, we decided that it would be sensible to ‘go
with the flow’ and explain our nuanced use of ‘connection’ as and when necessary.
We had reduced the names of all the dimensions to single words by the end of
2003 (Rowland, Huckstep, & Thwaites, 2003). A further refinement of the codes
was judged to be necessary in 2004 as we prepared the manuscript for a journal

114

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

article (Rowland et al., 2005). As we listed the elements of each dimension, we could
not avoid the feeling that a code that we had labelled ‘making connections’ (MC)
now seemed rather limp as a component (admittedly one of four) of the dimension
that we were calling Connection. Looking back at our data, we saw that the code
‘making connections’ had been used – in keeping with Askew et al. (1997) – to refer
to the participant teachers’ efforts (or missed opportunities) to make connections
between concepts, or to show how two procedures might be related. Consequently,
the rather tautologous making connections was subdivided into two codes: making
connections between concepts (MCC) and making connections between procedures
(MCP). I should emphasise that there were instances of both of these new codes
in our original data: it was not a case merely of imagining ways to make ‘making
connections’ more focused. Indeed, a clear instance of making connections between
procedures can be seen in the case of Laura (Rowland et al., 2004).
The substantive conceptualisation of the four dimensions of the Knowledge Quartet
at length can be found in Rowland et al. (2005). However, a brief characterisation
of each unit of the Knowledge Quartet is as follows. The first category, foundation,
consists of prospective teachers’ knowledge, beliefs and understanding acquired
‘in the academy’, in preparation (intentionally or otherwise) for their role in the
classroom. The key components of this theoretical background are: knowledge
and understanding of mathematics per se and knowledge of significant tracts of
the literature on the teaching and learning of mathematics, together with beliefs
concerning the nature of mathematical knowledge, the purposes of mathematics
education, and the conditions under which pupils will best learn mathematics.
The second category, transformation, concerns knowledge-in-action as
demonstrated both in planning to teach and in the act of teaching itself. As Shulman
indicates, the presentation of ideas to learners entails their re-presentation (our hyphen)
in the form of analogies, illustrations, examples, explanations and demonstrations
(Shulman, 1986, p. 9). Of particular importance is the prospective teachers’ choice
and use of examples presented to pupils to assist their concept formation, language
acquisition and to demonstrate procedures (Rowland, Thwaites, & Huckstep, 2003).
The third category, connection, binds together certain choices and decisions
that are made for the more or less discrete parts of mathematical content. In her
discussion of ‘profound understanding of fundamental mathematics’, Ma (1999,
p. 121) cites Duckworth’s observation that intellectual ‘depth’ and ‘breadth’ “is a
matter of making connections.” Our conception of this coherence also includes the
sequencing of material for instruction, and an awareness of the relative cognitive
demands of different topics and tasks.
Our fourth and final category, contingency, is witnessed in classroom events that
are almost impossible to plan for. In commonplace language it is the ability to ‘think
on one’s feet’. As indicated earlier, in the comments on episodes with Jason, and
Naomi, it includes the readiness to respond to children’s ideas and a consequent
preparedness, when appropriate, to deviate from an agenda set out when the lesson
was prepared. The Knowledge Quartet framework is summarised in Table 4.1.

115

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

Table 4.1. The Knowledge Quartet (adapted from Rowland et al., 2005)

The Knowledge Quartet

Foundation Propositional knowledge and beliefs concerning:


‡ the meanings and descriptions of relevant mathematical concepts, and of
relationships between them;
‡ the multiple factors which research has shown to be significant in the
teaching and learning of mathematics;
‡ the ontological status of mathematics and the purposes of teaching it.
Contributory codes: awareness of purpose; identifying errors; overt
subject knowledge; theoretical underpinning of pedagogy; use of
terminology; use of textbook; reliance on procedures.
Transformation Knowledge-in-action as revealed in deliberation and choice in planning
and teaching. The teacher’s own meanings and descriptions are
transformed and presented in ways designed to enable students to learn
it. These ways include the use of representations, analogies, illustrations,
explanations and demonstrations. The choice of examples made by the
teacher is especially visible:
‡ for the optimal acquisition of mathematical concepts and procedures;
‡ for confronting and resolving common misconceptions;
‡ for the justification (by generic example) or refutation (by counter-
example) of mathematical conjectures.
Contributory codes: choice of representation; teacher demonstration;
choice of examples.
Connection Knowledge-in-action as revealed in deliberation and choice in planning
and teaching. Within a single lesson, or across a series of lessons, the
teacher unifies the subject matter and draws out coherence with respect to:
‡ connections between different meanings and descriptions of particular
concepts or between alternative ways of representing concepts and
carrying out procedures;
‡ the relative complexity and cognitive demands of mathematical concepts
and procedures, by attention to sequencing of the content.
Contributory codes: making connections between procedures; making
connections between concepts; anticipation of complexity; decisions
about sequencing; recognition of conceptual appropriateness.
Contingency Knowledge-in-interaction as revealed by the ability of the teacher to
‘think on her feet’ and respond appropriately to the contributions made
by her students during a teaching episode. On occasion this can be seen
in the teacher’s willingness to deviate from her own agenda when to
develop a student’s unanticipated contribution might be of special benefit
to that pupil, or might suggest a particularly fruitful avenue of enquiry.
Contributory codes: responding to children’s ideas; use of opportunities;
deviation from agenda; teacher realisation.

116

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

THEORETICAL SAMPLING: THE DEMOCRATIC EVOLUTION OF THE


KNOWLEDGE QUARTET

Since its initial development, researchers in several countries have used the
Knowledge Quartet as a framework for the analysis of their own classroom data, and
some have corresponded with us about their findings, and shared their experiences
of the comprehensiveness of the theory in relation to their own data. Although
our experience to date indicates that the fundamental anatomy of the Knowledge
Quartet is complete, we take the view that the details of its component codes, and
the conceptualisation of each of its dimensions, are perpetually open to revision.
In grounded theory methodology, it is also inherent in the notion of ‘theoretical
sampling’ (Glaser & Strauss, 1967), whereby the application of the theory exposes
some shortcoming, and thereby lays it open to refinement, modification and possible
improvement until (perhaps) it achieves saturation.
As a consequence of this process, and the possibility of electronic global
communication, four additional codes have been added to the original 17. Either
the teacher behaviours captured in these codes were absent in our 2002–2003 video
data, or else we failed to note it in our analysis of that data. The names of the four
new codes, the Knowledge Quartet dimension which they enrich, the year in which
they emerged, and the researchers who brought them to our attention, are as follows:
‡ teacher insight during instruction (Contingency). 2005, Dolores Corcoran –
Ireland
‡ (mis)use of instructional materials (Transformation), 2006, Marilena Petrou –
Cyprus
‡ responding to the (un)availability of tools and resources (Contingency). 2009,
Libby Jared et al. – United Kingdom
‡ making connections between representations (Connection). 2015, Abraham de la
Fuente – Spain
An elaboration of the first and last of these codes now follows. Accounts of the other
two are given in Petrou (2010) and in Rowland et al. (2011), respectively.

Contingency: Teacher Insight during Instruction

The first instance of this incremental process of enrichment is the case of Máire, a
prospective teacher participant in the study of Dolores Corcoran, located in Ireland.
Máire was observed teaching a lesson on whole-number division to a class of girls
aged 9–10 years (see Corcoran, 2007). She had written worksheets on division, set
in a fantasy Harry Potter scenario. The first problem for one of the groups was as
follows:
Ron has 18 Galleons7 and a pack of cards costs 3 Galleons. How many packs
can he buy?

117

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

There are two principal division problem structures, variously called partition
(or sharing) and quotition (or measurement or grouping). In the problem under
discussion, the problem structure is quotition. Máire had provided butter beans as
manipulatives, to represent the Galleons. One pupil, Rosin, read out the problem,
while Megan volunteered to count out 18 butter beans.
Máire offered a few words of explanation about the “wizard money,” then she
asked:
Máire: How many groups does she [Megan] need to break it into and can
you tell me why? Hannah, what do you think?
Hannah: Into three groups.
Máire: Into three groups. Well done, and why? You can read the question
again if you want.
Máire’s query here about the number of groups, and not to their size, points
inappropriately to a partition structure, and this is picked up by Hannah. Máire
congratulates the child (“Well done”) on her inappropriate suggestion. Máire asks
Hannah to explain (“and why?”), and the interaction then takes a different direction.
Hannah: Because there’s three packs of cards.
Máire: It’s not that there’s three packs of cards. But what is it about the
cards?
Hannah: It costs three galleons.
Máire is pulled up short at this point. She knows that there are not three packs of
cards. Máire has inadvertently directed the pupils to the wrong division structure,
she realises that this is so, and she resolves to find a way out:
Máire: It costs three galleons. […] You’ve got 18 and what are you doing?
Máire is attempting to alter the direction of the discussion, but the child who
answers has not altered course:
Child: Splitting them up into three groups …
Maire responds with a direct correction, and her language is now correctly aligned
with quotition/grouping
Máire: Ahh …? Into groups of three [she nods]. And how many groups do
you have?
Child: Six.
Máire: So how many packs of cards could Ron buy?
Child: Six.
Máire: He could buy six packs of cards. Can everybody follow that? What
sentence would you write to explain what we just did?
This ability to change course as a result of reflection had not been noted in the
lessons that were the data for our original study. We see an instance of reflection-

118

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

in-action (Schön, 1983) in this episode, and in what we would call a ‘contingent
moment’. Máire could not have prepared (in her planning) for what she did at that
moment, but what she did say and do brought about a significant and pedagogically
important shift in the discourse and the cognitive content of the lesson. This was
possible because Máire seems to have experienced a pedagogical insight of some
kind. The Contingency dimension of the Knowledge Quartet was rooted, as it arose
from the data in our original study, in the teacher’s response to children’s insights
and misconceptions. In this instance we seem to have a moment where Máire
herself suddenly realises that the problem, the child’s suggestion, and her approval,
are in contradiction. Máire’s moment of insight is an instance where theoretical
sampling found the existing Knowledge Quartet theory to be wanting and caused it
to be rethought and enhanced. Consequently, we added an additional code – teacher
insight during instruction (TII) – to those previously associated with Contingency.

Connection: Making Connections between Representations

Our second instance of theory evolution draws on the doctoral study of Abraham de
la Fuente in Spain. The participants were mathematics teachers working in the first
two years of the secondary stage. Specifically, this research aimed to understand how
the teachers used their knowledge to help students to learn to use algebraic language
in a problem-solving environment. In the episode described here the intention was
that students would learn to solve simultaneous linear equations by engaging with
iconic, algebraic and tabular representations of key information. For example, the
equation 3a + 3b =12 was represented initially in a picture of 3 slices of pizza and 3
drinks costing 12 euros. Various student responses included listing various prices of
drinks and slices that would satisfy each of the two equations.
After working on several problems like this, the teachers devised a ‘test’ for the
students, consisting of three simultaneous linear equations problems. The second
gave 3x + y = 55 and 2x + 2y = 62 in precisely that symbolic form; the first was
isomorphic to it, but with an iconic form (involving two different types of ‘Star
Wars’ figures and total costs in euros). The third was a different pair of equations in
conventional symbolic form only.
After the students had spent 20 minutes or so working on the problems, the
teacher led a whole-class discussion about solving the first two problems, drawing
out the fact that problem 2 is the ‘same’ problem as problem 1.
For further details and a Knowledge Quartet analysis of the lesson, see de la Fuente
et al. (2016). On the basis of this analysis, de la Fuente proposed the additional code
making connections between representations in the Connection dimension of the
Knowledge Quartet.
Once this code had been brought to light, Fay Turner (a participant in our 2002–
2003 study) was able to identify instances of it in the data in her own doctoral
study (Turner, 2010). For example, in a lesson on the comparison (or ‘difference’)
subtraction structure with a Year 2 (pupil age 6–7) class, Kate began by comparing

119

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

the heights of two towers of interlocking ‘multilink’ cubes. She then displayed two-
dimensional representations of pairs of towers of cubes on the classroom interactive
whiteboard, before showing pairs of lines numbered, respectively, to 5 and to 9, on
the interactive whiteboard. In this way she made connection between enactive, iconic
and symbolic representations (Bruner, 1974) of the difference between 5 and 9.

Conclusion

While the Cambridge-based originators of the Knowledge Quartet cannot possibly


claim ‘ownership’ of the theory, we take a deep interest in proposals to develop
it, of the kind described above. Having said that, we think it only proper that such
proposals be empirically-based outcomes of theoretical sampling as opposed to the
result of speculations about what codes could be added to those (21 so far) that have
emerged from focused analysis of classroom data.
Table 4.2 summarises the Knowledge Quartet dimensions and their contributory
codes, at the time of writing. The 17 shown in ‘normal’ font emerged in our analysis
of the 24 lessons in 2002–2003. Those shown in italics came from the process of
theoretical sampling described above.

Table 4.2. The Knowledge Quartet – Dimensions and contributory codes

Dimension Contributory codes

Foundation: awareness of purpose; adheres to textbook;


knowledge and understanding of concentration on procedures; identifying
mathematics per se and of mathematics- errors; overt display of subject knowledge;
specific pedagogy; beliefs concerning theoretical underpinning of pedagogy; use of
effective mathematics instruction, the mathematical terminology
nature of mathematics, and the purposes of
mathematics education.
Transformation: choice of examples; choice of representation;
the presentation of ideas to learners in the (mis)use of instructional materials; teacher
form of analogies, illustrations, examples, demonstration (to explain a procedure)
explanations and demonstrations
Connection: anticipation of complexity; decisions about
the sequencing of material for instruction, sequencing; recognition of conceptual
and an awareness of the relative cognitive appropriateness; making connections
demands of different topics and tasks between procedures; making connections
between concepts; making connections
between representations
Contingency: deviation from agenda; responding to
the ability to make cogent, reasoned and students’ ideas; use of opportunities; teacher
well-informed responses to unanticipated insight during instruction; responding to the
and unplanned events (un)availability of tools and resources

120

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

FINAL REMARKS

From Observation to Evaluation

In a paper presented to a meeting of teacher education researchers, Skott (2006)


highlighted the changing nature of the relationship between theory and practice
in teacher education. He observed that the theories brought to bear on the task of
improving teaching increasingly derive from studies of teaching and coined the term
‘theoretical loop’ to capture this dialectical relationship between theory and practice
in teacher education. In the case of the research which has been the focus of this
chapter, the Knowledge Quartet came about as the outcome of systematic analysis of
mathematics teaching. Initially, we viewed it as a way of managing the complexity of
describing the role of teachers’ content knowledge in their teaching. In keeping with
Skott’s theoretical loop, we subsequently developed ways of using the Knowledge
Quartet as a framework to facilitate analysis and discussion of mathematics teaching
among prospective teachers, their mentors and teacher educators (see Rowland &
Turner, 2006).
The progression from observation of teaching to its description and analysis is
clear, but, thus far, I have been less explicit about the evaluation of teaching. In the
spirit of reflective practice, the most important evaluation must be that of the teacher
him/herself. However, this self-evaluation is usefully provoked and assisted by a
colleague or mentor. In a number of papers (e.g., Rowland et al., 2004; Rowland &
Turner, 2006) we have exemplified this provocation through the identification, using
the Knowledge Quartet, of tightly-focused discussion points to be raised in a post-
observation review. We have suggested that these points be framed in a relatively
neutral way, such as “Could you tell me why you etc?” or “What were you thinking
when etc?” It would be naïve, however, or a kind of self-delusion, to suggest
that the mentor, or teacher educator, makes no evaluation of what they observe.
Indeed, the observer’s evaluation is likely to be a key factor in the identification and
prioritisation of the discussion points. In post-observation review, it is expected that
the ‘more knowledgeable other’ will indicate what the novice did well, what they did
not do and might have, and what they might have done differently. The Knowledge
Quartet is a framework to organise such evaluative comments, and to identify ways
of learning from them.

Methodology
In his China Lectures, Hans Freudenthal (1991) ranted against the (then) new breed
of professional methodologists. His words capture my own experience of research
‘design’ far better than I could express it myself:
I don’t remember when it happened but I do remember, as though it were
yesterday, the bewilderment that struck me when I first heard that the training
of future educationalists includes a course on “methodology.” This is at any

121

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

rate the custom in our country but, judging from the literature in general, this
brain-washing policy is an international feature. Please imagine a student of
mathematics, of physics, of – let me be cautious, as I am not sure how far
this list extends – impregnated, in any other way than implicitly, with the
methodology of the science that he sets out to study; in any other way than by
having him act out the methodology that he has to learn! In no way do I object
to a methodology as such – I have even stimulated the cultivation of it, but it
should be the result of a posteriori reflecting on one’s methods, rather than an
a priori doctrine that has been imposed on the learner. (Freudenthal, 1991, pp.
150–151)
Many readers will have little sympathy for Freudenthal’s suggestion that research
methodology might be the outcome of reflection on research action, arguing that
educational research has reached new heights of scientific sophistication since
Freudenthal composed this diatribe against “the pure methodologists, whose strength
consists in knowing all about research and nothing about education” (op. cit.). At the
same time, Freudenthal’s version of events agrees reasonably well with my own
experience. Much of the account that I have given of the research processes that the
SKIMA team followed in arriving at what we came to call the Knowledge Quartet
has been possible with the benefit of hindsight, “the result of a posteriori reflecting
on one’s methods.” I have done my best to be true to our intellectual and practical
experience, as documented and remembered, rather than to offer some idealised,
even sanitised, version of events.

Applications and Next Steps

Research originally fuelled by curiosity about teacher knowledge and classroom


practices led to the development of the Knowledge Quartet, a manageable
framework within which to observe, analyse and discuss mathematics teaching
from the perspective of teachers’ mathematical content knowledge, both subject
matter knowledge and pedagogical content knowledge. Those who use it need to
be acquainted with the details of its conceptualisation, because mere labels such as
‘connection’ may, for each individual, connote meanings other than those intended.
Initial indications are that this development has been well-received by teacher-
mentors, who appreciate the specific focus on mathematics content and pedagogy.
They observe that it compares favourably with government guidance on mathematics
lesson observation, which focuses on more generic issues such as “a crisp start, a
well-planned middle and a rounded end. Time is used well. The teacher keeps up a
suitable pace and spends very little time on class organisation, administration and
control” (Department for Education and Employment, 2000, p. 11).
It is all too easy for analysis of a lesson taught by a novice teacher to be (or
to be perceived to be) gratuitously critical, and it is important to emphasise that
the Knowledge Quartet is intended as a tool to support teacher development, with

122

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

a sharp and structured focus on the impact of their subject matter knowledge and
pedagogical content knowledge on teaching. The post-observation review meeting
usefully focuses on a lesson fragment, and on only one or two dimensions of the
Knowledge Quartet, to avoid overloading the prospective teacher with action points.

Using the Knowledge Quartet in other Contexts and in Other Places


Although the Knowledge Quartet came into being in a primary/elementary context
in the United Kingdom, it has been tested and found to be relevant and useful in
other contexts. This is inherent in the earlier account of Theoretical Sampling in
connection with the Knowledge Quartet.
For example, Karagöz-Akar (2016) used the Knowledge Quartet to investigate
the relationship between Turkish prospective secondary mathematics teachers’
perspectives and their mathematical knowledge for teaching. Two prospective
secondary mathematics teachers were observed, one teaching geometry, the other
algebra. These two teachers were selected for observation because they were
judged to hold a Progressive Incorporation Perspective on mathematics teaching,
viewing mathematics learning as an active mental process. It was found that they
demonstrated “all the aspects of mathematical knowledge for teaching (Rowland
et al., 2005; Thwaites et al., 2011) and vice versa. Particularly, data showed the
reasons behind the nature of prospective teachers’ foundational knowledge and other
domains in Knowledge Quartet” (Karagöz-Akar, 2016, pp. 20–21). Data samples
given in the paper include a lesson in which Sarah (the prospective teacher) asks
her class of Grade 10 students to solve sin(x) = ½. Suffice it here to say that they
have difficulty finding solutions other than x = 300. Sarah is driven to respond to
this Contingent situation, specifically responding to students’ ideas (RSI). Karagöz-
Akar also readily identifies three codes from the Foundation dimension in Sarah’s
behaviour – namely awareness of purpose, identifying errors, and overt display of
subject knowledge.
In a study of university mathematics teaching in Ireland, Breen, Meehan and
O’Shea use the Knowledge Quartet framework to analyse accounts of their practice
as mathematics lecturers. They had each written accounts of their teaching of one
module: two were in Real Analysis and one in Differential Calculus. Their purpose
was to investigate what insights the Knowledge Quartet analysis might add to their
earlier investigation of lecturers’ decision-making when teaching (Breen et al., 2014).
They felt the need to adapt the scope of a few of the Knowledge Quartet codes: for
example, teacher demonstration (to explain a procedure) was extended to encompass
teacher demonstration to explain a proof. The accounts of all three lecturers were
often about giving a task to the class or instigating a whole class discussion around a
task, recording some students’ responses (or lack of responses) to the task, including
the lecturer’s reflections about student thinking. Choice of Examples was the most
frequently occurring code for all three lecturers. Thus, in a session focused on the

123

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

relationship between bounded and convergent sequences, Lecturer 1 instructed the


class to give examples of bounded and convergent sequences, and to consider which
were both. They conjectured that every convergent sequence is bounded but found
bounded sequences that were not convergent. This is an interesting case of choice
of example since, in this case, the students are choosing the examples. It also seems,
however, to be a good case of Theoretical Underpinning of Pedagogy (Foundation
Dimension). The authors conclude that
use of the KQ [Knowledge Quartet] as a reflection tool could afford
mathematicians (with no formal pedagogical training) an opportunity to
develop pedagogical knowledge. It is interesting to contrast this with Turner
and Rowland’s (2011) finding that the KQ afforded preservice primary
teachers (typically non-specialists in mathematics) an opportunity to develop
mathematical content knowledge. (Breen et al., 2018, p. 391)
In a study in Australia, Sharyn Livy used the Knowledge Quartet to analyse a
prospective teacher’s management of a Bingo game with third grade students. The
teacher (Lisa) gave a subtraction problem to identify each of the numbers for the
students to find (or not) on their cards. Subtraction problems given (for the student
ts to answer and find on their Bingo board) included: 20 – 8, 14 – 2, 14 – 4, 28 – 8,
8 – 3. Lisa advised the class:
If you want to use some scrap piece of paper and write it out you can do that or
if you want to do it mentally, you might count by twos or fives. If you need to
draw a number line you can do that. (Livy, 2010, p. 346)
Livy uses the Knowledge Quartet dimensions and codes to analyse (and evaluate)
Lisa’s planning and execution of the lesson, focusing on her subject matter knowledge
and pedagogical content knowledge and their application. She thereby identifies a
number of events in the lesson which could be discussed with Lisa and modified in
the future. These include, for example, her introduction of a vertical ‘subtraction
ladder’ intended to assist the subtractions.
Finally, our teacher education colleagues with different subject specialisms –
including English, science and modern foreign languages – tell us that they find the
Knowledge Quartet relevant and useful in relation to their own work, especially in
their lesson observations and review meetings. What might the conceptualisations of
the dimensions of the Knowledge Quartet look like in these and other disciplines? In
any case, it would seem that the cross-disciplinary character of Shulman’s ground-
breaking 1980s insights assures the relevance and usefulness of the Knowledge
Quartet beyond the domain in which it was first developed. This would be fertile
ground for research into the application of the Knowledge Quartet across the
curriculum.

124

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

NOTES
1
By ‘content’ knowledge, we include any kind of disciplinary (in this case mathematics-related)
knowledge. In particular, ‘content’ knowledge encompasses both subject matter knowledge and
pedagogical content knowledge.
2
Slightly preceding hand-held digital cameras.
3
What our colleagues in other parts of Europe would more likely call ‘didactics’.
4
We probably depart from Ball et al. (2008) here, in that they seem to regard these aspects of mathematics
content knowledge as aspects of specialized content (i.e., subject matter, for them) knowledge. This is
unimportant, for the present purpose at least: we could re-frame our observation by saying that what
Ball et al. call common content knowledge was rarely an issue, either for concern or celebration, in
the 24 lessons we analysed.
5
This later became Responding to Students’ Ideas (RSI) to include learners of all ages.
6
We reflected that this particular example, with equal subtrahend and difference, is pedagogically
problematic (see e.g. Rowland, Thwaites and Huckstep, 2003).
7
The Harry Potter novels by J. K. Rowling are well-known in Ireland. Galleons are the fictional
currency in use at Hogwarts – Harry Potter’s school.

REFERENCES
Askew, M., Brown, M., Rhodes, V., Wiliam, D., & Johnson, D. (1997). Effective teachers of numeracy.
London: King’s College, University of London.
Ball, D. L. (1990). Prospective elementary and secondary teachers’ understanding of division. Journal for
Research in Mathematics Education, 21(2), 132–144.
Ball, D. L., & Bass, H. (2003). Toward a practice-based theory of mathematical knowledge for teaching.
In B. Davis & E. Simmt (Eds.), Proceedings of the 2002 Annual Meeting of the Canadian Mathematics
Education Study Group (pp. 3–14). Edmonton, Canada: Canadian Mathematics Education Study
Group.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Bishop, A. J. (2001). Educating student teachers about values in mathematics education. In F.-L Lin &
T. J. Cooney (Eds.), Making sense of mathematics teacher education (pp. 233–246). Dordrecht:
Kluwer Academic Publishers.
Breen, S., McCluskey, A., Meehan, M., O’Donovan, J., & O’Shea, A. (2014). A year of engaging with the
discipline of noticing: Five mathematics lecturers’ reflections. Teaching in Higher Education, 19(3),
289–300, doi:10.1080/13562517.2013.860107
Breen, S., Meehan, M., O’Shea, A., & Rowland, T. (2018). An analysis of university mathematics teaching
using the Knowledge Quartet. In V. Durand-Guerrier, R. Hochmuth, S. Goodchild, & N. M. Hogstad
(Eds.), Proceedings of the Second Conference of the International Network for Didactic Research in
University Mathematics (pp. 383–392). Kristiansand: University of Agder and INDRUM.
Brown, T., Mcnamara, O., Jones, L., & Hanley, U. (1999). Primary student teachers’ understanding of
mathematics and its teaching. British Education Research Journal, 25(3), 299–322.
Calderhead, J. (1981). Stimulated recall: A method for research on teaching. British Journal of Educational
Psychology, 51(2), 211–217.
Carpenter, T. P., & Moser, J. M. (1983). The acquisition of addition and subtraction concepts.
In R. Lesh & M. Landau (Eds.), The acquisition of mathematical concepts and processes (pp. 7–44).
New York, NY: Academic Press.
Corcoran, D. (2007). “You don’t need a tables book when you have butter beans!” Is there a need for
mathematics pedagogy here? In D. Pitta-Pantazi & G. Philippou (Eds.), Proceedings of the Fifth
Congress of the European Society for Research in Mathematics Education (pp. 1856–1865). Nicosia,
Cyprus: ERME.
Department for Education and Employment. (1997). Teaching: High status, high standards: Circular
10/97. London: HMSO.

125

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

Department for Education and Employment. (1998). Teaching: High status, high standards: Circular
4/98. London: HMSO.
Department for Education and Employment. (1999). The national numeracy strategy: Framework for
teaching mathematics from Reception to Year 6. Sudbury: Department for Education and Employment
Publications.
Department for Education and Employment. (2000). Auditing mathematics in your school. Sudbury:
Department for Education and Employment Publications.
Dey, I. (1999). Grounding grounded theory: Guidelines for qualitative inquiry. London: Academic Press.
Dickson, L., Brown, M., & Gibson, O. (1984). Children learning mathematics. Eastbourne: Holt
Education.
Freudenthal, H. (1991). Revisiting mathematics education: The China lectures. Dordrecht: Kluwer
Academic Publishers.
Glaser, B. G., & Strauss, A. L. (1967). The discovery of grounded theory: Strategies for qualitative
research. New York, NY: Aldine de Gruyter.
Goulding, M., Rowland, T., & Barber, P. (2002). Does it matter? Primary teacher trainees’ subject
knowledge in mathematics. British Educational Research Journal, 28(5), 689–704.
Goulding, M., & Suggate, J. (2001). Opening a can of worms: Investigating primary teachers’ subject
knowledge in mathematics. Mathematics Education Review, 13, 41–54.
Gray, E. M., & Tall, D. O. (1994). Duality, ambiguity and flexibility: A proceptual view of simple
arithmetic. Journal for Research in Mathematics Education, 25(2), 115–141.
Huckstep, P., Rowland, T., & Thwaites, A. (2002, September). Primary teachers’ mathematics content
knowledge: What does it look like in the classroom? Symposium paper presented at the Annual
Conference of the British Educational Research Association, University of Exeter, London.
Huckstep, P., Rowland, T., & Thwaites, A. (2003). Observing subject knowledge in primary mathematics
teaching. In BSRLM (Eds.), Proceedings of the British Society for Research into Learning Mathematics
(Vol. 23, Issue 1, pp. 37–42). South Yorkshire: BSRLM.
Karagöz-Akar, G. (2016). Prospective secondary mathematics teachers’ perspectives and mathematical
knowledge for teaching. Eurasia Journal of Mathematics, Science & Technology Education, 12(1),
3–24
Lakatos, I. (1976). Proofs and refutations. Cambridge: Cambridge University Press.
Livy, S. (2011). A ‘Knowledge Quartet’ used to identify a second-year pre-service teacher’s primary
mathematical content knowledge. In L. Sparrow, B. Kissane, & C. Hurst (Eds.), Proceedings of the
33rd Annual Conference of the Mathematics Education Research Group of Australasia (pp. 344–351).
Fremantle: MERGA.
Ma, L. (1999). Knowing and teaching elementary mathematics: Teachers’ understanding of fundamental
mathematics in China and the United States. London: Lawrence Erlbaum Associates.
McNamara, D. (1991). Subject knowledge and its application: Problems and possibilities for teacher
educators, Journal of Education for Teaching, 17(2), 113–128.
Morris, H. J. (2001). Issues raised by testing trainee primary teachers’ mathematical knowledge.
Mathematics Teacher Education and Development, 3, 37–48.
Petrou, M. (2010). Adapting the Knowledge Quartet in the Cypriot mathematics classroom. In
V. Durand-Guerrier, S. Soury-Lavergne, & F. Arzarello (Eds.), Proceedings of the Sixth Congress of
the European Society for Research in Mathematics Education (pp. 2020–2029). Lyon, France: Institut
National de Recherche Pédagogique and ERME.
Rowland, T. (2006). Subtraction – Difference or comparison? Mathematics in School, 35(2), 32–35.
Rowland, T. (2007). Developing knowledge for mathematics teaching: A theoretical loop. In S. Close,
D. Corcoran, & T. Dooley (Eds.), Proceedings of the Second National Conference on Research in
Mathematics Education (pp. 13–26). Dublin: St Patrick’s College.
Rowland, T. (2009). Beliefs and actions in university mathematics teaching. In M. Tzekaki,
M. Kaldrimidou, & C. Sakonidis (Eds.), Proceedings of the 33rd Conference of the International
Group for the Psychology of Mathematics Education (Vol. 5, pp. 17–24). Thessaloniki: PME.

126

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RESEARCHING MATHEMATICAL KNOWLEDGE IN TEACHING

Rowland, T., Huckstep, P., & Thwaites, A. (2003). The Knowledge Quartet. In BSRLM (Eds.), Proceedings
of the British Society for Research into Learning Mathematics). South Yorkshire: BSRLM.
Rowland, T., Huckstep, P., & Thwaites, A. (2004). Reflecting on prospective elementary teachers’
mathematics content knowledge: The case of Laura. In M. J. Høines & A. B. Fugelstad (Eds.),
Proceedings of the 28th Conference of the International Group for the Psychology of mathematics
education (Vol. 4, pp. 121–128). Bergen: Bergen University College.
Rowland, T., Huckstep, P., & Thwaites, A. (2005). Elementary teachers’ mathematics subject knowledge:
The Knowledge Quartet and the case of Naomi. Journal of Mathematics Teacher Education, 8(3),
255–281.
Rowland, T., Martyn, S., Barber, P., & Heal, C. (2000). Primary teacher trainees’ mathematics subject
knowledge and classroom performance. In T. Rowland & C. Morgan (Eds.), Research in Mathematics
Education (Vol. 2, pp. 3–18). London: BSRLM.
Rowland, T., Martyn, S., Barber, P., & Heal, C. (2001). Investigating the mathematics subject matter
knowledge of pre-service elementary school teachers. In M. van den Heuvel-Panhuizen (Ed.),
Proceedings of the 25th Conference of the International Group for the Psychology of mathematics
education (Vol. 4, pp. 121–128). Utrecht: PME.
Rowland, T., Thwaites, A., & Huckstep, P. (2003). Novices’ choice of examples in the teaching of
elementary mathematics. In A. Rogerson (Ed.), Proceedings of the International Conference on the
Decidable and the Undecidable in mathematics education (pp. 242–245). Brno, Czech Republic: The
Mathematics Education into the 21st Century Project.
Rowland, T., Thwaites, A., & Jared, L. (2011). Triggers of contingency in mathematics teaching.
In B. Ubuz (Ed.), Proceedings of the 35th Conference of the International Group for the Psychology
of Mathematics Education (Vol. 4, pp. 73–80). Ankara: PME.
Rowland, T., & Turner, F. (2006). A framework for the observation and review of mathematics teaching.
Mathematics Education Review, 18, 3–17.
Sanders, S., & Morris, H. (2000). Exposing student teachers’ content knowledge: Empowerment or
debilitation? Educational Studies, 26(4), 397–408.
Schön, D. (1983). The reflective practitioner: How professionals think in action. New York, NY: Basic
Books Inc.
Schwab, J. J. (1978). Education and the structure of the disciplines. In I. Westbury & N. J. Wilkof (Eds.),
Science, curriculum and liberal education (pp. 229–272). Chicago, IL: University of Chicago Press.
Shulman, L. (1986). Those who understand, knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Shulman, L. S. (1987). Knowledge and teaching: Foundations of the new reform. Harvard Educational
Review, 57(1), 1–22.
Skott, J. (2006). The role of the practice of theorising practice. In M. Bosch (Ed.), Proceedings of the
Fourth Congress of the European Society for Research in Mathematics Education (pp. 1598–1608).
Barcelona: FUNDEMI IQS, Universitat Ramon Llull.
Strong, M., & Baron, W. (2004). An analysis of mentoring conversations with beginning teachers:
Suggestions and responses. Teaching and Teacher Education, 20(1), 47–57.
Thwaites, A., Huckstep, P., & Rowland, T. (2005). The Knowledge Quartet: Sonia’s reflections.
In D. Hewitt & A. Noyes (Eds.), Proceedings of the Sixth British Congress of mathematics education
(pp. 168–175). London: BSRLM.
Thwaites, A., Jared, L., & Rowland, T. (2011). Analysing secondary mathematics teaching with the
Knowledge Quartet. In BSRLM (Eds.), Proceedings of the British Society for Research into Learning
Mathematics (Vol. 30, Issue 3, pp. 85–90). Cambridge: BSRLM.
Teacher Training Agency. (2002). Qualifying to teach: Professional standards for the award of qualified
teacher status. Handbook of guidance. London: Teacher Training Agency.
Turner, F. (2007). Development in the mathematics teaching of beginning elementary school teachers:
An approach based on focused reflections. In S. Close, D. Corcoran, & T. Dooley (Eds.), Proceedings
of the Second National Conference on Research in Mathematics Education (pp. 377–386). Dublin:
St Patrick’s College.

127

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TIM ROWLAND

Turner, F. (2010). Using the knowledge Quartet to develop early career primary teachers’ mathematical
content knowledge: A longitudinal study (Unpublished PhD thesis). University of Cambridge, London.
Vergnaud, G. (1983). Multiplicative structures. In R. Lesh & M. Landau (Eds.), Acquisition of mathematics
concepts and processes (pp. 127–175). New York, NY: Academic Press.

Tim Rowland
Faculty of Education
University of Cambridge
and
Faculty of Education
Norwegian University of Science and Technology

128

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PART 2
MATHEMATICS TEACHER BELIEFS ABOUT
MATHEMATICS AND ITS TEACHING

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI, SILVIA FUNGHI AND
ALESSANDRO RAMPLOUD

5. MATHEMATICS TEACHERS’
CULTURAL BELIEFS
The Case of Lesson Study

The aim of this chapter is to discuss teachers’ cultural beliefs, that is, the ones which
are strongly dependent on teachers’ culture. We have identified a popular model
of teacher education and development – Lesson Study – one that has the potential
of raising the issue of cultural beliefs. This chapter contains a short review of the
lesson study literature, a short discussion of the deep roots of lesson study in the
Confucian Heritage Culture, the presentation of the theoretical construct of cultural
transposition useful in interpreting/designing programs of lesson study outside the
original context and the early findings of a programme in mathematics teacher
education and development through lesson study that has been realized in Italy.
We also address the features of the adaptation of the original model of lesson study
in the West and the possible construction of a theory of lesson study suitable for
Western approaches.

INTRODUCTION

According to Baba (2007), a Lesson Study cycle


consists in preparation, actual class and class review sessions […]. This
process begins with finding and selecting materials relevant to the purpose
of, and is then followed by refining the class design based on the actual needs
of, the students and tying all this information together into a lesson plan. The
significance of [lesson study] is that all of these processes are performed
in collaboration with other teachers. A classroom is then taught based on
the teaching plan devised. The class is observed by many teachers, who are
sometimes joined by university instructors and supervisors from the board of
education, and a review session is held for all observers after the class. (p. 2)
The following is a short excerpt from an essay by an Italian teacher after
participating in a lesson study activity for some months. The essay was prepared
as a partial fulfillment of her compulsory in-service education in the first year of
enrolment as a primary school teacher.

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_006

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

The added value of the [lesson study] as an educational opportunity for novice
teachers consists exactly in its power to destabilise teachers’ beliefs, and so to
reduce the risk to oversimplify mathematical content. This work methodology
invites teachers to rethink their profession from a new viewpoint, and it is
suitable for lessons about every subject, since it is based on teaching intentions
and learning processes, on an idea of learning as a collective construction of
meaning. […]
The teacher is adopting a position on some issues that will be reconsidered later in the
chapter. Even a short experience of lesson study activity has prompted an awareness
in this teacher about her profession and a reflection on her past experiences, hence
the discussion of her own development too. The excerpt from her essay summarises
some beliefs that many of the other teachers exposed in their short essays. In
consequence, their thinking throws light on the culture of teaching shared by the
participants in a programme of teacher development, addressing the introduction of
lesson study in dozens of Italian schools. This is our starting point.
We claim on a more general level that when teachers in a certain culture
encounter tools or teaching methods originating from a different culture, they
can derive profound reflections on their own way of conceiving didactics from
this meeting. In particular, we believe that experimental methods from another
culture, such as lesson study, can make teachers more aware of their beliefs about
teaching and learning that are dependent on the socio-cultural context to which they
themselves belong. This awareness, in some instances, can also lead to a change in
their perspectives on teaching and learning. We elaborate on this in later sections
of the chapter.

LITERATURE REVIEW

Lesson study reflects a model of teacher education and development initiated in


Japan many decades ago. Later, it appeared all over the world as one of the most
widespread models of teacher education and development. In 2006, a large project
about lesson study was launched by the Asia-Pacific Economic Cooperation (APEC),
an international forum which includes more than twenty economies and nearly three
billion inhabitants (see http://www.criced.tsukuba.ac.jp/math/apec/). At least twice
a year, conferences were held alternatively either in Japan or in Thailand. The World
Association for Lesson Studies (walsnet.org), was established in 2007 with annual
conferences and with a specific journal (The International Journal of Lesson and
Learning Studies) from 2011. According to Takahashi (in Isoda, Stephens, Ohara, &
Miyakawa, 2007, p. 194), in the United States a large movement was started for the
dissemination of lesson study with more than 140 lesson study groups in more than
twenty-nine states. The assistance for developing countries in Asia and Africa was
introduced as a major perspective into the lesson study; thanks to the involvement in
APEC project, as stated by Baba, Ueda, Ninomiya, and Hino (2018).

132

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

Several volumes and collections of papers about lesson study have appeared
in the last decades. We can mention at least the following volumes: Fernandez
& Yoshida, 2004; Hart, Alston, & Murata, 2011; Isoda et al., 2007; Lewis, 2002;
Inprasitha, Masami, & Ban-Har, 2015; Wang-Iverson & Yoshida, 2005. The Journal
of Mathematics Teacher Education also has published several papers on lesson study
(e.g., Inoue, 2011; Lewis, Perry, & Hurd, 2009; Ricks, 2011). A recent survey (Robutti
et al., 2016) that includes a focus on lesson study has been published on behalf of the
13th International Congress of Mathematics Education (ICME-13). After this survey,
a new International Commission on Mathematical Instruction (ICMI) study entitled
Mathematics teachers working and learning in collaborative groups, explicitly
referring to lesson study, was launched (see https://www.mathunion.org/icmi/icmi-
news-november-2018). In spite of this dissemination, there is only one chapter on
lesson study, by Yoshida (pp. 85–106), in the second volume of the International
Handbook of Mathematics Teacher Education (Sullivan, Wood, & Tirosh, 2008).
The volume Mathematical lesson study around the world, edited by Quaresma
et al. (2018), aims at reporting the outcomes of a discussion group held at ICME-
13. The editors summarise the aims of the discussion group and of this volume as
follows:
The discussion group had two main foci of work. The first was dedicated to
presenting and discussing regional and national particularities and approaches
of lesson and learning studies in mathematics in Japan and other Asian
countries, North and South America and Europe. The second was dedicated
to discussing theoretical, methodological, and epistemological issues involved
in organizing and carrying out research on and with lesson and learning study.
(pp. vii–viii)
In the same volume, Winsløw, Bahn, and Rasmussen (2018) report on a
comprehensive review of the literature on lesson study from the second half of the
1990s. The review (updated March 2017) identified between 165 papers and a few
less than 300 papers (the former number referring to the papers with ‘lesson study’
as a part of the title and the latter to papers which have some connection with the
topic). This large dissemination, promoted by Japanese experts, required adaptations
to local contexts, given that many realisations diverged substantially from Japanese
lesson study (Lewis, 2004, as mentioned by Winsløw et al., 2018). This difference
may be related to many factors, such as school organisation, standards, teachers’
preparation, and so on, but we claim that cultural influence (and the teachers’ cultural
beliefs) plays a role that is much larger than expected.
Recently, another model of lesson study appeared in the literature: the Chinese
lesson study. A special issue of the International Journal of Lesson and Learning
Studies (issue 4, 2017) was devoted to the Theory and practice of Chinese lesson study,
and a symposium at the American Educational Research Association 2018 conference
was organized on Chinese lesson study: Theories, practices, and its adaptation,
with a reaction by Alan Schoenfeld. There are differences between the Japanese and

133

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

Chinese lesson study (see, for instance, Bartolini Bussi, Bertolini, Ramploud, & Sun,
2017), although they share the same Confucian Heritage Culture, discussed in the
next section after some consideration of what constitutes the construct of beliefs.
The problem of identifying a shared definition for the construct of beliefs has been
widely discussed in mathematical education. However, there is no internationally
accepted definition about beliefs (Zhang & Morselli, in Goldin et al., 2016, pp. 11–
13). According to Ruffell, Mason and Allen (1998), beliefs are “observer constructs
fashioned in order to make sense of the behaviour of others” (p. 2). So the presence
of several definitions can be seen as a consequence of the attempts by researchers
to define the construct depending upon the specific objectives of their research.
Skott (2015) highlights that it is difficult to grasp all aspects of the ‘belief’ concept
in a single, explicit definition. According to him, rather than searching for such a
definition, it can be more useful to recognise the main properties of the concept
itself, according to what has been discussed in the literature:
The core of the beliefs concept may, then, be defined as subjectively true,
value-laden mental constructs that are the relatively stable results of substantial
prior experiences and that have significant impact on practice. (p. 6)
The issue of the impact of beliefs on practice in particular constitutes an important
aspect to take into account when dealing with teacher education and development.
In fact, much research has highlighted that both practising and prospective teachers
often have beliefs related to mathematics and mathematics education that do not
accord with what is supported by research in education (Handal, 2003; Philipp,
2007). Therefore, assuming beliefs have an influence on practice provides teacher
educators with the hope to improve (prospective and practising) teachers’ practice by
prompting a change in their beliefs (Richardson, 1996). However, despite the large
number of studies about change in teachers’ belief and teacher education courses,
and despite the spread of lesson study as a teacher education tool, there still have
been only a few studies focusing on teachers’ beliefs in relation to the participation
in lesson study experiences (Changsri, Inprasitha, Pattanajak, & Changtong, 2012;
Inprasitha & Changsri, 2014). The present study aims to make a contribution to the
research from this perspective.

THEORETICAL PERSPECTIVE

Cultural Beliefs

Abelson (1979) and Nespor (1987) have both argued that beliefs may originate from
culture or personal experience. What seems to be missing in the literature is research
on a particular kind of teacher beliefs, namely those originating from culture. We
might adopt for them the term cultural beliefs, something that was mentioned
by Bruner (1996) in his elaboration of “folk pedagogy,” which he specified as:
“taken-for-granted practices that emerge from embedded cultural beliefs about how

134

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

children learn and how teachers should teach” (p. 46; emphasis added). We observe
that neither Bruner nor other scholars who refer to cultural beliefs defined them as
if they were a shared construct by teachers, researchers, students and, generally,
by the members of each society. For instance, Stigler and Perry (1988) highlighted
differences among American, Chinese and Japanese schools, based on observing
primary school mathematics lessons and focusing on differences in mathematics
knowledge and methods. In their conclusion, the authors addressed the following
question:
Why do we organize our U.S. mathematics classes in the way that we do?
Some part of the answer can be tied to cultural beliefs about the nature of
individual differences and the nature of learning. […] American mothers are
more likely to see mathematical ability as innately determined than are Asian
mothers. Because we tend to think individual children as inherently unique in
their limitations, we believe that the education appropriate for one child may
not be appropriate for another, and thus we tend to emphasize individualized
learning. Asian educators are more comfortable in the belief that all children,
with proper effort, can take advantage of a uniform educational experience,
and so they are able to focus on providing the same high quality experience to
all students. (pp. 51–52; emphasis added)
The aim of this paragraph was to try to suggest a working definition of cultural
beliefs, in the hope that this construct may help us to understand better why the
many Western realisations of lesson study diverged substantially from the Japanese
one. Even if the construct of cultural beliefs is not explicitly addressed in the
current mathematics education literature, we can find many papers engaging with
the difference between the so-called Confucian Heritage Culture (CHC) and the so-
called Western Culture. [We are aware that this clustering is a bit naïve. Although
there are surely many similarities among Chinese, Japanese and Korean cultures
(just to mention a few) that are considered representative of Confucian Heritage
Culture, there are also differences. Accordingly, also referring to ‘Western Culture’
is naïve: there are surely differences between United States and European cultures
and, among European cultures, there are differences between the cultures of different
countries (see Andrews, 2010). But for the purposes of this chapter, concerning
the realisation of lesson study in the West, we accept this simplification.] So, for
instance, Leung, Park, Shimizu, and Xu (2015) devoted a paragraph of the panel on
Mathematics Education in East Asia (in the Proceedings of ICME-13) to Confucian
Heritage Culture, summarising it as follows:
A major characteristic of CHC is the social orientation of its people, in contrast
to individual orientation typically found in Western societies. Social orientation
is a “tendency to act in accordance with external expectations or social
norms, rather than with internal wishes or personal integrity.” It emphasizes
integration and harmony, in contrast to independence and individualism in

135

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

Western cultures. People in CHC treasure the community, much more so than
the individual. Related characteristics of CHC include compliance, obedience,
respect for superiors, and filial piety. (p. 137)
A difference mentioned in the comparison between United States teachers on
the one hand and Japanese and Chinese teachers on the other hand is the belief
about teaching as a private versus public activity (Ferreras & Olson, 2010; Isoda
et al., 2007). Another difference concerns group work. In the East, group work
(at school, but also in life) is implicitly conditioned by hierarchical relationships
among group members, whereas, in the West, group work is often intended as a
setting where every member has the same rights, according to the rules that regulate
democratic discussions. Phuong-Mai, Terlouw and Pilot (2005), for instance, believe
the distance from the tradition of cooperative learning (Johnson & Johnson, 1999)
is not effective in Confucian Heritage Culture contexts, offering instances from
Vietnam and other Asian countries. Also, recourse to student discussion is contrasted
in the Korean tradition with the so-called “pedagogy of silence” (Lee & Sriraman,
2013). In particular, Korean students are not encouraged to express their thoughts
immediately after being given a task, since, the authors claim:
Even today, Koreans hold tight to the belief that it is more virtuous to express
one’s thought politely after having mulled over an idea for some period of
time than to impulsively speak incomplete thoughts. […] Kim (2002) […]
claims talking and thinking are considered to be interdependent in the Western
intellectual tradition, but not in the East. (p. 153)
So, although several studies have raised profound cultural differences between
East and West, there are not many studies that show how these differences in
beliefs of different cultures are reflected in the organisation of the educational
institution and the curriculum, and in the purpose of school education. An important
instance of a study that instead seeks to bring to light the cultural, philosophical
and pedagogical assumptions from which the institutional choices on school and
education originate is that of Xie and Carspecken (2008). They compared Chinese
mathematics curriculum (as representative of Confucian Heritage Culture) and the
National Council of Teachers of Mathematics (NCTM) perspective of mathematics
curricula (as representative of Western Culture). They did not simply highlight the
differences between the two curricula but deepen in various chapters the more or
less explicit assumptions of Chinese and American culture, in order to trace the
profound reasons for these differences in terms of a conception of the world. To
clarify what we are saying, we present the diagram in Figure 5.1 adapted from Xie
and Carspecken (2008), which highlights a series of cultural premises on which the
Chinese and the NCTM Mathematics Standards depend – represented in the boxes at
the bottom of the diagram (for further discussion, see Bartolini Bussi & Martignone,
2013). According to Xie and Carspecken’s viewpoint, no analysis of the curriculum
may be made without referring to the umbrella cultural themes, including world-

136

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

Figure 5.1. The complexity of relationships in mathematics education (adapted from Xie &
Carspecken, 2008, p. 17)

views. We therefore believe that no comparative analysis of lesson study in Japan or


China and in the West should be made without referring to and taking into account
the umbrella themes underlying different cultures.
The discussion about how to define “culture” is complex (e.g., Bishop, 1988,
p. 4). We shall adopt a working definition, given by Bates and Plog (1990), that
seems to encompass what is required for our purposes:
[culture is] the system of shared beliefs, values, customs, behaviours, and
artifacts that the members of society use to cope with their world and with
one another, and that are transmitted from generation to generation through
learning. (p. 7)
Within this frame, we propose to connect the definition of cultural beliefs to
the “system of shared beliefs” mentioned in this definition of culture, adopting the
following working definition:
Cultural beliefs are, on the one hand, beliefs which are socially shared within a
given culture and are considered as characterising this culture by its members,
and, on the other hand, individual beliefs which are reconnectable and

137

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

adherent – within a certain degree of re-elaboration and personalisation – with


beliefs proper to the culture of the society of which the individual is a member.
(Funghi, in preparation, unpublished doctoral thesis)
To determine this definition, we draw on Bruner’s (1996) discussion about
culture and how it influences the human mind. In particular, he describes culture
as something both deriving from a process of social negotiation of meaning and
something that is essential for individual thinking processes, as it gives a set of
meanings that are necessary for individuals to think and communicate.
[Culture] shapes the minds of individuals as well. Its individual expression
inheres in meaning making, assigning meanings to things in different settings
on particular occasions. […] Although meanings are “in the mind,” they have
origins and their significance in the culture in which they are created. It is
cultural situatedness of meanings that assures their negotiability and, ultimately,
their communicability. […] For however much the individual may seem to
operate on his or her own in carrying out the quest for meanings, nobody can
do it unaided by the culture’s symbolic systems. It is culture that provides the
tools for organizing and understanding our world in communicable ways. (p. 3)
It is important to underline that cultural beliefs are usually invisible to the eyes of
the members of that culture because they provide those shared and implicit meanings
that allow them to think, communicate and share a common perspective on the world.
So, among the people coming from the same cultural context, they take these shared
meanings as taken for granted. Our claim is that when we deal with people, methods
or tools coming from another cultural context, we are forced to reflect on those
meanings we usually take as given in order to reach a greater awareness about what
we believe as members of a certain culture. Furthermore, we suddenly realise that
what we usually take for granted is not that, and that other meanings are possible.
In our opinion, even when lesson study is carried out, taking into account the
different cultural background of the country from which it was realised, it still remains
an object which is foreign to the local culture. From this perspective, therefore,
experimenting on lesson study for teachers is an opportunity for reflection upon and
a re-thinking of one’s own cultural beliefs. This is the point we will develop through
the analysis of data discussed in the next section of the chapter, after considering the
construct of cultural transposition.

Cultural Transposition

To cope with the above problems, we propose the idea of cultural transposition
as a process activated by researchers, teacher educators and teachers who begin to
deconstruct those educational practices adopted in other cultural contexts, in order
to reconsider the issues of educational intentionality, which is the background of
any educational practice. As a matter of fact, the role of researchers and educators

138

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

is crucial, since they can introduce original interpretative keys for “the same”
educational practice related to their own cultural context through the deconstruction
of the several levels in which an educational practice is stratified. The different
cultural backgrounds generate different possibilities of meaning and different
mathematics education perspectives that, in turn, organise contexts and school
mathematics practices in different ways (Mellone & Ramploud, 2015).
Our research adopted the perspective of the philosopher and sinologist François
Jullien (2004, 2009), who turned his gaze towards differences and beyond the borders
of Western thought using deconstruction from the outside as a methodological tool.
He traced a path along which anyone could meet and successfully put one’s own
‘un-thought’:
[…] Il ne s’agit pas là de philosophie comparée, par mise en parallèle des
conceptions; mais d’un dialogue philosophique, où chaque pensée, à la
rencontre de l’autre, s’interroge sur son impensé. (Jullien, 2006, p. 8)
[This is not about comparative philosophy, about paralleling different
conceptions, but about a philosophical dialogue in which every thought,
when coming towards the other, questions itself about its own unthought. (our
translation)]
Thus, this meeting is between different philosophical perspectives and thoughts,
but above all, it deals with the difference or the gap between cultures. The contact
with distant and hard-to-conceive didactic practices represents what the French
philosopher Jullien defines as impensé (Jullien, 2006). With this term (‘un-thought’
in English), Jullien refers to all the implicit assumptions in which a cultural paradigm
is rooted. People remain unaware of these implicit assumptions while they reside
in the same cultural paradigm, which becomes even unthinkable; however, while
moving towards a different cultural paradigm, they can become aware of them.
As previously stated, we proposed the tool of cultural transposition with the
aim of using the differences among mathematical education practices adopted in
different cultures and societies to design professional development, which aims
to develop teachers’ awareness and, eventually, to change their lesson design and
implementation (Mellone et al., 2018). Hence, we claim that it is not possible to
transport a method such as lesson study from one cultural context to another, without
taking into account the difference between the umbrella themes that underlie the
different cultures. An operation of cultural transposition can therefore be carried out
by a researcher or a teacher educator only if she or he is aware of the dependence of
the transposed object upon the beliefs proper to the culture from which it originates.
This statement is consistent with Wang and Lin (2005), who claimed:
[an] approach from one country to another may not be useful in producing
similar students’ performances without careful consideration of the cultural
tradition and foundation upon which the practice or approach was conceived,
developed and implemented. (p. 4)

139

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

We shall exploit the analysis of teachers’ cultural beliefs to know more about this
process.

An Example: Twelve Minutes More

In this subsection, we discuss an example of Japanese lesson study that allows the
seeing of cultural constraints at work. A classical book on Japanese lesson study,
one of the first to be published in English (Fernandez & Yoshida, 2004), offers
much interesting information about the realisation of just one lesson study cycle in a
Japanese elementary school. All the processes of preparation, actual class and lesson
review were carefully reported by the two observers. They focused on the features
of the lesson plan, that included a careful description of the class where the lesson
was realised and of the manipulatives to be used, as well as a very detailed analysis
of the ways of posing the problem and of responding to students’ reactions. Their
report answers a lot of Westerners’ questions about which features are indispensable
for the realisation of a lesson study in Japan.
For instance, the mismatch between the planned duration (forty-five minutes) and
the actual total length of the observed lesson (twelve minutes more) was carefully
analysed. Several pages (pp. 111–114) were devoted to the issue of “improving the
use of time in the lesson,” trying to find ways of cutting some parts that, according
to the observers, have, on the one hand, “broken the planned harmony between form
and order, and, on the other hand, made the students confused and slowed down in
the process.”
The discussion of “twelve minutes longer” might hardly find the same attention
and conclusion in a different cultural context. For instance, in Bartolini Bussi et al.
(2017), an even larger time mismatch appeared in a lesson study cycle: seventeen
minutes more than the sixty originally planned were observed in a lesson on fractions.
The lesson review aimed not to determine what to cut in order to match the form of
the lesson plan (i.e., the planned time), but rather to find whether the time given to
each step in the real lesson was suitable for the intended processes, that included the
intervention of all the small groups to explain their solution of the given problem
and the comments of the pupils. This need fits a teacher belief that a good model of
teaching is to involve individual pupils deeply and to have time enough to compare
solutions, while also on commenting similarities and differences.
How can we interpret this attention to the “harmony between order and form”
that shaped the lesson review in Japan? We have tried to reconstruct the “umbrella
cultural themes” with the help of a yamatologist, who introduced us to the issue of
kata or shikata. She mentioned the tradition of the practice of martial arts:
Form is a significant word within this practice: form refers to a coded sequence
of movements through which we exercise body and mind at the same time: the
body’s movement is marked by a sequence of positions that are representative

140

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

of a given reality, as well as being both a mental and an emotional situation.


[…] Practice and teaching within Martial Arts are built on kata (form): the
coded sequence of precise movements, evoking real situations within which
the practice learns to harmoniously exercise the whole body: movement,
breathing, attention. […] In Martial Arts, practice reserves special attention to
the transmission of principles through techniques. (Casadei, 2014, p. 19)
This excerpt highlights the importance of imitation (e.g., imitation of animals) and
of emulation (of the instructor) who shows the movements of his/her body. Kata (or
shikata) is indeed a much more general word, one that applies to the whole life of
Japanese. According to De Mente (1990):
[the] shikata concept includes more than just the mechanical process of doing
something. It also incorporates the physical and spiritual laws of the cosmos. It
refers to the way things are supposed to be done, both the form and the order,
as a means of expressing and maintaining harmony in society and the universe.
(pp. 13–14)
Shikata may well describe the strict structure of the lesson plan, with a strong wish
to design the lesson perfectly, and may explain the reaction in the post-lesson review to
any deviation of the actual lesson from the designed lesson. We might express this by
saying that the protagonists of the lesson to be observed are neither the students nor the
teacher, but the form and the order. In a booklet published on the occasion of ICME–12
(ESUT, 2012), there is a detailed annotated list of terms to be used in order to guide the
classroom processes: they may be interpreted as elements of lesson study kata.
The hidden (and mostly implicit) kata may also explain the disappointment of
Japanese scholars when they look at lesson study as developed in other countries.
For instance, in the introduction to a chapter on task design in lesson study, Fujii
(2015) wrote:
It is becoming clear that there are aspects of Lesson Study that are implicitly
understood by Japanese teachers that have not transferred easily to other
countries. For Lesson Study to be successful, these aspects should be made
explicit. (p. 273)
This analysis is consistent with Lewis (2011):
The term “lesson study” connotes, to many non-Japanese, a focus on the
lesson plan. While the lesson plan is certainly central to lesson study, the word
“lesson” always refers to live lessons, not to something that can be captured
on paper (as it sometimes does in English, when we hand someone a piece of
paper and say “take a look at my lesson”). (p. 239)
Considering this perspective, our research group started a research programme in
mathematics teacher education and development through lesson study.

141

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

A RESARCH PROGRAMME ON LESSON STUDY

Since 2014, our research group has undertaken dozens of lesson study experiments
with Italian teachers working in the region of Modena and Reggio Emilia. Most
teachers had already worked with the university research group in the past or were
school tutors of the prospective teachers during their internship in schools. Hence,
they were already familiar with the general idea that “the simple transplantation of
a particular kind of teaching practice […] from one country to another may not be
useful” (Wang & Lin, 2005, p. 4). In other words, they were familiar with the idea
of cultural transposition.

The Programme of Teacher Development

This research programme, started by our group in 2014, aims to study how lesson
study can be integrated into the Italian school system, especially in primary schools.
Over the last two years (2016–2017), our main goal has been to spread the practice
of lesson study among local teachers as much as possible, in order to encourage
more and more teachers to participate in lesson study and to comprise autonomous
lesson study groups within schools, with the support of one or two experts (called, in
Confucian Heritage Culture, “knowledgeable others”). In our case, these experts were
people belonging to the research group in mathematics education at the University of
Modena and Reggio Emilia. The structure of the lesson plan we proposed to teachers
can be found in Bartolini Bussi et al. (2017). In short, the differences between the
Chinese lesson plan and the Italian lesson plan can be seen by comparing sections of
the two instruments, as shown in Table 5.1, where the phases of the Chinese lesson
plan are mandatory.
In fact, the two structures are very similar, especially in the final part of the plan.
Our changes concern mainly the following parts, that were partly inspired by the
Japanese lesson study (Fernandez & Yoshida, 2004), as they met the needs of Italian
teachers:
‡ There is a section to be filled in with data about the class where the lesson has
to be taught. This section was introduced to let the teachers circumscribe their
fieldwork from a disciplinary viewpoint and to let them take into account the
needs of the specific class during the planning and also for those teachers that do
not know the class.
‡ There is also a section dedicated to the careful design and analysis of materials to be
used during the lesson. This section was inserted to push teachers to reflect on the
potential of materials they could use; for expert teachers, this section could be also
dedicated to the analysis of the manipulatives, according to the Theory of Semiotic
Mediation (Bartolini Bussi & Mariotti, 2008; see also Bartolini Bussi et al., 2017).
Moreover, we decided to insert an additional column into the lesson plan, where
teachers must describe the reasons for their didactical choices made in the planning

142

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

Table 5.1. Comparison between sections of the Chinese lesson plan and
the Italian lesson plan

Chinese lesson study: Lesson plan Italian lesson study: Lesson plan

Not found in the Chinese literature Contextualisation [this section is not a phase
of the lesson; it is a section necessary for the
teachers’ planning meeting]: description of
class composition; contextualisation of the
lesson to be taught into the educational path;
core topic of the lesson; objectives of the
lesson; observation foci.
Not found in the Chinese literature Analysis of materials to be used [this section
is not a phase of the lesson; it is a section
necessary for the teachers’ planning meeting]
Summary and/or revision of previous
lesson
Homework check Homework check (non-mandatory phase)
Presentation of the core topic Introduction of the lesson and presentation of
the core topic
Explanation of the problem of the day Explanation and clarification of the problem
of the day
Work on sub-problems Work on sub-problems (non-mandatory
phase)
Work on the main problem Work on the main problem
Students’ presentation of their work Students’ presentation of their work
Discussion about different solving Discussion about different solving strategies
strategies
Exercises Exercises (non-mandatory phase)
Teacher’s summary and emphasis of the Teacher’s summary and emphasis of the main
main results of the lesson results of the lesson
Homework assignment Homework assignment (non-mandatory
phase)
Anticipation of the topic of the next lesson Anticipation of the topic of the next lesson
(non-mandatory phase)

of each phase. This column was introduced because, in Italy, the teacher is in charge
of choosing the teaching methodology, materials, and classroom organisation for
each of his/her lessons. Therefore, our main concern (as teacher educators) was to
make teachers aware of their didactical choices, in their planning of the lesson.
Italian lesson study cycles were planned for lessons lasting sixty minutes, and
not forty-five minutes, as it is in Eastern lesson study. At the primary level, it is not

143

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

unusual for Italian teachers to plan lessons lasting even more than an hour, since
teachers usually teach more than one subject, and so they often spend more than an
hour with the same class.

METHODOLOGY

In order to study teachers’ beliefs, we decided to collect qualitative data (Kaasila,


2007) by means of a questionnaire we created with mostly open-ended questions.
The reason for choosing this research tool was that, unlike interviews, the use of the
questionnaire allowed us to gather qualitative data from relatively many respondents.
Moreover, even though questionnaires cannot gain the deepness of information a
researcher can obtain with interviews (Di Martino & Funghi, 2016; Kaasila, 2007),
open-ended questionnaires are – among the various kinds of questionnaires – ones
that allow us to catch “authenticity, richness, depth of response, honesty and candor
which […] are the hallmarks of qualitative data” (Cohen, Manion, & Morrison,
2002, p. 255). In fact, they give respondents the opportunity to express their thoughts
with the words they prefer and to focus on the aspects that they consider important
(Di Martino & Sabena, 2011).
In its final version, the questionnaire comprised two closed questions concerning
the respondent’s profession and degree of knowledge about lesson study before the
experimentation, one open question asking respondents to list as many words as
possible related to “Lesson Study,” and the following further open questions:
‡ How did you feel about the key topic of the lesson to be designed? Did you
find it difficult to follow the interventions made by your collaborators during the
planning?
‡ How did you get on dealing with people from other professional backgrounds and/
or training paths (e.g., university tutors, teachers working in other school levels)?
In particular, what benefits and critical issues did you find in this confrontation?
‡ What do you think of the lesson plan as a working tool? How did you get on with
it? Are there any changes that you would make to it, and if so, which ones?
‡ Making an overall balance of the planning meeting, do you think this working
method was functional and effective for the realisation of the lesson? Why?
‡ Can you briefly describe what your experience was within the Lesson Study you
participated in? In particular, what was the most difficult moment for you and the
one where you were most at ease?
In this chapter, we focus on the analysis of the answers to these last open questions,
since they returned data suitable to reveal teachers’ cultural beliefs.
During the 2016–2017 academic year, our staff prompted around thirty Italian
lesson study cycles. At the end, we sent the questionnaire to every participant (some
one hundred teachers), but we received response data from just thirty-nine.
Concerning the analysis of data, Lieblich, Tuval-Mashiach, and Zilber (1998)
identify four possible approaches to analyse narrative data, depending on the two

144

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

dichotomies form versus content, and holistic approach vs categorical approach.


The first dichotomy regards the focus of the analysis, which can be on formal aspects
of the narrative (e.g., structure of the plot, figures of speech) or on its content (e.g.,
events, characters). The second dichotomy concerns the unit of analysis, which can
be the whole narrative text (holistic approach) or single sentences extracted from
the context (categorical approach): in particular, in the second case, “the original
story is dissected, and sections or single words belonging to a defined category
are collected from the entire story or from several texts belonging to a number of
narrators” (p. 12). We choose the content-categorical approach to analyse our data,
since this approach is considered to be particularly suitable to study a phenomenon
common to a group of people (Kaasila, 2007). In particular, in our data, we looked
for aspects of lesson study considered suited to or desirable for the Italian context,
aspects that on the contrary were evaluated as not fitting into the Italian context, and
the reasons respondents gave to justify their evaluations. For example, consider the
following excerpt:
At the beginning, it [i.e. lesson study] seemed a bit ‘mechanical’ because to
each activity within the lesson is given a certain amount of time, and I think
that in an Italian class giving a certain amount of time is not always easy,
it depends on the class, on the situation and on the type of activity that is
proposed, so this thing made me doubtful. Actually, I realised that planning in
this way the management of the time is better, that is, there is no dead time, and
therefore time is used to the maximum […]
We can see that, in the first sentence, the respondent identifies the strict planning
of the time for each lesson phase as a critical point of the implementation of lesson
study in the Italian context and the reason for this is the high number of variables the
respondent feels she or he has to cope with. Instead, in the second sentence, we can
see that the respondent changes his or her mind and describes the strict planning of
time in the lesson as a benefit of lesson study, because she or he realises that this way
of planning allows to exploit all the time available.

SOME RESULTS FOR THE PRACTISING TEACHERS

We summarise here the main results of the analysis of data. In particular, we focus on
the aspects that emerged from the answers that best highlight the teachers’ awareness
of beliefs proper to Italian culture and, eventually, their rethinking of them, according
to the perspective we illustrated earlier in the chapter.
The general judgement about lesson study was positive. Yet some teachers were
doubtful not only about the functionality of a single lesson plan for the learning
of all children, but also about the functionality of a single plan for more than one
teacher. Around a quarter of the participants (ten out of thirty-nine) were critical
about the possibility of designing a very precise division in slots of time, stating the
impossibility of anticipating the time needed to pay attention to every single child,

145

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

depending on their individual needs, and to unexpected events. These considerations


are in line with the features of Italian culture, since in Italy teachers are free to
choose the teaching methods they prefer and organise the contents of the curricula
as s/he wishes. Moreover, in Italy, classes are generally inclusive, that is, each class
can include children with physical disabilities, learning disabilities, foreign children
who do not speak Italian as a mother tongue, and so on, so teachers frequently feel
the need to have time to cope with different needs of the children.
Most teachers (twenty-six out of thirty-nine) expressed a positive judgment about
the lesson plan as a design instrument. Some teachers appreciated it because it
helped to focus and keep in mind the important points of the lesson. Others instead
underlined its usefulness to enhance their awareness of their own choices and to
make systematic revisions during the post-lesson meeting or stressed the possibility
to collect data about several aspects of the lesson. In some cases, it was emphasised
that the lesson plan induced teachers to reflect on their management of time in
everyday lessons. These claims are important because, in Italy, teachers tend to be
very flexible during their lessons and so they do not prepare a strict programme for
them. In this case, we can see that experimenting with lesson study allowed them
to rethink their habits about lesson planning and management of time, and to gain a
deeper awareness of their educational choices in their way of teaching.
All teachers considered the collaboration with other members of the group very
positive. Many of them underlined especially that other members had been open
to listen to everyone, in spite of different levels of competence and expertise in
mathematics and/or in education. In some cases, the design meeting lasted more
than two hours, in order to give all of the participants the possibility to express their
ideas about the lesson and sometimes also to discuss about individual pedagogical
and educational opinions. The level of colleagues’ expertise at the very beginning,
however, caused a feeling of anxiety in some participants.
Nevertheless, the collective nature of lesson study was appreciated for many
reasons. A fair number of the teachers in our sample (fifteen out of thirty-nine)
underlined that the discussion has been interesting precisely because it had provided
an authentic opportunity to exchange ideas, allowing them to clarify their thoughts
about the core topic of the lesson. Others stated that collective discussion permitted
making the lesson plan more refined and effective or that it brought aspects of
teaching to the fore that usually are implicit or that otherwise could have been
neglected or ignored. The involvement of university researchers during the planning
and revision of the lesson in seven cases was described as particularly meaningful
to create links between theory and practice and to stimulate teachers to explore
innovative teaching methods.
Some teachers seemed to discover for the first time that colleagues can be a
support to overcome difficulties, and described their previous sense of loneliness,
the lack of real feedback from colleagues and scarcity of possibilities to discuss,
all issues that were partially resolved by the experience of the lesson study. These
considerations are important, because teachers in Italy tend to work alone for most

146

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

of the time; they are not used either to being observed during their lessons or to
collaborating closely with their colleagues. The experience of lesson study allowed
some of them to rethink about the possibility of working with others in a fruitful and
non-evaluative manner.

THE EXTENSION TO PROSPECTIVE TEACHER EDUCATION

During the last three years, lesson study has also been introduced into prospective
primary school teachers’ education at the University of Modena and Reggio Emilia
(a five-year university programme). In particular, lesson study as a teaching
approach is presented and explained to prospective teachers during the course on
general education in the second year and the course on mathematics education in the
third year; for the last course, prospective teachers attend a twelve-hour mandatory
workshop dedicated to a simulation of a lesson study cycle. They just ‘simulate’ the
planning phase in order to design a lesson about an assigned topic by means of the
Italian lesson plan for a prescribed class. Eventually, in the fourth and fifth year of
their university education, the prospective teachers may also choose a project on
lesson study as part of their internship in primary school, during which they take part
in real lesson study cycles.
The paragraphs below summarise part of the findings regarding the prospective
teachers’ beliefs about these different experiences. They concern two different sets
of data: one is related to the responses to an open-ended questionnaire distributed to
seventy prospective teachers at the end of the workshop of the third year (sample A),
while the other concerns the interviews of five prospective teachers who took part
in an actual lesson study project within their internship at the fifth year of education
(sample B).
Nearly half of the prospective teachers in sample A (thirty-two out of seventy)
initially thought that lesson study was unrealistic to be realised in Italian schools,
mainly because teachers need to dedicate much time to it. These prospective
teachers complained about the absence of experiencing a real lesson study in a class:
some of them affirmed that the simulation did not give information about children’s
responses and difficulties. Also, the three prospective teachers in sample B (who
had participated in a workshop similar to the one described above) confirmed that
the potential of lesson study can be seen only when one takes part in a real lesson.
However, all the interviewees in sample B retained the lesson study experience as an
extremely useful opportunity for prospective teachers to be prepared for their future
profession, because it gave them the possibility to receive specific feedback from
expert teachers and to understand what they need to improve in their way of teaching
– whereas many of the prospective teachers in sample A were doubtful about the
usefulness of the method in relation to the Italian school system.
All prospective teachers in sample B appreciated the lesson plan as a design
instrument, even if they confirmed that the management of time remained a crucial
issue in the realisation of lesson study cycles. Nevertheless, they recognised that

147

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

the strict timing of the lesson plan can be useful to face some of those issues met by
practising teachers. A prospective teacher underlined that the strict planning of time
made children faster and more attentive to the task; another prospective teacher stated
that lesson study and lesson plan actually helped the teacher to avoid downtime,
and that lesson study prevented the teacher from getting lost in the development
of the lesson, since it allowed the maintenance of a train of thought throughout the
entire lesson. Moreover, this interviewee explained that she also found lesson study
useful because it helped to exploit children’s responses to achieve the teacher’s goals
without leaving the development of the lesson completely dependent upon those
responses. Three of the prospective teachers in sample B even chose the lesson study
as a major focus for their master’s dissertation involving a teaching experiment in
primary school.
Similar to the practising teachers, nearly half of the prospective teachers in sample
A (thirty out of seventy) appreciated the collective management of the process
permitting an overcoming of an initial feeling of disorientation. In some cases, the
prospective teachers confirmed that, especially at the beginning, being observed was
something scary for them, especially if the observers were expert and more than
one. All prospective teachers (both from sample A and B) appreciated the collective
nature of the lesson study. In particular, the relationships that the lesson study created
among teachers was highlighted by and contrasted with perceived teachers’ division
and isolation. Nevertheless, this does not mean harmony or agreement: for instance,
one of the prospective teachers in sample B emphasised the difficulty of having a
relaxed discussion among different members of her group. Moreover, discussing
proved to be problematic for her because she did not feel legitimate talking because
of her lack of experience.

DISCUSSION: COMPARISON ACROSS THE THREE SAMPLES

At the end of their experience, the three samples (practising teachers; prospective
teachers in sample A, and prospective teachers in sample B) showed common beliefs,
with some exceptions. In the following, we highlight those ones that seem to be most
related to features of Italian culture.
The most positive comments shared across all three samples concerned the
collective feature of the process, one that was able to contrast the difficulties in
communication among teachers, and therefore the isolation, that was reported also
by Stigler and Hiebert (2009, p. 123) for United States teachers. This comment
emphasises a common critical belief regarding the Western ‘individualistic’ position
of teachers.
The three samples also reflected some common beliefs about observation. This
was sometimes considered a means of evaluation – as it happens, for example,
for Italian teachers in their first year of teaching – and this is probably why some
practising teachers at the beginning seemed to misunderstand the observers’ role.
The fear of being evaluated can also be related to the fact that, in Western culture,

148

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

criticising someone can be considered offensive to the person’s self-esteem,


especially when it happens in front of others. For example, Stigler and Perry (1988)
claim: “American teachers generally do not call children to the board to display their
errors, because they fear the possible damage it might do to the child’s self-esteem”
(p. 29); whereas Li (2012), in relation to the Eastern viewpoint, claims: “when the
self is regarded as malleable and improvable through learning, children know that a
mistake does not indicate an inherent flaw in themselves, but rather an opportunity
to self-improve” (p. 318).
A difference between the three samples appeared to be concerning the management
of time. The strict time limits designed in the lesson plan were criticised by some
practising teachers and by some prospective teachers in sample A. The structure
itself was perceived as limiting the creative process that teachers and prospective
teachers were expected to start in the planning phase. They were not accustomed
to planning lessons at this level of detail. Yet, in some cases, the very structure of
the lesson plan was perceived as a way to exploit as much as possible the time and
the resources at a teacher’s disposal, not as a means of forcing the teacher to hurry
within the lesson or to cut parts of the lesson that were considered relevant. In other
cases, it was perceived as a limitation to individual design.
In this respect, our respondents seemed to be similar to the sample of United
States teachers in the study by Cai (2006). Comparing lesson plans written by nine
Chinese teachers and eleven United States teachers concerning an introduction to
the concept of ratio, he found that United States teachers gave a general outline of
the lesson, whereas Chinese teachers wrote precise plans, stating their questions as if
they were the lines of a screenplay, and predicting at every phase children’s answers.
In fact, it is not hard to infer that if Chinese teachers plan an everyday lesson at
this level of detail, it would not be a problem for them to complete the lesson plan
starting from one of these scripts; moreover, it is also possible that Chinese teachers’
expertise with lesson study influences the way they think about everyday lesson
plans.
It must be noticed that some Chinese teachers are well aware of the need to
combine strict plans with unexpected events, and therefore claim that real
coherence is also related to a teacher’s ability to adapt the lesson to contingencies
(Cai & Wang, 2010). Prospective teachers in sample B seemed to be less doubtful
about the possibility to combine flexibility and strictness of a lesson plan when
compared with practising teachers, who stressed their wish to stay flexible to meet
students’ individual needs. In this regard, practising teachers expressed views more
in line with those highlighted by Bryan, Wang, Perry, Wong, and Cai (2007) for
United States and Australian teachers, but, in our context, this need seemed both
to be challenged and the teachers exasperated by the extreme heterogeneity in the
composition of the class and the presence in each classroom of many students with
special needs.
The difference between sample A and sample B prospective teachers may depend
on the lack of classroom practice of those in sample A, a fact which induced these

149

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

prospective teachers to consider lesson study as an abstract exercise of teacher


education, whereas prospective teachers in sample B stated that experiencing lesson
study during their education allowed them to look critically at routines that seemed
to be well-established among Italian teachers. In the interviews, some other aspects
appeared which are likely to be related to specific features of the Italian context – in
this sense, different from other Western contexts. This issue is not considered here,
for space limitation reasons (see Bartolini Bussi et al., 2017).
For all the reasons expressed above, even if we proposed a lesson plan with
adaptations to the Italian context to teachers in our samples, they had to face a
work method that was very unusual for them, since it drew on different pedagogical
principles. To say this in terms of cultural transposition, the teachers’ encounter with
lesson study was an opportunity for them to reflect on some ‘un-thought issues’ and
also on some reasons why teaching in Italy and in the far East can be so different.
This aspect is particularly evident in the protocol discussed in the first section of
this chapter, which hinted at the collective and collaborative process, at mutual
observation, at the consistency between the lesson plan and Italian educational aims,
and at the need for challenging one’s own beliefs.
Another reflection may be made. Murata (2011) claims that lesson study may
provide the opportunity to understand “the existing educational system and cultural
values and beliefs that support the system” (p. 10). The prospective and practising
teachers in our samples were representatives of the Italian educational system and,
on the one hand, were put in the position of understanding better the cultural values
and beliefs that support it and, on the other, of making the researchers aware of them.

CONCLUSION AND IMPLICATIONS

Our programme has shown that lesson study could serve as a means to improve
the professional education and development of teachers in the Italian context. The
total sample comprises practising and prospective teachers who are supposed to be
in the process of distancing themselves from their beliefs and reconsidering their
un-thoughts. It is not surprising that the process seems to be more advanced for
practising teachers and for prospective teachers in sample B who actually had the
experience of practising lesson study in the school, whereas it seemed still to be in
progress for prospective teachers in sample A, who had just a workshop on lesson
study without the possibility of testing the lesson plan in school. The practical
impossibility of offering the experience to the prospective teachers in sample A
during their internship in school has surely altered the sense of lesson study, that
in the Confucian Heritage Culture version is intentionally meant as a unit of doing
and knowing. This result is now being analysed by our research group, in order to
introduce some changes to the workshop.
We agree with Takahashi and McDougal (2018) that the term ‘Lesson Study’ is
likely to be substituted by Collaborative Lesson Research in order to distinguish

150

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

lesson study from its diverse and less-than-faithful overseas adaptations. This
clarification might help out “building a general theory of [lesson study], if any, in
a critical way” (Clivaz & Takahashi, 2018, p. 156). If the analysis that we have
outlined in this chapter is correct, in every attempt to construct a ‘theory’ of lesson
study around the world, we should consider that, in both Japanese and Chinese
lesson study, the experience is a life-long process, spanning thousands of hours,
whilst the lesson study programmes carried out in the West are still limited, both in
size and in time.

ACKNOWLEDGEMENTS

We thank all the prospective and practising teachers who took part in this study.
The chapter was prepared by the three authors collaboratively. The sections from
Methodology to Conclusions draw on the data collected by S. Funghi for her Ph.D.
thesis. We thank the three reviewers for their careful reading and commenting of the
first draft of the chapter, and David Pimm for his invaluable help in editing the final
version.

REFERENCES
Abelson, R. P. (1979). Differences between belief and knowledge systems. Cognitive Science, 3(4),
355–366.
Andrews, P. (2010). The importance of acknowledging the cultural dimension in mathematics teaching
and learning research. Acta Didactica Napocensia, 3(2), 3–16.
Baba, T. (2007). How is lesson study implemented? In M. Isoda, M. Stephens, Y. Ohara, & T. Miyakawa
(Eds.), Japanese lesson study in mathematics: Its impact, diversity and potential for educational
improvement (pp. 2–7). Singapore: World Scientific.
Baba, T., Ueda, A., Ninomiya, H., & Hino, K. (2018). Mathematics education lesson study in Japan
from historical, community, institutional and development assistance perspectives. In M. Quaresma,
C. Winsløw, S. Clivaz, J. P. Da Ponte, A. Ní Shúilleabháin, & A. Takahashi (Eds.), Mathematics lesson
study around the world (pp. 23–45). Cham: Springer.
Bartolini Bussi, M. G., Bertolini, C., Ramploud, A., & Sun, X. (2017). Cultural transposition of Chinese
lesson study to Italy. An exploratory study on fractions in a fourth-grade classroom. International
Journal for Lesson and Learning Studies, 6(4), 380–395.
Bartolini Bussi, M. G., & Mariotti, M. A. (2008). Semiotic mediation in the mathematics classroom:
Artifacts and signs after a Vygotskian perspective. In L. D. English & D. Kirshner (Eds.), Handbook
of international research in mathematics education (pp. 746–783). New York, NY: Routledge.
Bartolini Bussi, M. G., & Martignone, F. (2013). Cultural issues in the communication of research on
mathematics education. For the Learning of Mathematics, 33(1), 2–8.
Bates, D. G., & Plog, F. (1990). Cultural anthropology. New York, NY: McGraw-Hill.
Bishop, A. J. (1988). Mathematical enculturation: A cultural perspective on mathematics education.
Dordrecht: Kluwer Academic.
Bruner, J. (1996). The culture of education. Cambridge, MA: Harvard University Press.
Bryan, C. A., Wang, T., Perry, B., Wong, N., & Cai, J. (2007). Comparison and contrast: Similarities
and differences of teachers’ views of effective mathematics teaching and learning from four regions.
ZDM-Mathematics Education, 39(4), 329–340.
Cai, J. (2006). U.S. and Chinese teachers’ conceptions and constructions of representations: A case of
teaching ratio concept. International Journal of Science and Mathematics Education, 4(1), 145–186.

151

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

Cai, J., & Wang, T. (2010). Conceptions of effective mathematics teaching within a cultural context:
Perspectives of teachers from China and the United States. Journal of Mathematics Teacher Education,
13(3), 265–287.
Casadei, R. (2014). Education throught movement: Self-learning and the ability to feel, know and
change oneself. The contribution of Far Eastern traditional performing arts in Western pedagogic
thought. In L. Cavana & R. Casadei (Eds.), Proceedings of the seminar WAZA GENGO: Mente,
corpo, conoscenza e relazione educativa – Scambi fra tradizioni d’Oriente ed esperienze innovative
d’Occidente (in English, Italian and Japanese) (pp. 13–22). Bologna, Italy: Scuola di Psicologia e
Scienze della Formazione, Alma Mater Studiorum Università di Bologna.
Changsri, N., Inprasitha, M., Pattanajak, A., & Changtong, K. (2012). A study of teachers’ perceived
beliefs regarding teaching practice. Psychology, 3(4), 346–351.
Clivaz, S., & Takahashi, A. (2018). Mathematics lesson study around the world: Conclusions and looking
ahead. In M. Quaresma, C. Winsløw, S. Clivaz, J. P. Da Ponte, A. Ní Shúilleabháin, & A. Takahashi
(Eds.), Mathematics lesson study around the world (pp. 153–164). Cham: Springer.
Cohen, L., Manion, L., & Morrison, K. (2002). Research methods in education. London: Routledge
Falmer.
De Mente, B. L. (1990). Japan’s secret weapon: The cultural programming that made the Japanese a
superior people (1st ed.). Phoenix, AZ: Phoenix Books.
Di Martino, P., & Funghi, S. (2016). “Think about your maths teachers”: A narrative bridge between future
primary teachers’ identity and their school experience. In C. Csíkos, A. Rausch, & J. Szitányi (Eds.),
Proceedings of the 40th Conference of the International Group for the Psychology of Mathematics
Education (Vol. 2, pp. 211–218). Szeged: PME.
Di Martino, P., & Sabena, C. (2011). Elementary pre-service teachers’ emotions: Shadows from the past
to the future. In K. Kislenko (Ed.), Proceedings of MAVI 16 Conference: Current state of research on
mathematical beliefs XVI (pp. 89–105). Tallin: Tallinn University of Applied Sciences.
Elementary School University of Tsukuba [ESUT]. (2012). Lesson study of Japanese style in elementary
school mathematics, issue printed for the 12th International Congress of Mathematics Education.
Seoul: Fujiwara Printing.
Fernandez, C., & Yoshida, M. (2004). Lesson study: A Japanese approach to improving mathematics
teaching and learning. Mahwah, NJ: Lawrence Erlbaum Associates.
Ferreras, A., & Olson, S. (2010). The teacher development continuum in the United States and China:
Summary of a workshop. Washington, DC: The National Academies Press.
Fujii, T. (2015). The critical role of task design in lesson study. In A. Watson & O. Minoru (Eds.), Task
design in mathematics education, an ICMI study 22 (pp. 273–286). Cham: Springer.
Goldin, G. A., Hannula, M. S., Heyd-Metzuyanim, E., Jansen, A., Kaasila, R., Lutovac, S., … Zhang, Q.
(2016). Attitudes, beliefs, motivation and identity in mathematics education. An overview of the field
and future directions – ICME13 Topical Surveys. Cham: Springer.
Handal, B. (2003). Teachers’ mathematical beliefs: A review. The Mathematics Educator, 13(2), 47–57.
Hart, L. C., Alston, A. S., & Murata, A. (Eds.). (2011). Lesson study research and practice in mathematics
education. Dordrecht: Springer.
Inoue, N. (2011). Zen and the art of neriage: Facilitating consensus building in mathematics inquiry
lessons through lesson study. Journal of Mathematics Teacher Education, 14(1), 5–23.
Inprasitha, M., & Changsri, N. (2014). Teachers’ beliefs about teaching practices in the context of lesson
study and ppen Approach. Procedia – Social and Behavioral Sciences, 116, 4637–4642.
Inprasitha, M., Masami, I., & Ban-Har, Y. (Eds.). (2015). Lesson study: Challenges in mathematics
education. Singapore: World Scientific.
Isoda, M., Stephens, M., Ohara, Y., & Miyakawa, T. (Eds.). (2007). Japanese lesson study in mathematics:
Its impact, diversity and potential for educational improvement. Singapore: World Scientific
Publishing Company.
Johnson, D. W., & Johnson, R. (1999). Learning together and alone: Cooperative, competitive, and
individualistic learning (5th ed.). Boston, MA: Allyn & Bacon.
Jullien, F. (2004). In praise of blandness proceeding from Chinese thought and aesthetics. New York,
NY: Zone Books.

152

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS’ CULTURAL BELIEFS

Jullien, F. (2006). Si parler va sans dire. Du logos et d’autres ressources. Paris, France: Seuil.
Jullien, F. (2009). The great image has no form, or on the nonobject through painting. Chicago, IL: The
University of Chicago Press.
Kaasila, R. (2007). Using narrative inquiry for investigating the becoming of a mathematics teacher.
ZDM-Mathematics Education, 39(3), 205–213.
Lee, K., & Sriraman, B. (2013). An Eastern learning paradox: Paradoxes in two Korean mathematics
teachers’ pedagogy of silence in the classroom. Interchange, 43(2), 147–166.
Leung, F. K., Park, K., Shimizu, Y., & Xu, B. (2015). Mathematics education in East Asia. In S. J. Cho
(Ed.), The Proceedings of the 12th International Congress on Mathematical Education (pp. 123–143).
Cham: Springer.
Lewis, C. (2011). Response to part IV: Seeing the whole iceberg – The critical role of tasks, inquiry
stance, and teacher learning in lesson study. In L. C. Hart, A. S. Alston, & A. Murata (Eds.), Lesson
study research and practice in mathematics education (pp. 235–240). Dordrecht: Springer.
Lewis, C. C. (2002). Lesson study: A handbook for teacher-led improvement of instruction. Philadelphia,
PA: Research for Better Schools Inc.
Lewis, C. C. (2004). How does lesson study improve mathematics instruction? ZDM-Mathematics
Education, 48(4), 571–580.
Lewis, C. C., Perry, R. R., & Hurd, J. (2009). Improving mathematics instruction through lesson study:
A theoretical model and North American case. Journal of Mathematics Teacher Education, 12(4),
285–304.
Li, J. (2012). Cultural foundations of learning: East and West. New York, NY: Cambridge University
Press.
Lieblich, A., Tuval-Mashiach, R., & Zilber, T. (1998). Narrative research: Reading, analysis, and
interpretation. Thousand Oaks, CA: Sage.
Mellone, M., & Ramploud, A. (2015). Additive structure: An educational experience of cultural
transposition. In X. H. Sun, B. Kaur, & J. Novotna (Eds.), Conference proceedings of the ICMI study
23: Primary mathematics study on whole numbers (pp. 567–574). Macao: University of Macau.
Mellone, M., Ramploud, A., Di Paola, B., & Martignone, F. (2018). Cultural transposition: Italian
didactic experiences inspired by Chinese and Russian perspectives on whole number aritmetic. ZDM-
Mathematics Education, 51(1), 199–212.
Murata, A. (2011). Introduction: Conceptual overview of lesson study. In L. C. Hart, A. S. Alston, & A.
Murata (Eds.), Lesson study research and practice in mathematics education (pp. 1–12). Dortdretch:
Springer.
Nespor, J. (1987). The role of beliefs in the practice of teaching. Journal of Curriculum Studies, 19(4),
317–328.
Philipp, R. A. (2007). Mathematics teachers’ beliefs and affect. In F. K. Lester (Ed.), Second handbook of
research on mathematics teaching and learning (pp. 257–315). Reston, VA: NCTM.
Phuong-Mai, N., Terlouw, C., & Pilot, A. (2005). Cooperative learning vs confucian heritage culture’s
collectivism: Confrontation to reveal some cultural conflicts and mismatch. Asia Europe Journal,
3(3), 403–419.
Quaresma, M., Winsløw, C., Clivaz, S., da Ponte, J. P., Ní Shúilleabháin, A., & Takahashi, A. (Eds.).
(2018). Mathematics lesson study around the world. Cham: Springer.
Richardson, V. (1996). The role of attitudes and beliefs in learning to teach. In J. Sikula (Ed.), Handbook
of research on teacher education (2nd ed., pp. 102–119). New York, NY: Macmillian.
Ricks, T. E. (2011). Process reflection during Japanese lesson study experiences by prospective secondary
mathematics teachers. Journal of Mathematics Teacher Education, 14(4), 251–267.
Robutti, O., Cusi, A., Clark-Wilson, A., Jaworski, B., Chapman, O., Esteley, C., … Joubert, M. (2016).
ICME international survey on teachers working and learning through collaboration. ZDM-Mathematics
Education, 48(5), 651–690.
Ruffell, M., Mason, J., & Allen, B. (1998). Studying attitude to mathematics. Educational Studies in
Mathematics, 35(1), 1–18.

153

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARIA G. BARTOLINI BUSSI ET AL.

Skott, J. (2015). Towards a participatory approach to ‘beliefs’ in mathematics education. In B. Pepin &
B. Roesken-Winter (Eds.), From beliefs to dynamic affect systems in mathematics education (pp.
3–23). Cham: Springer.
Stigler, J. W., & Hiebert, J. (2009). The teaching gap: Best ideas from the world’s teachers for improving
education in the classroom. New York, NY: Free Press.
Stigler, J. W., & Perry, M. (1988). Mathematics learning in Japanese, Chinese, and American classrooms.
New Directions for Child and Adolescent Development, 41, 27–54.
Sullivan, P., Wood, T. L., & Tirosh, D. (Eds.). (2008). The international handbook of mathematics teacher
education. Rotterdam, The Netherlands: Sense Publishers.
Takahashi, A., & McDougal, T. (2018). Collaborative Lesson Research (CLR). In M. Quaresma,
C. Winsløw, S. Clivaz, J. P. Da Ponte, A. Ní Shúilleabháin, & A. Takahashi (Eds.), Mathematics lesson
study around the world (pp. 143–152). Cham: Springer.
Wang-Iverson, P., & Yoshida, M. (Eds.). (2005). Building our understanding of lesson study. Philadelphia,
PA: Research for Better Schools.
Wang, J., & Lin, E. (2005). Comparative studies on U.S. and Chinese mathematics learning and the
implications for standards-based mathematics teaching reform. Educational Researcher, 34(5), 3–13.
Winsløw, C., Bahn, J., & Rasmussen, K. (2018). Theorizing lesson study: Two related frameworks and
two Danish case studies. In M. Quaresma, C. Winsløw, S. Clivaz, J. P. Da Ponte, A. Ní Shúilleabháin,
& A. Takahashi (Eds.), Mathematics lesson study around the world (pp. 123–142). Cham: Springer.
Xie, X., & Carspecken, P. F. (2008). Philosophy, learning and the mathematics curriculum: Dialectical
materialism and pragmatism related to Chinese and American mathematics curriculums. Rotterdam,
The Netherlands: Sense Publishers.
Yoshida, M. (2008). Exploring ideas for a mathematics teacher educator’s contribution to lesson study:
Towards improving teachers’ mathematical content and pedagogical knowledge. In P. Sullivan,
T. L. Wood, & D. Tirosh (Eds.), International handbook of mathematics teacher education: Tools
and processes in mathematics teacher education (pp. 85–106). Rotterdam, The Netherlands: Sense
Publishers.

Maria G. Bartolini Bussi


Department of Education and Humanities
University of Modena and Reggio Emilia

Silvia Funghi
Doctoral School in Humanities
University of Modena and Reggio Emilia

Alessandro Ramploud
Comprehensive School J. F. Kennedy
Reggio Emilia

154

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

6. MATHEMATICAL CREATIVITY IN THE


CLASSROOM
Teachers’ Conceptions and Professional Development

One of our aims as mathematics educators is to promote students’ mathematical


creativity. Mathematics teachers of all ages hold various beliefs regarding
the nature of creativity in general and mathematics creativity specifically, the
characteristics of mathematically creative students, how to promote mathematical
creativity, and more. The first part of this chapter surveys studies (taken from the
general creativity literature as well as from the mathematics education literature)
which have investigated both prospective and practicing teachers’ beliefs regarding
mathematical creativity in school. The second part of the chapter reviews studies
related to professional development and mathematical creativity, discussing different
approaches to enhancing both prospective and practicing teachers’ appreciation for
mathematical creativity.

INTRODUCTION

The study of beliefs related to mathematics education, and specifically, teachers’


beliefs, has been recognized for some time as a major topic of interest to the
mathematics education community. In the first edition of the International Handbook
of Mathematics Teacher Education, Forgasz and Leder (2008) devoted a chapter to
Beliefs about mathematics and mathematics teaching. That chapter began with a
historical overview of the field of beliefs, which I will not repeat here. However,
it should be noted that while in the 1960s and 1970s, the word “belief” was hardly
heard in connection with education, let alone mathematics education, from the 1980s
on, it has become a recognized and continuously growing field of study. Today,
there are specific topic study groups devoted to beliefs in mathematics education
at several international mathematics education conferences including the Congress
of the European Society for Research in Mathematics Education (CERME) and the
International Congress on Mathematics Education (ICME), as well as the MAVI
conference (MAthematical VIews), a conference entirely devoted to studying
affective issues in mathematics education.
Another fast-growing interest of the mathematics education community is the
study of mathematical creativity in classrooms. As stated by Sriraman (2009),

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_007

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

“mathematical creativity ensures the growth of the field of mathematics as a whole”


(p. 13). While there is no single accepted definition of creativity or creative thinking
(e.g., Runco & Jaegaer, 2012), many researchers from several domains, including
mathematics, but not exclusive to mathematics, note that creativity involves at
least two attributes: originality and appropriateness (e.g., Runco & Jaeger, 2012).
Originality is often related to creating new ideas, along with aspects of divergent
thinking, such as fluency and flexibility (Runco & Acar, 2012). As educators, we
are less concerned with the creativity of a few eminent persons (Big-C creativity)
and more concerned with creativity as it manifests itself in the classroom (mini-c
creativity) (Kaufman & Beghetto, 2009). As students learn new concepts, they
may come up with “novel and personally meaningful interpretation of experiences,
actions, and events” (p. 3). With regard to mathematics classrooms, this aspect of
creativity may manifest itself when a student examines many solutions to a problem,
methods or answers, and then generates another that is different (Silver, 1997).
The teacher has a prominent role in promoting mathematical creativity. It is up
to her or him to choose tasks that may occasion mathematical creativity, create a
supportive environment, and adjust planned lessons according to student responses
(Leikin & Dinur, 2007; Levenson, 2011). Taking this into consideration, this chapter
focuses on teachers’ beliefs related to mathematics creativity in the classroom.
Forgasz and Leder (2008) note that the concern with teachers’ beliefs is not only
theoretical, but practical. Studies have found that teachers’ beliefs may affect their
instructional decisions (Frost, 2010; Schoenfeld, 2011), as well as students’ learning
(Cross, 2009). For example, beliefs about mathematics as a discipline, about the
teacher’s role, and about students’ capabilities, can impact on the type of classroom
environment created by the teacher (Beswick, 2007). Thus, if we are to encourage
teachers to promote mathematical creativity in their classroom, it is relevant to study
their beliefs regarding this issue.
When authoring a chapter for a handbook, one has to decide if the chapter will be
an overview of the existing body of research in the chosen area or a more personal
report of one’s own research in the field. In this chapter, I decided to do a little of
both. Following this introduction, I next review studies of teachers’ beliefs related to
creativity in general, as well as mathematics creativity, specifically. I include beliefs
about creativity in general since creativity research is not limited to mathematics
classrooms, and since, especially in the elementary school grades, generalist teachers
are often responsible for teaching mathematics. In addition, as, beliefs regarding
different issues are not held in isolation, but may be thought of as existing in clusters
(Philipp, 2007), so too, conceptions of mathematical creativity are related to both
conceptions of mathematics education, as well as to conceptions of creativity in
general. I next offer a specific research example, taken from my own practice, of the
complexity of studying beliefs related to mathematical creativity. I then offer a review
of studies related to professional development and mathematical creativity, including
my own, exploring similarities and differences in approaches. The chapter concludes
by noting what was not included in this chapter and issues that need further research.

156

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

TEACHERS’ CONCEPTIONS OF CREATIVITY

Reviewing the Research Field

As mentioned earlier, this chapter reviews studies of teachers’ conceptions related


to both general and mathematics creativity. As such, two major lines of query were
followed. The first involved searching major mathematics education journals, books,
and conferences, and the second involved searching major publications in the field
of creativity research. Within mathematics education, I first searched for studies
mentioning creativity and related notions such as divergent thinking and flexibility,
and then sought out how those may convey findings related to teachers’ conceptions
of creativity within mathematics education. When searching creativity publications,
I searched first for articles related to education, and within those articles searched for
teachers’ conceptions of creativity, and then teachers’ conceptions of creativity and
mathematics. It should also be noted that although studies reviewed in this chapter
were carried out by researchers from various countries, it is beyond the scope of this
chapter to investigate the relationship between cultural differences and conceptions
of creativity.
One of the difficulties in studying beliefs, as well as reviewing studies of beliefs,
is the plethora of terms used to capture this notion. In my search of publications, I
used key words such as conceptions, views, beliefs, and perceptions. In the title of
this section, I ultimately decided to use the term “conceptions,” in line with Philipp
(2007) who wrote that the term conception is “a general notion or mental structure
encompassing beliefs, meanings, concepts, propositions, rules, mental images,
and preferences” (p. 259). This fits many of the methodologies used in studies of
teachers’ beliefs related to creativity. For example, mental images of creativity
may be captured in studies that requested teachers to describe creative students, list
characteristics of creative students, or choose from a list of attributes those they think
are most associated with creative students (e.g., Aljughaiman & Mowrer-Reynolds,
2005; Leikin, Subotnik, Pitta-Pantazi, Singer, & Pelczer, 2013). What it means to
be creative, or mathematically creative, may be captured in studies that requested
teachers to define creativity or to describe the essence of creativity (Aljughaiman
& Mowrer-Reynolds, 2005; Rubenstein, Ridgley, Callan, Karami, & Ehlinger,
2018). In accordance with the notion that, “beliefs, unlike knowledge, may be held
with varying degrees of conviction …” (Philipp, 2007, p. 259), several studies
utilized Likert-scale questionnaires to assess the degree to which teachers agreed
or disagreed with propositions (e.g., Leikin et al., 2013). To summarize, studies of
teachers’ conceptions regarding creativity and mathematical creativity used both
closed and open questionnaires, and referred to teachers’ beliefs, conceptions, views,
perceptions, and values. Within mathematics education, a few studies also included
interviews and qualitative methodologies.
Another challenge when researching teachers’ conceptions of creativity, is the
multitude of issues related to creativity and its promotion. Some of the issues

157

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

discussed include teachers’ definitions of creativity (Kampylis, Berki, & Saariluoma,


2009; Rubenstein et al., 2018), teachers’ conceptions of personality traits associated
with creativity (Aljughaiman & Mowrer-Reynolds, 2005), and factors, such as the
environment, which might affect the manifestation of creativity (Diakidoy & Kanari,
1999). In an attempt to categorize and organize the various conceptions about
creativity, I look towards previous studies which attempted to set apart and unwind
different aspects related to creativity. Rhodes (1961), for example, conceived of what
is called the 4Ps model. The 4Ps refer to (1) the creative Person, including cognitive
abilities, personality traits and biographical experiences; (2) creative Processes or
the methodology that produces a creative product; (3) creative Products which are
unique, novel, and useful ideas; and (4) Press (sometimes referred to Place), which
refers to the environment. This model has been adopted by other researchers when
LQYHVWLJDWLQJFUHDWLYLW\LQHGXFDWLRQ HJ3DYORYLü0DNVLü %RGURå DVZHOO
as in mathematics education (Pitta-Pantazi, Kattou, & Christou, 2018). However,
when studying teachers’ beliefs, there seems to be an important element missing
from this model – teachers’ beliefs related to the nature of creativity, specifically,
whether or not teachers believe creativity can be developed. Because this belief
may directly impact on one’s inclination to promote creativity, this section follows
a different conceptual framework, which incorporates views related to the nature of
creativity as a specific category.
The framework developed by Andiliou and Murphy (2010) and refined by
Berecski and Kárpáti (2018) divides teachers’ beliefs of creativity into three main
strands: beliefs related to (1) the nature of creativity, (2) creative individuals, and
(3) creative environments. The strands of creative individuals and environments are
aligned with strands of creative persons and places from Rhodes’ (1961) model.
Creative processes (Rhodes, 1961) are included under creative individuals. Creative
products (Rhodes, 1961), are included under the category of the nature of creativity.
In the following sections, each strand is sub-divided into additional, more specific
beliefs, related to the main strand, and each sub-division begins with a review of
general creativity studies and then continues with studies related to mathematical
creativity. As will be seen, these strands may overlap and the reader should keep in
mind that there is no one absolute way to categorize the many issues involved.
Finally, when studying teachers’ beliefs, there is always the question of whether
or not these beliefs are in line with research-based findings. With regard to the field
of creativity, there are not only several studies investigating different aspects of
creativity, but at times there are studies which investigate the same aspect of creativity,
yet with different results. For example, is mathematical creativity dependent on
academic ability? Livne and Milgram (2006) found that general academic ability (as
represented by IQ scores), predicted academic ability in mathematics, but did not
predict creativity in mathematics. Leikin and Lev (2013) found that gifted students
(also represented by IQ scores), have higher mathematical creativity scores than
other students, including those learning mathematics at a high level. How can these

158

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

differences be explained? Although both studies took place in Israel, the instruments
used to assess mathematical creativity differed. Assessing creativity is dependent
on how creativity is defined. Kozbelt, Beghetto, and Runco (2010) categorized ten
major theories of creativity, each with their own ways of assessing creativity. Are
creative people risk-takers? Does the environment affect creativity? To answer these
questions, you must answer additional questions. Do you hold by cognitive theories
of creativity, or by systems theories, or by problem finding theories? Taking this
into consideration, this chapter does not judge teachers’ beliefs, but rather describes
them.

Conceptions Related to the Nature of Creativity

Can everyone be creative? One of the main beliefs investigated by several studies
is related to what Berecski and Kárpáti (2018) call the question of distribution, or,
who can be creative. Can everyone be creative or must a person be born that way?
This is related to, although not quite the same as, the question of malleability, or in
other words, can creativity be developed? In an early study of prospective teachers
in Cyprus (Diakidoy & Kanari, 1999), it was found that approximately 75% of the
participants believed that creativity is not a characteristic of all people and that some
children are more creative than others. Yet, about 90% believed that creativity can
be facilitated amongst all children. More recently, in a study of Greek prospective
and practicing elementary school teachers (Kampylis, Berki, & Saariluoma, 2009),
half of the participants believed that only a few students have the “gift” of creativity,
and yet the vast majority agreed that creativity can be developed in all students. In
other words, while teachers may believe that some people are born more creative
than others, most believe that everyone can learn to be creative. Similar beliefs
were found in additional studies (e.g., Aljughaiman & Mowrer-Reynolds, 2005;
Rubenstein, McCoach, & Siegle, 2013).
Regarding creativity in the mathematics classroom, studies found that prospective
secondary school mathematics teachers (Shriki, 2010), as well as practicing
elementary and middle school mathematics teachers (Lev-Zamir & Leikin, 2007)
believe that mathematical creativity can be developed. However, some mathematics
teachers believe that not all students can develop creativity (Levenson, 2017), or that
relatively few students are capable of being creative (Shriki & Lavy, 2012). In one
comparative study of secondary school mathematics teachers in different countries
(Leikin et al., 2013), participants were requested to rate their level of agreement on
a scale of 1–6 with the statement “a creative person is born that way.” The average
across all participants was 4.23, indicating a tendency to agree. Participants from
Mexico were less likely to agree with this statement than participants from India,
Cyprus, Israel, Latvia, and Romania. Reasons mathematics teachers give for why
only some students exhibit mathematical creativity are related to views of individual
characteristics and will be discussed later in the next section of this chapter.

159

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

What are the products of creativity? Conceptions of whether or not a product is


creative depends on one’s view of creativity. Perhaps the most frequent description
of a creative product is that it is original, novel, or unique (Berecski & Kárpáti,
2018). However, as Aljughaiman and Mowrer-Reynolds (2005) point out, it is
not always clear what teachers mean by original. In a previous study (Levenson,
2013), noted that the terms novel and original seem synonymous and are sometimes
used interchangeably in the research literature. However, there might be a subtle
difference between the two. Novel refers to ‘new’, while original may refer to ‘one
of a kind’ or ‘different from the norm.’ Although a ‘one of a kind’ idea may also be
‘new’ and vice versa, it is sometimes the case that an idea, especially one raised in
the classroom, may be new to a student, but if other students have the same idea, it
may not be original.
Berecski and Kárpáti (2018) noted that what is missing from teachers’
conceptions of creative products are additional descriptors, such as appropriateness,
usefulness, and value. Yet, appropriateness is necessary if we are to differentiate
between unique ideas that are of value to learning, and random useless ideas that
may be unique, but are pointless and irrelevant. Rubenstein et al. (2010) found
that among prospective and practicing teachers, of those who described creativity
according to the product, 65% mentioned originality, such as “creativity is an
original thought” (p. 104) and only 21% mentioned usefulness. In Aljughaiman and
Mowrer-Reynolds’ (2005) study, none of the teachers mentioned appropriateness
or usefulness when describing their views of creative production. The researchers
attributed this oversight to teachers’ stressing about academic achievement and thus
viewing creativity as perhaps interesting, but not necessarily of value. However, it
could be that teachers view creativity in the classroom as did Kaufman and Beghetto
(2009), who described students’ creativity as being personally meaningful to the
student, and not necessarily of value or useful to others.
Notions of originality also come up when describing the creative products of
students learning mathematics. In a study of teachers for grades K-3 (Shen &
Edwards, 2017), 70% of teachers spoke about mathematical creativity in terms
of students’ production of ideas that were different from others. Some of those
teachers explicitly mentioned the production of unconventional strategies and
representations. Additional studies have shown that mathematics teachers also
associate mathematical creativity with unusual or original solutions, explanations,
RUSURRIV (PUH$NGR÷DQ <D]JDQ6D÷6KULNL /DY\ ([SODLQLQJ
what was meant by unusual or original, some teachers exclaimed that “it is a proof
I have never seen before”; “a solution that surprises me”; or “explanations I haven’t
thought of” (pp. 4–94). Notably, as with studies of teachers’ perceptions of general
creativity, appropriateness, usefulness, and value were not mentioned. Perhaps this is
due to teachers’ implicitly relating mathematical creativity to the practice of solving
problems, and thus did not feel it necessary to mention usefulness. It could also be
that mathematics teachers, like mathematicians, believe that the results of creative

160

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

work may not always have immediately useful implications, as in a creative proof
for a theorem that has already been proven (Sriraman, 2009).

Creativity and domain specificity. Another major aspect related to the nature of
creativity is whether or not creativity is domain specific. Can creativity be occasioned
in all subjects or is it inherently related to only particular subjects? In two studies
(Diakidoy & Kanari, 1999; Kamplylis et al., 2009) prospective and practicing
elementary teachers were explicitly requested to identify school subjects and
domains that they believed likely to elicit creativity. Amongst the top subjects listed
by approximately 90% of participants were art, theatre, and music; mathematics
was named by less than 65% of participants. As one prospective teacher stated,
“creativity has strong relations with subjects such as art, while mathematics is related
with logic” (Panaoura & Panaoura, 2014, p. 6). Another domain often associated
with creativity at the elementary school level, is language and writing (Aljughaiman
& Mowrer-Reynolds, 2005). Prospective primary school teachers commented that
language classes (in their case, English) offer many opportunities for discussing and
exploring ideas, where students have freedom, can use their imagination, and where
no ‘correct answer’ exists (Bolden, Harries, & Newton, 2010). Yet, associating
creativity with a particular domain does not necessary mean that teachers believe
creativity cannot be occasioned in other subjects. In Kamplysis et al.’s (2009) study,
over 90% of participants claimed that, in general, students can exhibit creativity in a
variety of domains in a variety of ways.
Teachers’ reluctance to associate creativity with mathematics classes, may be due
to their beliefs related to the domain of mathematics, and not necessarily related to
their beliefs of creativity. Some prospective middle- and high-school mathematics
teachers believe mathematics to be a closed domain, where all possible concepts
have previously been invented by mathematicians (Shriki, 2010). Prospective
primary school teachers often suggest that mathematics is a body of knowledge,
based on facts, figures, and rules, with little room for developing independent ideas
(Bolden, Harries, & Newton, 2009). These participants sometimes base their beliefs
on their past experiences, recalling that for them, mathematics in school was always
about getting the right answer, checking the answers in back of the textbook, and
then moving on to the next problem.
Similar sentiments were found when investigating the place of creativity in
classroom discussions (Beghetto, 2007). Prospective middle and secondary school
mathematics teachers were significantly less likely to value unique contributions to
discussions, than other subject teachers. The future mathematics teachers considered
unique responses as potentially disruptive, believing it more important to focus on
the problem at hand, and to follow the curriculum. While this belief is not the same
as believing that mathematics itself is not a creative domain, it may lead students
learning in those classes to believe that mathematics offers limited opportunities for
creativity.

161

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

Conceptions Related to Creative Individuals

What are the personal characteristics of creative students? When discussing


creativity, teachers often describe the personal traits of an individual they consider
creative. Some of these traits have to do with a person’s personality or disposition.
For example, being curious and having an imagination are almost always mentioned
as character traits of creative individuals by teachers coming from different countries,
VXFKDV&KLQD &KDQ &KDQ 6HUELD 3DYORYLü0DNVLü %RGURå WKH
United States and India (Runco & Johnson, 2002). Other characteristics of creative
students identified by United States teachers include: risk-taker, has a sense of
humor (Aljughaiman & Mowrer-Reynolds, 2005), always questions (Chan & Chan,
 UHVRXUFHIXOH[SUHVVLYH 3DYORYLü0DNVLü %RGURå DQGZLOOLQJWR
explore new possibilities (Rubenstein et al., 2018). Runco and Johnson (2002) also
investigated teachers’ beliefs regarding contraindicative traits, or traits that would
not be associated with creative individuals. Some of the adjectives used were lazy,
conforming, and indifferent.
Additional traits mentioned by teachers may be seen as related to intellect or
intelligence. For example, in one study, approximately a third of elementary school
teachers described creative students as intelligent and went on to describe intelligent
students as bright and quick to respond (Aljughaiman & Mowrer-Reynolds, 2005).
Similar descriptions include high-achievers, having good grades, and cleverness
(Chan & Chan, 1999; Gralewski, & Karwowski, 2013). Teachers also mentioned
specific intellectual abilities related to specific domains, such as being artistic,
having a rich vocabulary (Aljughaiman & Mowrer-Reynolds, 2005), likes reading,
and has a high verbal ability (Chan & Chan, 1999). Such beliefs may be related to
teachers’ beliefs regarding creative domains.
Few of the personal character traits mentioned by teachers with regard to general
creativity, were also mentioned with regard to mathematical creativity. This could
be because of the different methodologies used in studies investigating mathematical
creativity. Where studies of teachers’ beliefs about general creativity often included
Likert-scale questionnaires with specific reference to students’ character traits, studies
of teachers’ beliefs about mathematical creativity tended to be more qualitative and
interpretive in nature, and thus perhaps elicited fewer character-related adjectives.
One exception was Leikin et al.’s study (2013). In that study, cultural differences
between teachers’ perceptions, including their perceptions of creative characteristics,
were highlighted. For example, Cypriot teachers viewed the relationship between
depth of knowledge and creativity more strongly than Israeli and Mexican teachers.
Romanian teachers associated mathematical creativity with achieving high scores in
mathematics contests. In Israel, a few prospective teachers associated mathematical
creativity with curiosity (Shriki, 2010). As one participant stated, “People cannot
be creative unless they are curious about what they do” (p. 171). What mathematics
teachers do stress, are the processes involved in mathematical creativity. This is
discussed in the following section.

162

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

What are the processes and associated affect involved in producing a creative
product? At times, it is difficult to separate teachers’ perceptions of character
traits from behavior associated with creativity. For example, in Runco’s (2002)
study, teachers listed “flexible” as a character trait, in the same way that they listed
“friendly.” Flexible can mean that a student is amenable to change in general. It does
not necessarily mean that the student is flexible in his or her thinking processes.
The same may be said of “always questioning.” Is this a personality trait, as listed
by Chan and Chan (1999), or is it part of the creative process that leads to a creative
product? Perhaps it is both.
Both prospective and practicing teachers describe creative students as those
who generate ideas, think differently, and think outside of the box (Aljughaiman &
Mowrer-Reynolds (2005; Rubenstein et al., 2018), although what it means to think
differently was not explicated. Aljughaiman and Mowrer-Reynolds (2005) pointed
out that participants did not associate creativity with divergent thinking processes
such as approaching a task from different directions or arriving at multiple solutions.
In contrast to teachers’ perceptions of general creativity processes, when it comes
to creative mathematical processes, teachers seem to have more explicit conceptions.
Prospective elementary school teachers, for example, believe that practical activities,
such as physically measuring the lengths and widths of objects in the classroom, can
lead to creativity (Bolden et al., 2009). Similarly, some teachers mention the use
of manipulatives, such as the use of geo-boards, as a factor in promoting creativity
(Levenson, 2013). Other teachers claim that visual thinking, or as one teacher called
it “creative thinking of space,” is part of the creative process (Levenson, 2013, p.
279). Perhaps the most mentioned process by mathematics teachers was flexibility.
Specifically, prospective elementary school teachers mentioned computation
flexibility, such as computing an addition problem in several ways (Bolden et al.,
2009). However, in general, teachers connect mathematical creativity with flexibly
solving problems, that is, solving problems in multiple ways, and using a variety of
strategies (Leikin et al., 2013). Teachers also mention using non-algorithmic solution
methods and solving problems that integrate different mathematical domains (Leiken
et al., 2013; Levenson, 2013) or integrating mathematics with other school subjects
(PUH$NGR÷DQ <D]JDQ6D÷/HYHQVRQ 
Studies have shown that when teachers discuss cognitive processes, they often
raise associated affective issues. As with processes, affective issues can sometimes
be called personality traits. For example, as mentioned above, teachers often believed
that a curious person is likely to be creative (Runco & Johnson, 2002). However,
curiosity might also be seen as the motivator of creative processes. In Levenson’s
(2013) study, a certain task was thought to promote creativity because “You can
see here [in the task] an element of a riddle and of fun, which creates challenge and
curiosity” (p. 286). Among prospective and practicing teachers, there seems to be a
consensus that mathematical creativity also involves enjoyment. In Shriki’s (2010)
study, one participant stated, “I think that working creatively is concerned with
fun. … Isn’t that what creativity is about? I think it is. Enjoy while you create and

163

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

love your creations” (p. 171). In Leikin et al.’s study (2013), participants mentioned
enjoyment in solving problems at various levels of difficulty.

Who is a creative teacher? As opposed to studying teachers’ conceptions of


creative students, a few studies investigated teachers’ conceptions of what it
means to be a creative teacher. In a Portuguese study (Morais & Azevedo, 2011),
elementary and secondary school teachers characterized creative teachers as those
who exhibit enthusiasm, do not have a set routine, and promote student autonomy.
Lev-Zamir and Leikin (2011) found that elementary and middle-school mathematics
teachers perceive a creative teacher as one who is flexible, i.e., can transform a
given task, either mathematically or pedagogically. A teacher exhibits originality
when generating new tasks and ideas during a lesson. In Bolden’s (2010) study,
primary school teachers linked creative teaching in mathematics to the creative
use of resources (e.g., manipulatives and technology) and to linking mathematics
to everyday examples. Amongst mathematics teachers, there is also agreement
that a creative mathematics teacher is one who enjoys mathematics (Leikin et al.,
2013).

Conceptions Related to Creative Environments

When discussing environmental factors related to creativity, there are both factors
which support and those that hinder creativity. Prospective and participating
teachers acknowledge that the environment plays an important role in promoting
creativity (e.g., Diakidoy & Kanari, 1999; Kampylis et al., 20009). They do not
agree, however, that the school is the best environment for doing so. Kampylis et
al. (2009) theorized, that the reason teachers were skeptical about the ability of
the school environment to support creativity was the Greek tradition of teaching
through lectures and recitation methodologies. Many participants in that study
also referred to the textbooks and other curriculum materials which do not allow
for creative thinking. Other studies found that when asked to state hindrances to
developing creativity, teachers most often cited external regulation, such as not
having enough classroom time, testing, and a standardized curriculum (de Souza
Fleith, 2000; Rubenstein et al., 2018).
A standardized curriculum might also be related to why some teachers do not
believe creativity may be developed in mathematics classes. For example, in Bolden
et al.’s (2009) study, participants claimed that English, art, and even science had
fewer set goals, and were therefore more conducive to creativity development. Other
hindrances mentioned by teachers, although to a much lesser extent, were the lack
of support from parents, administration, and other colleagues (Rubenstein et al.,
2018). Finally, a small percent of respondents acknowledged that students who are
concerned about peer or teacher recognition, may hold back on creativity. Perhaps
the above perceived environmental constraints add to teachers’ feelings that it is not

164

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

their responsibility to foster students’ creativity (Aljughaiman & Mowrer-Reynolds,


2005).
The types of environments which teachers do perceive as supporting creativity
are those that emphasize autonomy and independence, where students are given the
opportunity to make choices and ask questions (Diakidoy & Kanari, 1999). Another
factor according to some teachers is allowing students some unstructured time (de
Souza Fleith, 2000).
With regard to mathematical creativity and the environment, little has been said
by teachers. The physical environment was mentioned by a teacher in Levenson’s
(2013) study, who seemed to think that being indoors in a classroom restricts
creativity. She claimed that if she wanted to promote creativity, she would implement
mathematical activities outdoors, “I believe that by just being outside the classroom
walls frees you from the shackles of convention, which in turn encourages all kinds
of creativity” (p. 278). In general, however, most mathematics teachers agreed with
other teachers that standardized curricula and tests, impede on students’ creativity in
mathematics (e.g., Bolden et al., 2009).

INTERTWINING BELIEFS: IS MATHEMATICAL CREATIVITY RELATED TO


MATHEMATICAL EXCELLENCE?1

In many countries, including Israel, students are often tracked in mathematics


according to achievement levels (Cogan, Schmidt, & Wiley, 2001). Even when
students are not tracked, there is great concern among teachers regarding how to
handle students of mixed abilities in their mathematics classroom. As seen in the
above review, there are a variety of beliefs that may be associated with this aspect
of teaching mathematics and the promotion of creativity. Regarding the nature of
creativity, for example, some teachers claim that only high achievers can be creative
(Chan & Chan, 1999). Regarding creative individuals, some teachers believe that
a creative student is of high intelligence, verbal ability, and intrinsic motivation
(Aljughaiman & Mowrer-Reynolds, 2005). These beliefs may affect if and how
creativity is promoted for students learning mathematics at different levels.
This section presents a study that focuses on mathematics teachers’ perceptions
regarding the relationship between mathematical creativity and mathematical
excellence. If we wish to encourage teachers to promote mathematical creativity
among all students, not only among the highest mathematics achievers, then it is
important to investigate teachers’ beliefs regarding this issue. The specific research
questions of the study were: (1) Do teachers believe that there is a relationship
between mathematical creativity and mathematical excellence, and if so, what types
of relationships do they believe exist? (2) What beliefs regarding mathematical
creativity surface, as teachers describe the relationship between mathematical
creativity and excellence? (3) Are different beliefs regarding creativity associated
with different beliefs regarding the relationship between mathematical creativity and
excellence?

165

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

Setting

For the past several years, I have been the instructor for an elective graduate
course, Creativity in Mathematics Education. Some students are already practicing
mathematics teachers, while others concurrently study toward their teaching certificate.
At the beginning of the course, I always give the following assignment: Choose a task
or activity from a mathematics textbook that in your opinion promotes mathematical
creativity and explain why this task has the potential to promote mathematical
creativity. Results of this part of the assignment were reported in Levenson (2013)
and showed that most teachers associated creativity with being different and unusual.
Relevant to this chapter and the focus on beliefs, several teachers raised the
issue of accessibility, that is, they chose the particular task, among other reasons,
because all students in their class would be able to engage in the activity. Samples
of teachers’ reasoning include: “… there isn’t a student who cannot participate in
this activity, even special-needs students [can participate],” “Every student can find
his own unique solution method,” and “the task promotes the use of … graphs,
algebra, numbers, and words and does not limit the solution to a specific media thus
allowing students the possibility of expressing themselves in the area where they are
strongest” (Levenson, 2013, p. 286). Another value mentioned by several teachers
was encouraging individuality. One teacher chose a certain task because “every
student can look at the given [data] in a different way and can create a different
formula from the others” (p. 286).
The above statements hint at an implicit belief that all children should have access
to tasks that can promote creativity and furthermore, that such tasks are beneficial,
specifically when taking into account the diversity of the classroom. In an effort to
investigate these beliefs explicitly, in the last two years, the following question was
added to this homework assignment: “There are those who say that mathematical
creativity is related to excellence in mathematics. What is your opinion?” The term
“excellent” is a commonly used term amongst teachers when describing students
who have high grades in mathematics.
In the following section, the findings from 45 participants are reported. The range
of teaching experience was from 0–25 years. None of the participants had previously
taken a formal course related to creativity.

Data Analysis and Findings

A grounded theory approach was used to analyse the data. The initial reading
categorized participants’ responses into “yes, excellence in mathematics is related
to creativity,” “no, there is no relationship,” and “undecided.” Further readings led
to a finer categorization scheme based on the type of relationship teachers claimed
to exist between mathematical creativity and mathematical excellence. The author
and another mathematics education researcher independently categorized all
participants’ responses. Where there was disagreement, a discussion ensued until

166

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

Table 6.1. Types of relationships between creativity and excellence

Category F (%) Examples of teachers’ statements

A: Mathematical 13 (29) “When a student is good at mathematics, his


excellence precedes self-confidence rises, which causes him to dare more
mathematical creativity and to try different solution methods without fear of
failure.”
B: Mathematical 19 (42) “Creativity comes from having an open mind,
creativity precedes solving problems in many different ways, which
mathematics excellence leads to excellence.”
“Mathematical creativity promotes excellence in
mathematics … However, excellence in mathematics
does not promote creativity because creativity is
genetic and cannot be acquired.”
C: Creativity and 4 (9) “The relationship between creativity and excellence
excellence in is two-way. While it may be that stronger
mathematics are mathematics students exhibit more mathematical
reciprocally related creativity, if teachers promote creativity among
the weaker students, they will become stronger in
mathematics.”
D: There is a non- 2 (4) “There is a relationship between mathematical
influential relationship creativity and mathematical excellence, but one is
between mathematical not a sufficient condition for the other.”
creativity and excellence
E: Mathematical 4 (9) “Creativity in mathematics can be developed and
creativity and excellence acquired, even among lower achieving mathematics
are not related students. It is dependent mostly on a supportive
environment of which the teacher is responsible.”
Undecided 3 (7) “It depends on how one defines excellence in
mathematics.”

agreement was reached. A third mathematics education researcher then categorized


20% of the responses, ending in 90% agreement between researchers.
Table 6.1 presents the six different categories of relationships found in the data,
along with their frequencies (F in %), and examples of participants’ statements. The
majority of teachers (84%) believed that some relationship between mathematical
creativity and mathematical excellence exists. In addition, considering all those who
believed that mathematical creativity can promote mathematical excellence, including
those that believed the relationship to be mutual, we find that half of the participants
believed mathematical creativity to have some influence on mathematical excellence.
After completing the above categorization, a second analysis investigated
inferred beliefs related to creativity. These were assigned to one of four categories:
beliefs related to (1) creative processes, (2) the product of creativity, (3) the nature

167

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

of creativity, including how creativity might be affected by the environment (e.g.,


opportunities afforded in some classroom), and (4) affective issues.
In Table 6.2 we see that the same creative processes were mentioned more or
less by all teachers, regardless of their beliefs concerning excellence and creativity.
Several teachers pointed to opportunities given to different students as impacting on

Table 6.2. Inferred beliefs related to each category

Category Associated implicit beliefs


Creative Creative The nature of Affective
processes products creativity issues

A: Mathematical Breaking down a – Opportunities Not fearing


excellence problem, thinking provided especially failure,
precedes out of the box, for excellent contributes to
mathematical trying different mathematics creativity
creativity solution methods students promote
creativity.
B: Mathematical Solving problems Meaningful Creativity is an Creativity is
creativity in different ways, and un- innate trait. enjoyable.
precedes using various conventional
mathematics thinking processes solutions
excellence
C: Creativity Thinking out of Interesting Creativity can –
and excellence the box solutions be developed;
in mathematics opportunities
are reciprocally provided in
related different classes
can have a
strong impact
on mathematical
creativity.
D: There is a – – – –
non-influential
rela-tionship
between
mathematical
creativity and
excellence
E: Mathematical Thinking – Creativity is –
creativity and differently dependent
excellence are mostly on the
not related environment of
which the teacher is
responsible.

168

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

the development of mathematical creativity, believing that higher-achieving students


have greater opportunities. Regarding affective issues, not fearing failure, was not
mentioned as an attribute of a creative individual, but rather as an outcome of being
excellent in mathematics, which can then contribute to creativity. In the case of
enjoyment, one teacher claimed that being creative is enjoyable, and this would
motivate the student to excel in mathematics. In other words, for those teachers, the
affective side was an intervening variable between excellence and creativity.

Discussion

Most mathematics teachers believed that mathematical creativity and mathematical


excellence are related. This is not surprising as previous studies pointed out that
teachers often associate creativity with high intelligence and achievement (e.g., Chan
& Chan, 1999). This study adds to those studies by exploring in depth how teachers
perceive of this relationship. For example, teachers do not necessarily believe that
ability directly affects creativity. Some believe that ability leads to greater motivation
and less fear of failure, which can then lead to creativity. Knowing this aspect of
teachers’ beliefs may help teacher educators address this issue, discussing with
teachers how to mitigate fear of failure among all students, possibly then motivating
those same teachers to promote creativity among all students.
This study also showed that approximately half of the teachers who believed there
is a relationship between creativity and excellence, also believed that creativity can
lead to excellence in mathematics. In other words, it is not only high-achieving
mathematics students who can be mathematically creative. This belief can be a first
steppingstone when encouraging teachers to promote creativity in their classroom;
if they think that creativity can lead to excellence, they may be more interested in
learning how to promote mathematical creativity. That being said, not all teachers
who believed that creativity can lead to mathematical excellence expressed a belief
that creativity can be promoted. One teacher believed that mathematical creativity
leads to excellence, but explicitly stated that creativity is an innate trait.
In answering the second research question, many beliefs found in this study were
aligned with finding of previous studies, such as mathematically creative people
are open-minded (Aljughaiman & Mowrer-Reynolds, 2005), come up with unique
solutions (Leikin et al., 2013), and propose different approaches to solving problems
(PUH$NGR÷DQ  <D]JDQ6D÷   /HVV GLVFXVVHG LQ SUHYLRXV VWXGLHV ZHUH
mathematics teachers’ beliefs regarding the environment and mathematics creativity.
In addition to curriculum and testing constraints mentioned in previous studies (de
Souza Fleith, 2000; Rubenstein et al., 2018), this study showed that mathematics
teachers view the different environments of different level mathematics classrooms,
as also affecting creativity. Students at all levels of mathematics must take mandated
tests, and all are constricted by a mandated curriculum. However, as teachers in
this study claimed, opportunities afforded at different levels, may be in the hands

169

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

of the teacher. This is another positive finding of this study, as it hints that teachers
recognize their role in supporting mathematics creativity.
As to the third research question, findings indicated that the same beliefs about
creativity were held by participants with opposing beliefs regarding the relationship
between excellence and creativity. One possible reason for this, is that beliefs are not
isolated entities, but part of a system of beliefs with complex relationships (Green,
1971). Thus, for example, a belief that the environment is a factor in promoting
mathematical creativity, may lead one teacher to claim that excellence (because of
opportunities given to excellent students) leads to mathematical creativity, while
another teacher claims that since it is up to the environment, mathematical creativity
and excellence are unrelated. A second possible reason is that participants’ various
teaching experiences might have affected the variety of belief interactions. Beswick,
Callingham, and Watson (2012) found that teaching experience can have a great
influence on one’s beliefs, including beliefs about mathematics as a school subject.
This in turn might affect teachers’ beliefs regarding what it means to excel in
mathematics. What this study does bring to light is how beliefs interact, and if we
want to encourage teachers to promote mathematical creativity, we need to take into
consideration their beliefs regarding creativity and teaching mathematics. The next
section focuses on this issue.

PROFESSIONAL DEVELOPMENT AND MATHEMATICAL


CREATIVITY: SOME EXAMPLES

In the previous sections of this chapter, I discussed both general and mathematical
creativity. However, in the case of professional development for mathematics
teachers, I consider specifically the work of the mathematics teacher. This section
focuses on prospective teachers and practicing teachers in separate sub-sections,
respectively. In both of these sub-sections, participants’ conceptions related to
mathematical creativity are highlighted. The last sub-section offers reflections.

Educating Prospective Teachers

Efforts for raising teachers’ awareness of mathematics creativity and their ability
to support students’ mathematical creativity may begin before participants become
actual teachers, often during content and methods courses. In Turkey, for example,
Kandemir and Gur (2007), worked with prospective secondary school mathematics
teachers during a content knowledge course. Participants were introduced to methods
for developing creativity, such as brainstorming, question production, using your
imagination, and thinking of alternatives. Participants reported that the course and
the methods learned during the course raised their interest in mathematics, showed
them how to see problems from different perspectives, and led them to believe that
creativity can be developed.

170

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

Lee (2017) worked with prospective secondary school mathematics teachers


in Korea during a methods course. The explicit aim of the course was to develop
participants’ awareness for how textbook tasks can be modified to increase students’
creativity. Lee viewed creativity as the confluence of divergent and convergent
thinking, recognizing that a balance is necessary. Too much divergence may result
in less meaningful mathematics. Too much convergence, can lead to stagnation.
During the course, participants analyzed tasks from a creativity perspective, studied
examples of modified tasks, engaged in modifying tasks, and finally implemented
modified tasks during micro-teaching. For example, in one case, participants chose
the following standard task: Find the nth term of the sequence 1, 3, 5, 7, 9, … and
the nth term of the sequence 1, 2, 4, 8, 18, … To this task, participants added a few
questions in order to promote divergent thinking, such as: “What is the nth term
of the sequence in which the first two terms are 3 and 9? Explain why. Find as
many as possible sequences that have 3 and 9 as their first two terms” (p. 1002).
Participants were requested to explicitly focus on mathematical and pedagogical
aspects of divergent thinking, such as extending students’ existing knowledge and
encouraging risk-taking, as well as convergent thinking, such as making use of
existing knowledge and following algorithms.
While Lee’s (2017) study did not focus on participants’ beliefs related to creativity,
the findings brought to light several perspectives regarding mathematics creativity
in the classroom. For example, in the beginning of the course, many participants
regarded problem-solving tasks as appropriate for promoting creativity, regardless
of the demand for divergent or convergent thinking. Participants believed these tasks
to be cognitively challenging and associated challenge with creativity. At the end
of the course, some teachers continued to choose such tasks as having the potential
to promote creativity. Participants who modified tasks and placed emphasis on
convergence, did so out of concern for slow learners, reminding us of teachers who
associate creativity with high achievers (Levenson, 2017). Participants who tended
to modify tasks by placing emphasis on divergence, expressed their concern with
students’ low motivation. These participants believed that tasks, which are open in
terms of representations, strategies, and solutions, are more student-friendly and can
increase students’ motivation for learning mathematics. This is in line with studies
showing teachers’ beliefs of creativity as relating to fun and enjoyment (Levenson,
2013). In general, we see how teachers’ beliefs related to creativity can impact on the
practice of task implementation.
Another intervention with prospective teachers was described by Panaoura and
Panaoura (2014), who investigated the outcomes of an elective course given to
third-year prospective teachers in Cyprus. The explicit aim of the course was to
raise participants’ awareness of recent trends in mathematics education and reforms
in school mathematics. These trends included enabling young students to think
critically and creativity in mathematics. In other words, enhancing future teachers’
knowledge of mathematical creativity was part of a larger aim to raise awareness
of school mathematics reforms. During the course, prospective teachers discussed

171

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

their conceptions of creativity in mathematics, were introduced to research and


theories related to mathematical creativity, solved problems which may occasion
mathematical creativity, and analyzed textbooks that were published in accordance
with the new reforms in Cyprus.
Similar to Lee’s (2017) study, one of the assignments during the course was to
choose an existing textbook task and change it in such a way that it would be suitable
for promoting mathematical creativity. For this assignment, participants exhibited
knowledge of mathematical creativity, changing routine problems to those that could
be solved in multiple ways, fostering fluency and flexibility. For the final assignment,
participants were asked to develop two lesson plans for teaching mathematics.
Most prospective teachers included one exploration activity in their lesson plan,
with the rest being routine activities. Upon further investigation, it seemed that the
participants did not believe that creativity could be a specific aim of a lesson plan.
Furthermore, although their initial conceptions of mathematical creativity focused
on originality, participants expressed that working this aspect into a lesson plan
was most difficult. Panaoura and Panaoura (2014) concluded that appreciating the
importance of creativity may not be enough for applying theory to practice.
A different type of intervention was described by Shriki (2010) who worked
with prospective secondary school mathematics teachers in a methods course. The
explicit aim of the methods course was to introduce students to didactical methods
and approaches for teaching geometry and algebra. However, the author also stated
that she designed the course to help prospective teachers rethink their conceptions
of what mathematics is (recall that teachers sometimes claim mathematics to be a
closed domain (e.g., Bolden, 2009)), and to develop their awareness of mathematical
creativity. Thus, creativity is a specific aim, although only a part of other, more
general aims.
Shriki describes that during the course, prospective secondary teachers were
encouraged to “work like mathematicians.” Participants had to come up with a new
geometrical concept, find as many properties as they could for their concept, and
to formulate relevant theorems. During the process, participants were also required
to reflect on various aspects of creativity. Participants worked on this task during
the course and also at home. Over a period of a few weeks, participants shared
their creations and investigations with the group and the instructor, who explicitly
discussed with participants their beliefs regarding the nature of mathematics and the
meaning of creativity.
At the beginning of the course, participants’ beliefs related to creativity focused
on the product of creativity, such as inventing new theorems or coming up with
original proofs. According to Shriki, as the prospective teachers reflected on their
own efforts to create a new concept, their beliefs related to creativity began to shift.
Participants began to describe mathematical creativity in terms of processes, such
as thinking differently and asking questions. By the time participants shared their
work in class, they began talking about the affective side of creativity, such as
enjoyment, curiosity, and taking risks. Creative products were now described as new

172

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

for oneself or a result of one’s own thinking. The project also enhanced participants’
mathematics knowledge, specifically knowledge of definitions and their role in
mathematics. Equally noteworthy, Shriki also describes the prospective teachers’
shifts in their beliefs of mathematics as a domain. In the beginning of the course,
participants viewed mathematics as a closed domain, where only mathematicians
may be creative. By the end of the process, participants viewed mathematics as
beautiful, full of patterns, where students may also pose their own problems and
come up with new concepts.
To sum up, for teacher educators, developing prospective teachers’ appreciation
for mathematical creativity is bound up in numerous other aspects of teacher
preparation. While Lee (2017) was able to dedicate an entire course for enhancing
future teachers’ ability to promote mathematical creativity, most teacher education
programs do not have this option. Both Shriki (2010) and Panaoura and Panaoura
(2014) found ways to educate future teachers about mathematical creativity during
regular methods courses. In general, all of the studies mentioned the importance
of raising future mathematics teachers’ awareness of mathematical creativity, while
acknowledging the challenges of doing so.

Professional Development for Practicing Teachers

When it comes to practicing teachers, time constraints, among other variables (e.g.,
national reforms) often dictate the length and intensity of professional development,
as well as the content and methods. Siswono (2015), for example, offered a
workshop to practicing elementary school teachers in Indonesia. The explicit aim of
the workshop was to introduce teachers to a problem solving-posing based learning
model of teaching mathematics, and developed teachers’ ability to implement
materials and design problems that would encourage students’ fluency, flexibility,
and novelty when solving problems.
Leikin and Levav-Waynberg (2009) studied a professional development program
for practicing secondary school mathematics teachers, that took place over a two-
year period. The explicit aim of the program was to develop teachers’ knowledge
for teaching school mathematics by employing multiple-solutions connecting tasks,
that is, multiple-solution tasks that can promote the “construction of mathematical
connections in the students’ minds” (p. 204). Implicitly, the course also aimed at
developing teachers’ ability to solve and identify such tasks and increase their
awareness of the importance of learning with these tasks. The authors inferred that
implementing multiple-solution tasks would require a paradigm shift for teachers,
and that such a shift is difficult for teachers who have never experienced learning
or teaching in such a way. Although multiple-solutions tasks are often associated
with developing creativity, fostering knowledge and appreciation for mathematical
creativity was not among the specific aims of the program.
The first stage of this study took place over a period of eight months. Teachers
solved multiple-solution tasks and developed their ability to solve problems in

173

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

multiple ways. They discussed what it meant to find different solutions, and
what types of differences can arise such as using different representations of a
mathematical concept, different properties of mathematical concepts from the same
branch of mathematics, and using different mathematical tools based on different
branches of mathematics. In the process, they also developed their ability to identify
such tasks, and an appreciation for this type of activity. At this point, we may say
that participants deepened their mathematical knowledge, especially regarding
mathematical connections, as well as their pedagogical knowledge related to tasks
(Ball, Thames, & Phelps, 2008).
In the second stage of the study, teachers were requested to implement such
activities in their classrooms. At this point, teachers began to notice how solving
multiple-solution tasks may contribute to their students’ mathematics knowledge,
as well as students’ developing creativity. Some teachers reflected on the benefits,
to what we may call, creative processes. For example, teachers specifically noted
students’ ability to think about many different topics at once or being able to see an
object from several angles. Other teachers began to pay attention to creative products,
such as the teacher who began to notice original solutions. Finally, teachers remarked
on students’ increased interest and motivation when solving multiple-solution tasks.
In other words, teachers also enhanced their knowledge regarding students and
mathematics (Ball, Thames, & Phelps, 2008), in this case knowledge of students
and mathematical creativity. Leikin and Levav-Waynberg also pointed out that as
teachers began to notice their students’ mathematical creativity, their conceptions of
which students are able to produce original ideas changed. Teachers also developed
a positive disposition towards implementing such tasks in their classes, which was
not present at the start of the program, and which may impact on their future practice
regarding the promotion of mathematics creativity.
The next study I review (Levenson, 2015), is the investigation of the graduate
course I instructed, entitled Creativity in mathematics education, mentioned in a
previous section. The explicit aims of the course were to introduce participants to
theoretical and practical perspectives of mathematical creativity. The course took
place over a semester and consisted of 14 ninety-minute lessons. Participants were
introduced to theories related to promoting and assessing mathematical creativity
in the classroom (e.g., Leikin, 2009; Silver, 1997; Sriraman, 2009), and engaged
in mostly short tasks, with the intention of having them experience mathematical
creativity in several ways, such as problem posing tasks (Silver & Cai, 2005),
opened-ended tasks (Kwon, Park, & Park, 2006), creating new geometrical concepts
(Shriki, 2010), and overcoming algorithmic fixation (Haylock, 1997). During the
course, groups of participants implemented creativity-promoting tasks with other
course participants, with the aim of gaining some experience implementing such
tasks. In the beginning, middle, and end of the course, participants were required to
choose a task or activity that they believed had the potential to promote mathematical
creativity and explain their choice.

174

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

The study follows one secondary school mathematics teacher, Ava, and reports
on her changing perspectives of tasks that may occasion mathematical creativity, as
well as her associated perspectives of creativity. At the beginning of the course, Ava
chose a task from the matriculation examination for 12th-grade students studying
mathematics at an advanced level, claiming that that in her opinion this task
promoted creativity because it required the student to think of an unconventional
solution path.
Triangle ABC is an equilateral triangle. Two of the vertices are A(0, a) and B(0,
-a), a > 0. Show that the sum of the distances from the sides of the triangle to
some point within the triangle, is dependent solely on a.
She believed that creativity was appropriate for talented students, who have sudden
insights. For the second task, she chose the following textbook problem: Calculate the
area of a triangle whose vertices are (3, 5) (-2, -2) (-1, 3). Ava claimed that because the
problem could be solved using trigonometry, analytical geometry, plane geometry,
and more, it is a task that connects different mathematical domains, promoting
fluency, flexibility, originality. She specifically related to theories introduced in the
course. By the third assignment, Ava chose a task that required generalization, had
an element of surprise, and, according to Ava, would be appropriate for students of
all levels, with the teacher playing an active role.
To sum up this section, working with practicing teachers can be challenging, but
also rewarding. When time is short, a workshop, such as that offered by Siswono
(2015), can offer practical advice for teachers, enabling them to enhance students’
engagement with mathematical creativity. When there is a bit more time, or when
the setting is more formal, theory can be introduced and integrated with practice. In
my own study, the teacher, Ava, specifically stated that the theoretical side of the
course was just as important to her as the tasks. However, Ava was not required to
implement in her classroom what was learned during the course. Studies have shown
that professional development that is not integrated into the daily life of the school
may be less effective (Garet et al., 2001; Tirosh & Graeber, 2003). In Leikin and
Levav-Waynberg’s (2009) study, in-class practice with creativity was an integral
part of the program.

Learning from Professional Development Studies

While the studies in the previous sections differed in several significant ways (e.g.,
participants, setting, aims, methods), we can learn from each about the possibilities
for teacher education to impact on teachers’ beliefs related to creativity. To begin
with, some of the studies (Leikin & Levav-Waynberg, 2009; Shriki, 2010) showed
how mathematics knowledge is intertwined with beliefs about mathematics, which
in turn may impact on beliefs about mathematical creativity. In Leikin and Levav-
Waynberg’s (2009) study, as teachers gained expertise with solving and identifying
multiple-solution tasks, they began to notice students’ creativity, and their conceptions

175

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

of students’ creativity began to change. In Shriki’s study, the prospective teachers’


conceptions of the nature of mathematics changed in tandem with their conceptions
of mathematical creativity. These findings are in line with other recent studies,
bringing attention to the interaction of knowledge and beliefs (e.g., Charalambous,
2015; Drageset, 2010).
At the core of each program, although to a different degree of frequency and
intensity, was engaging participants with tasks that have the potential to occasion
mathematical creativity. Watson and Sullivan (2008) claimed that mathematics
tasks for teachers have multiple purposes: to inform teachers about the variety and
purpose of classroom tasks, to provide opportunities to learn more mathematics, to
provide insight into the nature of mathematical activity, and to stimulate teachers’
theorizing about students’ learning. Not every program may use tasks in each one of
those ways. In Leikin and Levav-Waynberg’s (2009) program, as teachers engaged
with tasks, they began to theorize about students’ mathematical creativity. In Shriki’s
methods class, prospective teachers gained insight into the nature of mathematics, as
well as the nature of mathematical creativity. In Levenson’s (2015) and Lee’s (2017)
studies, participants became informed as to the variety of tasks that may be used to
occasion mathematical creativity in the classroom.
In addition to the purposes of tasks mentioned by Watson and Sullivan (2008),
there was an additional purpose of having participants engage in mathematical tasks,
and that was to have participants experience mathematical creativity for themselves.
Several of the studies above (e.g., Lee, 2017) mentioned that because participants
had not experienced creativity during their school years, changing their perspectives
of mathematical creativity, including who can be mathematically creative, and what
types of tasks to implement in the classroom, was very difficult. In fact, in Shriki’s
(2010) study, it was only as prospective teachers worked on their own creativity, did
they begin to change both their beliefs about mathematics and mathematical creativity.
Ambrose (2004), in her study of prospective teachers’ orientation to teaching
mathematics, claimed that instead of attempting to replace old beliefs with
new ones, it is more effective to build on existing beliefs. In her study, it was
the intense experience of the prospective teachers working with students, that
allowed beliefs to grow. It may be that in the case of mathematical creativity,
the catalyst for growth is participants’ experiencing mathematical creativity for
themselves. Citing Green (1971), Ambrose (2004) also points out that although
beliefs may be arranged in clusters, these clusters might be isolated from one
another. Thus, “The belief about the importance of creativity for learning may
not be connected to beliefs about mathematics because the prospective teachers
have not had creative experiences in mathematics” (Ambrose, 2004, p. 94). This
may be true of both prospective and practicing mathematics teachers. It is the
responsibility of the teacher educator to help teachers connect their beliefs, in this
case, their beliefs about mathematics, creativity, and mathematical creativity in the
classroom. Having teachers experience mathematical creativity may be one way of
stimulating this process.

176

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

SOME CONCLUDING REMARKS

A chapter dedicated to teachers’ beliefs, even when narrowing it down to beliefs


regarding creativity, cannot possibly cover all relative issues. The first section of this
chapter reviewed studies related to teachers’ conceptions of students’ creativity, with a
small section related to conceptions of teacher creativity. Such studies are also related
to teachers’ beliefs regarding what it means to learn and teach mathematics (e.g., Swars,
Smith, Smith, & Hart, 2009), an avenue not reviewed here. Another path not taken was
comparing teachers’ conceptions of creativity to theories developed by social scientists
which are based on studies of individuals’ creativity (e.g., Saracho, 2012). Finally, this
chapter only touched upon teachers’ beliefs regarding how creativity, in general and in
mathematics, can be promoted amongst students (e.g., Levenson, 2013); but there are
several other studies which focus on this more explicitly (Berecski & Kárpáti, 2018;
Kampylis et al., 2009), as well as on teachers’ self-efficacy for developing students’
creativity (Kampylis et al., 2009; Park, 2013), and the relationship between culture and
creativity (e.g., Kim, 2009; Leikin et al., 2013). I acknowledge these voids and suggest
to the interested reader additional avenues worth reviewing.
Missing not only from this chapter, but from the research literature as well, are
studies which investigate the relationship between teachers’ espoused beliefs related
to mathematical creativity and classroom practice. Taking into consideration that
several studies found a connection between teaches’ beliefs and various elements of
their practice (e.g., Beswick, 2007; Cross, 2009; Polly et al., 2013) it makes sense
to investigate this relationship with regard to mathematical creativity. To a limited
extent, Levenson (2013, 2015) found a relationship between teachers’ conceptions
of mathematical creativity and the types of tasks teachers would choose in order to
occasion mathematical creativity in their classrooms. But choosing tasks is only one
part of a teacher’s practice. Research inside the classroom is still needed.
An additional avenue for future research is the possibility of sustaining the
effects of professional developments programs, such as those described in this
chapter. At the end of their study, Leikin and Levav-Waynberg (2009) claimed that
the participants had transformed into members of a community of practice, with
shared goals, norms, and vocabulary. While this is a positive development, what
happened to the teachers when the program ended? Zehetmeier (2015) noted that
sustainment of a program’s impact may be negatively affected when teachers who
had mutually supported each other, move to different locations. Another question
regards the methodology of studying beliefs and practice. In my own study of the
graduate course I instructed, I argued that Ava, the teacher whom I studied, was able
at the end of the course to choose tasks in accordance with which aspect of creativity
she wanted to promote. While this is an important skill for a mathematics teacher,
can we say that this action reflects her changing perspectives? Will she implement
such tasks in her 12th grade mathematics class? Certainly, additional research is
needed that integrates teachers’ beliefs regarding creativity, mathematics education,
mathematical creativity, and classroom practices.

177

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

NOTE
1
This section is based on Levenson (2017), presented at the annual PME conference.

REFERENCES
Aljughaiman, A., & Mowerer-Reynolds, E. (2005). Teachers’ conceptions of creativity and creative
students. Journal of Creative Behaviour, 39(1), 17–34.
Ambrose, R. (2004). Initiating change in prospective elementary school teachers’ orientations to
mathematics teaching by building on beliefs. Journal of Mathematics Teacher Education, 7(2),
91–119.
Andiliou, A., & Murphy, P. K. (2010). Examining variations among researchers’ and teachers’
conceptualizations of creativity: A review and synthesis of contemporary research. Educational
Research Review, 5(3), 201–219.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Beghetto, R. A. (2007). Does creativity have a place in classroom discussions? Prospective teachers’
response preferences. Thinking Skills and Creativity, 2(1), 1–9.
Bereczki, E. O., & Kárpáti, A. (2018). Teachers’ beliefs about creativity and its nurture: A systematic
review of the recent research literature. Educational Research Review, 23, 25–56.
Beswick, K. (2007). Teachers’ beliefs that matter in secondary mathematics classrooms. Educational
Studies in Mathematics, 65(1), 95–120.
Beswick, K., Callingham, R., & Watson, J. (2012). The nature and development of middle school
mathematics teachers’ knowledge. Journal of Mathematics Teacher Education, 15(2), 131–157.
Bolden, D., Harries, T., & Newton, D. (2010). Preservice primary teachers’ conceptions of creativity in
mathematics. Educational Studies in Mathematics, 73(2), 143–157.
Chan, D. W., & Chan, L. K. (1999). Implicit theories of creativity: Teachers’ perception of student
characteristics in Hong Kong. Creativity Research Journal, 12(3), 185–195.
Charalambous, C. Y. (2015). Working at the intersection of teacher knowledge, teacher beliefs, and
teaching practice: A multiple-case study. Journal of Mathematics Teacher Education, 18(5), 427–445.
Cogan, L. S., Schmidt, W. H., & Wiley, D. E. (2001). Who takes what math and in which track? Using
TIMSS to characterize US students’ eighth-grade mathematics learning opportunities. Educational
Evaluation and Policy Analysis, 23(4), 323–341.
Cross, D. I. (2009). Alignment, cohesion, and change: Examining mathematics teachers’ belief structures
and their influence on instructional practices. Journal of Mathematics Teacher Education, 12(5),
325–346.
de Souza Fleith, D. (2000). Teacher and student perceptions of creativity in the classroom environment.
Roeper Review, 22(3), 148–153.
Diakidoy, I. A. N., & Kanari, E. (1999). Student teachers’ beliefs about creativity. British Educational
Research Journal, 25(2), 225–243.
Drageset, O. G. (2010). The interplay between the beliefs and the knowledge of mathematics teachers.
Mathematics Teacher Education and Development, 12(1), 30–49.
(PUH$NGR÷DQ (  <D]JDQ6D÷ *   3URVSHFWLYH WHDFKHUV¶ YLHZV RI FUHDWLYLW\ LQ VFKRRO
mathematics. In F. M. Singer, F. Toader, & C. Voica (Eds.), Proceedings of the 9th International
Mathematical Creativity and Giftedness International Conference (pp. 182–187). Sinaia, Romania:
MCG.
Forgasz, H. J., & Leder, G. C. (2008). Beliefs about mathematics and mathematics teaching. In
P. Sullivan & T. Wood (Eds.), International handbook of mathematics teacher education: Knowledge
and beliefs in mathematics teaching and teaching development (Vol. I, pp. 173–192). Rotterdam, The
Netherlands: Sense Publishers.
Frost, J. H. (2010). Looking through the lens of a teacher’s life: The power of prototypical stories in
understanding teachers’ instructional decisions in mathematics. Teaching and Teacher Education,
26(2), 225–233.

178

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

Gralewski, J., & Karwowski, M. (2013). Polite girls and creative boys? Students’ gender moderates
accuracy of teachers’ ratings of creativity. The Journal of Creative Behavior, 47(4), 290–304.
Green, T. F. (1971). The activities of teaching. New York, NY: McGraw-Hill.
Haylock, D. (1997). Recognizing mathematical creativity in schoolchildren. ZDM – The International
Journal on Mathematics Education, 27(2), 68–74.
Kampylis, P., Berki, E., & Saariluoma, P. (2009). In-service and prospective teachers’ conceptions of
creativity. Thinking Skills and Creativity, 4(1), 15–29.
Kandemir, M. A., & Gur, H. (2007). Creativity training in problem solving: A model of creativity in
mathematics teacher education. New Horizons in Education, 55(3), 107–122.
Kaufman, J. C., & Beghetto, R. A. (2009). Beyond big and little: The four C model of creativity. Review
of General Psychology, 13, 1–12.
Kim, K. H. (2009). Cultural influence on creativity: The relationship between Asian culture (Confucianism)
and creativity among Korean educators. The Journal of Creative Behavior, 43(2), 73–93.
Kozbelt, A., Beghetto, R. A., & Runco, M. A. (2010). Theories of creativity. In J. C. Kaufman &
R. J. Sternberg (Eds.), Cambridge handbook of creativity (pp. 20–47). New York, NY: Cambridge
University Press.
Kwon, O. N., Park, J. S., & Park, J. H. (2006). Cultivating divergent thinking in mathematics through an
open-ended approach. Asia Pacific Education Review, 7(1), 51–61.
Lee, K. H. (2017). Convergent and divergent thinking in task modification: A case of Korean prospective
mathematics teachers’ exploration. ZDM-The International Journal on Mathematics Education,
49(7), 995–1008.
Leikin, R. (2009). Exploring mathematical creativity using multiple solution tasks. In R. Leikin,
A. Berman, & B. Koichu (Eds.), Creativity in mathematics and the education of gifted students (pp.
129–135). Rotterdam, The Netherlands: Sense Publishers.
Leikin, R., & Dinur, S. (2007). Teacher flexibility in mathematical discussion. The Journal of
Mathematical Behavior, 26(4), 328–347.
Leikin, R., & Lev, M. (2013). Mathematical creativity in generally gifted and mathematically excelling
adolescents: What makes the difference? ZDM-The International Journal on Mathematics Education,
45(2), 183–197.
Leikin, R., & Levav-Waynberg, A. (2009). Development of teachers’ conceptions through learning
and teaching: The meaning and potential of multiple-solution tasks. Canadian Journal of Science,
Mathematics and Technology Education, 9(4), 203–223.
Leikin, R., Subotnik, R., Pitta-Pantazi, D., Singer, F. M., & Pelczer, I. (2013). Teachers’ views on creativity
in mathematics education: An international survey. ZDM-The International Journal on Mathematics
Education, 45(2), 309–324.
Lev-Zamir, H., & Leikin, R. (2011). Creative mathematics teaching in the eye of the beholder: Focusing
on teachers’ conceptions. Research in Mathematics Education, 13(1), 17–32.
Levenson, E. (2011). Exploring collective mathematical creativity in elementary school. Journal of
Creative Behaviour, 45(3), 215–234.
Levenson, E. (2013). Tasks that may occasion mathematical creativity: Teachers’ choices. Journal of
Mathematics Teacher Education, 16(4), 269–291.
Levenson, E. (2015). Exploring Ava’s developing sense for tasks that may occasion mathematical
creativity. Journal of Mathematics Teacher Education, 18, 1–25.
Levenson, E. (2017). Promoting mathematical creativity in heterogeneous classes. In SEMT ‘17 –
International Symposium Elementary Mathematics Teaching (pp. 42–52). Prague, Czech Republic.
Livne, N. L., & Milgram, R. M. (2006). Academic versus creative abilities in mathematics: Two
components of the same construct? Creativity Research Journal, 18, 199–212.
Morais, M. F., & Azevedo, I. (2011). What is a creative teacher and what is a creative pupil? Perceptions
of teachers. Procedia-Social and Behavioral Sciences, 12, 330–339.
Panaoura, A., & Panaoura, G. (2014). Teachers’ awareness of creativity in mathematical teaching and
their practice. IUMPST: The Journal, 4, 1–11.

179

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
ESTHER S. LEVENSON

3DYORYLü-0DNVLü6 %RGURåD%  ,PSOLFLWLQGLYLGXDOLVPLQWHDFKHUV¶WKHRULHVRIFUHDWLYLW\


Through the “four P’s” looking glass. International Journal of Creativity and Problem Solving, 23(1),
39–57.
Philipp, R. A. (2007). Mathematics teachers’ beliefs and affect. In F. Lester (Ed.), Second handbook
of research on mathematics teaching and learning (pp. 257–315). Reston, VA: National Council of
Teachers of Mathematics.
Pitta-Pantazi, D., Kattou, M., & Christou, C. (2018). Mathematical creativity: Product, person, process
and press. In F. Singer (Ed.), Mathematical creativity and mathematical giftedness. ICME-13
monographs (pp. 27–54). Cham: Springer.
Polly, D., McGee, J. R., Wang, C., Lambert, R. G., Pugalee, D. K., & Johnson, S. (2013). The association
between teachers’ beliefs, enacted practices, and student learning in mathematics. The Mathematics
Educator, 22(2), 11–30.
Rhodes, M. (1961). An analysis of creativity. The Phi Delta Kappan, 42(7), 305–310.
Rubenstein, L. D., McCoach, D. B., & Siegle, D. (2013). Teaching for creativity scales: An instrument to
examine teachers’ perceptions of factors that allow for the teaching of creativity. Creativity Research
Journal, 25(3), 324–334.
Rubenstein, L. D., Ridgley, L. M., Callan, G. L., Karami, S., & Ehlinger, J. (2018). How teachers perceive
factors that influence creativity development: Applying a social cognitive theory perspective. Teaching
and Teacher Education, 70, 100–110.
Runco, M. A., & Acar, S. (2012). Divergent thinking as an indicator of creative potential. Creativity
Research Journal, 24(1), 66–75.
Runco, M. A., & Jaeger, G. J. (2012). The standard definition of creativity. Creativity Research Journal,
24(1), 92–96.
Runco, M. A., & Johnson, D. J. (2002). Parents’ and teachers’ implicit theories of children’s creativity:
A cross-cultural perspective. Creativity Research Journal, 14(3–4), 427–438.
Saracho, O. (2012). Creativity theories and related teachers’ beliefs. Early Child Development and Care,
182(1), 35–44.
Schoenfeld, A. H. (2011). Toward professional development for teachers grounded in a theory of decision
making. ZDM-The International Journal on Mathematics Education, 43(4), 457–469.
Shen, Y., & Edwards, C. P. (2017). Mathematical Creativity for the youngest school children: Kindergarten
to third grade teachers’ interpretations of what it is and how to promote it. The Mathematics Enthusiast,
14(1–3), 325.
Shriki, A. (2010). Working like real mathematicians: Developing prospective teachers’ awareness of
mathematical creativity through generating new concept. Educational Studies in Mathematics, 73,
159–179.
Shriki, A., & Lavy, I. (2012). Teachers’ perceptions of mathematical creativity and its nurture. In
T. Y. Tso (Ed.), Proceedings of the 36th Conference of the International Group for the Psychology of
Mathematics Education (Vol. 4, pp. 91–98). Taipei, Taiwan: PME.
Silver, E. (1997). Fostering creativity through instruction rich in mathematical problem solving and
problem posing. ZDM-The International Journal on Mathematics Education, 3, 75–80.
Silver, E. A., & Cai, J. (2005). Assessing students’ mathematical problem posing. Teaching Children
Mathematics, 12, 129–135.
Siswono, T. Y. E. (2015, May 11–15). Improving elementary teacher competency to develop the abilities
of students’ creative thinking through mathematics problem posing and problem solving strategy.
Paper presented at the 7th ICMI-East Asia Regional Conference on Mathematics Education, Cebu
City, Philippines.
Sriraman, B. (2009). The characteristics of mathematical creativity. ZDM-The International Journal on
Mathematics Education, 41, 13–27.
Swars, S. L., Smith, S. Z., Smith, M. E., & Hart, L. C. (2009). A longitudinal study of effects of a
developmental teacher preparation program on elementary prospective teachers’ mathematics beliefs.
Journal of Mathematics Teacher Education, 12(1), 47–66.

180

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICAL CREATIVITY IN THE CLASSROOM

Watson, A., & Sullivan, P. (2008). Teachers learning about tasks and lessons. In D. Tirosh & T. Wood
(Eds.), Tools and resources in mathematics teacher education (pp. 109–135). Rotterdam, The
Netherlands: Sense Publishers.
Zehetmeier, S. (2015). Sustaining and scaling up the impact of professional development programmes.
ZDM-The International Journal on Mathematics Education, 47(1), 117–128.

Esther S. Levenson
Department of Mathematics, Science and Technology Education
Tel Aviv University

181

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PART 3
THE INTERPLAY OF MATHEMATICS TEACHER
IDENTITY, BELIEFS AND KNOWLEDGE

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

7. BELIEFS AND PEDAGOGICAL CONTENT


KNOWLEDGE FOR TEACHERS OF MATHEMATICS

In this chapter we discuss methodological and theoretical implications of considering


pedagogical content knowledge for teaching mathematics as inclusive of teachers’
relevant beliefs together. We present a refined framework that encompasses
pedagogical content knowledge and report its use with two focus groups of
experienced teachers of mathematics, with particular attention to connections
between their knowledge and beliefs. We believe our approach yielded a more
nuanced understanding of teachers’ knowledge and of the complexity of teaching
than a consideration of just knowledge (traditionally conceived) or beliefs alone.
In particular, it highlighted the importance of teachers’ beliefs about themselves as
mathematicians and teachers of mathematics, and about their students’ capacities
as mathematics learners and their affective reactions and motivations in relation to
the subject.

INTRODUCTION

In this chapter we use a slightly modified version of the framework for mathematics
teachers’ pedagogical content knowledge of Dyment, Chick, Walker and Macqueen
(2018) to illustrate the interaction of beliefs and knowledge, and the complexity
of teachers’ thinking about classroom scenarios. We draw on data from two focus
group discussions involving experienced secondary mathematics teachers and
mathematics education researchers. The discussions were stimulated by items
developed as described by Beswick and Callingham (2011) that included examples
of student responses to mathematical tasks, the selection of appropriate examples
to illustrate particular mathematics concepts, and knowledge of representations of
mathematical concepts. The pedagogical content knowledge framework of Dyment
et al. (2018) was based on an earlier version (Chick, 2007) to which knowledge of
assessment, beliefs about the nature of the content (mathematics), and knowledge
of student affect were added by Chick and Beswick (2013, 2018). Dyment et al.
(2018) added a further beliefs aspect that is retained in this chapter. In this analysis
of the discussion of one of the focus group items we highlight aspects of pedagogical
content knowledge that were evident and the complex interplay between beliefs and
other aspects of pedagogical content knowledge and discuss the implications for
teacher education.

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_008

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

There has been ongoing discussion about the nature of beliefs compared
with knowledge, with some researchers arguing that the two constructs are
indistinguishable for practical purposes (Beswick, 2005, 2007; Liljedahl, 2008).
That is, within their classrooms, teachers act as if they know things that others
might consider to be beliefs. This fact, and recognition of the powerful, albeit
subtle, influence of teachers’ beliefs on their practice, make it sensible to consider
knowledge and beliefs together. In this chapter we begin by defining knowledge
and beliefs before discussing pedagogical content knowledge; the knowledge that
distinguishes teachers of mathematics from teachers in general. We then discuss
connections between beliefs and pedagogical content knowledge and present our
study and findings from it. The discussion focuses on methodological and theoretical
implications of the study.

KNOWLEDGE AND BELIEFS

Whereas beliefs research is often criticised for focussing on an ill-defined


construct (Leder, 2015), the meaning of knowledge is typically taken for granted,
with discussion of its structure a common entry point in mathematics education
research (e.g., Blömeke & Kaiser, 2014). Any discussion of the interplay of
knowledge (specifically pedagogical content knowledge in this chapter) and beliefs,
however, depends upon an understanding of how these constructs are defined and
distinguished. Distinctions between knowledge and beliefs are based upon notions
of truth, certainty, and justification. This is illustrated by Philipp’s (2007) adoption
of Richardson’s (1996, p. 106) definition of beliefs (also used by Blömeke & Kaiser,
2014) as “psychologically held understandings, premises, or propositions about
the world that are felt to be true” and his definition of knowledge as a subset of
beliefs, namely those “held with certainty or justified true belief” (Philip, 2007, p.
259). As explained by Beswick (2011), truth, certainty and justification are linked
because justification must be based on some accepted criteria for establishing
truth, and the extent to which such criteria are satisfied informs the certainty with
which an understanding, premise, or proposition is held to be true. What counts as
acceptable warrants for the truth of a claim, and hence of its status as knowledge,
is not straightforward but varies with aspects of context including the nature of the
discipline (for example the veracity of mathematical propositions is established
according to different criteria from that of historical claims), audience, and culture.
Cilliers (2005) pointed to two broad views of the nature of knowledge: on the one
hand that objective truth exists and can, in principle at least, be uncovered, and on
the other hand that knowledge is culturally defined and hence necessarily contestable
and subjective. The latter is essentially a constructivist view. Lerman (1989) argued
that, rather than rendering the pursuit of knowledge pointless, constructivism makes
it even more important that claims are questioned and that preferences for one
idea over another are carefully justified. From a constructivist perspective, beliefs
are a type of knowledge that attract relatively less consensus as a consequence of

186

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

being based upon less and/or poorer information and being less powerful in terms
of making sense of the world (Guba & Lincoln, 1989). Conversely, and following
Beswick (2005, 2007) and Beswick, Callingham, and Watson (2012), we understand
beliefs essentially as defined by Ajzen and Fishbein (1980) to be anything that an
individual regards as true and, therefore, to encompass knowledge.
Current work on teachers’ knowledge, including that reported in this chapter,
draws upon the seminal work of Shulman (1987) but it is useful to consider “types
of knowledge” more broadly and go further back in time than Shulman. In 1949
the philosopher Gilbert Ryle distinguished between “knowing-that” (factual or
propositional knowledge) and “knowing-how” (practical or enacted knowledge),
although there have since been some debates about the distinctions and relationship
between the two (see Fantl, 2008, for some discussion about this). These distinctions
seem useful when talking about knowledge for teaching: it is one thing to “know-
that” Pythagoras’ theorem is taught in Grade 9, and a quite different thing to “know-
how” to teach it effectively. Mason and Spence (1999) suggested that by 1994
mathematics educators were also talking about “knowing-why” (knowledge of what
explains our knowledge; this type of knowledge was also posited by Fantl). Mason
and Spence were talking about content knowledge to be learned by students, but
the concept of “knowing-why” can, we suggest, be regarded as a third knowledge
type more broadly, and thus might apply to the knowledge held by teachers for
teaching. So, for example, “knowing-why” for teaching could include “knowing-
why” students might have difficulty with the formalities of recording a syntactically
valid solution to a problem involving Pythagoras’ theorem. Such knowledge may
draw on “knowing-that” and “knowing-how.”
Mason and Spence then added “knowing-to” as a fourth knowledge type
(knowledge of how to respond “in-the-moment” to a set of circumstances, such as to
“know-to” use a particular solution method in an unfamiliar problem type), to capture
the idea of being able to spontaneously recognise that a method or approach might
be useful in a new situation and to “know-to” try it. Again, it seems feasible, and
within Mason and Spence’s intent, to extend this idea of “knowing-to” to knowledge
generally, not just mathematics content knowledge, reflecting the observation that
sometimes individuals may “know-that,” or “know-how,” or “know-why,” but be
unable to “know-to” use that knowledge when required. In particular, we suggest
that this knowledge type also applies to teacher knowledge, in a very important
way. As an example, a teacher might “know-to” prompt a group of students, based
on awareness of their readiness and a textbook question they had just done, to
generalise Pythagoras’ theorem to three dimensions. This fourth type of knowledge
captures the more dynamic actions suggested by the Knowledge Quartet (Rowland,
Huckstep, & Thwaites, 2005), particularly contingency.
These categories of knowledge might also be seen to vary in the degree to which
they might be contentious and hence likely to be categorised as beliefs. Although
knowing-that is described as relating to factual (implying correctness) knowledge,
not all facts are universally agreed. A teacher, for example, might “know-that” Johnny

187

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

cannot learn sophisticated mathematics. How such a teacher approaches teaching


Johnny (“know-how”) is likely also to be controversial. Even if individuals agree
that something is so (“know-that”) or is done in a certain way (“know-how”) the
reasons they might offer could vary, perhaps because of other relevant propositional
(“knowing-that”) or procedural knowledge (“knowing-how”). In relation to Johnny,
for example, the teacher might also “know-that” Johnny seems very like his older
sister who was unsuccessful in mathematics, and thus “know-how,” from experience,
to teach such students. Another teacher could make similar assessments of Johnny
and teach him similarly but for different reasons, or they might teach him differently
despite “knowing-that” Johnny’s capacity to learn mathematics is limited; that is,
their “knowing-why” could be different. Knowledge in the “knowing-to” category
is also subjective. Even teachers aware in the moment of the same response options
may “know-to” do quite different things. Again, such choices would depend upon
the rest of the teachers’ knowledge including across the other three forms. All of
this sounds rather like discussions of teachers’ beliefs viewed as systems in which
teachers’ actions are not traceable in a linear fashion to beliefs, but rather beliefs
operate together with some taking a more influential role than others as the context
varies (Beswick, 2018).

PEDAGOGICAL CONTENT KNOWLEDGE

Concerns about identifying the knowledge that is needed for the teaching of
mathematics have existed for decades, motivated by the observation that content
knowledge alone did not seem to be a good predictor of successful teaching.
Shulman famously brought these concerns to a head in the middle of the 1980s,
when he identified and characterised a number of categories of teacher knowledge,
including pedagogical content knowledge which he saw as the “category most likely
to distinguish the understanding of the content specialist from that of the pedagogue”
(Shulman, 1987, p. 8). With a view to teaching the content of a particular subject
area, he included within pedagogical content knowledge
the most useful forms of representation of those [content] ideas, the most
powerful analogies, illustrations, examples, explanations, and demonstrations
– in a word, the ways of representing and formulating the subject that make
it comprehensible to others … [It] also includes an understanding of what
makes the learning of specific concepts easy or difficult: the conceptions and
preconceptions that students of different ages and backgrounds bring with
them to the learning. (Shulman, 1986, p. 9)
Since then there has been an explosion of research into knowledge for teaching,
not only within mathematics, but in other areas as well (see, for example, Loughran,
Berry, and Mulhall, (2012) for science education, Pitfield (2012) for drama education,
and Dyment et al. (2018) for outdoor education). In mathematics, the scope and focus
of these studies – and the “types” of knowledge being investigated – have varied

188

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

hugely, and a number of “knowledge frameworks” have arisen from these. We will
review four, one of which is the focus of this chapter and three which are arguably
most prominent in the mathematics education literature, to highlight what they offer
for research into pedagogical content knowledge. Together the frameworks represent
a range of approaches to conceptualising mathematics teachers’ knowledge in terms
of grain size and the extent to which they represent knowledge as relatively static
or dynamic. In each of these studies, knowledge has been regarded as something
that teachers possess and may acquire, in contrast to discursive approaches such as
suggested by Barwell (2013).
The “mathematical knowledge for teaching” framework (Ball, Thames, & Phelps,
2008; Hill, Ball, & Schilling, 2008) has, as its name suggests, a focus on subject
matter knowledge, but with an emphasis on what is germane for teaching. In its
pedagogical content knowledge component, it includes knowledge of content and
students, knowledge of content and teaching, and knowledge of curriculum, and
to this is added a “subject matter knowledge” component focussing on content
knowledge that teachers specifically need for their teaching work. It is worth noting
that this framework incorporates within pedagogical content knowledge both
knowledge of curriculum and knowledge of learners, which Shulman had identified
as categories of knowledge separate from pedagogical content knowledge. These
two categories are, however, fundamentally associated with the way that content
and pedagogy come together in sequencing content in a way that meets learners’
needs, and so a case can be made for including them as part of pedagogical content
knowledge.
The framework of Krauss et al. (2008) has content knowledge as its foundation
and, growing out of this, identifies three critical areas of pedagogical content
knowledge: knowledge of students’ conceptions and prior knowledge, knowledge of
the potential and implementation of mathematical tasks, and knowledge of subject-
specific instruction strategies. In common with Ball and colleagues, Krauss et al.
seem to view knowledge for teaching as a set of understandings that may or may
not be in the possession of a teacher, suggesting that pedagogical content knowledge
is static at any given point in time, and measurable. Although attempts have been
made to measure pedagogical content knowledge by both teams, it can be argued
that trying to do this efficiently on a large scale, perhaps by means of a written test,
cannot fully reveal what pedagogical content knowledge a teacher brings to bear in
the complex milieu of the classroom (see Chick, 2011, for further discussion; this
chapter will highlight additional complexities).
The “Knowledge Quartet” framework (Rowland et al., 2005; Rowland, this
volume of the handbook) sees knowledge for teaching somewhat more dynamically
than the previous two frameworks. A comprehensive set of categories contribute to
the Knowledge Quartet’s four components. The foundation component incorporates
possessed content knowledge, and so is rather static in nature, but the other three
components highlight actions associated with using that knowledge in the classroom.
The transformation component concerns transforming content knowledge into

189

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

forms for effective teaching, connection considers links and coherence among
topics and lessons, and contingency reflects the capacity to respond to unexpected
events. Nevertheless, these three more dynamic components of knowledge are
clearly dependent on held knowledge, both content knowledge and pedagogical
content knowledge, since the ability to transform knowledge, for example, requires
an understanding of the content and what representations might be effective for a
particular group of learners.
This brings us to the framework for pedagogical content knowledge that is the
focus of this chapter (Dyment et al., 2018). It dates back to Chick, Baker, Pham, and
Cheng (2006) and, more completely, to Chick (2007; see also Chick & Beswick,
2013 that first included beliefs and affect related categories). This framework
captures the complexity of pedagogical content knowledge as a blend of pedagogical
knowledge and content knowledge (Shulman, 1987), identifying a number of
categories loosely divided into three groups depending on the predominance of
pedagogical or content knowledge in the mix. It unpacks in more detail what
the amalgam of pedagogical and content knowledge that Shulman (1987) called
pedagogical content knowledge looks like for teachers of mathematics. In particular,
our three categories illustrate the differing emphases on pedagogical and content
knowledge that constitute different aspects of pedagogical content knowledge and
make the important point that aspects of pedagogical or content knowledge may be
more or less prominent in the pedagogical content knowledge that informs teachers’
decision making in relation to specific aspects of mathematics teaching. Table 7.1
shows the framework, with descriptors for each category, and illustrative examples.
The categories are more fine-grained than most other frameworks (although the
Knowledge Quartet of Rowland et al. (2005) is built on a not-dissimilar extensive
set of components) and are not intended to be hierarchical or discrete (overlap
among categories is acknowledged). It is also underpinned by a very broad view of
the scope of pedagogical content knowledge, including curriculum knowledge and
aspects of content knowledge.
Not surprisingly there are commonalities between this framework and the others
that have been discussed; the categories are about “know-how,” “know-that,” and
“know-why,” and so can be regarded as having a static, “held” knowledge view
of pedagogical content knowledge. However, the descriptors acknowledge that this
knowledge is evident in the use of it, so “knowing-to” is, perhaps, implicit. In practice,
the framework offers a set of well-defined “filters” through which teaching can be
viewed in order to identify evidence for teachers’ pedagogical content knowledge. It
has been applied to interview and questionnaire data as well as to classroom practice
(e.g., Chick, 2007), with evidence for the existence of all the categories, and with
situations in which teacher knowledge seemed evident in a teacher’s explanation
or actions being able to be described using one or more of the categories. The
framework has also been adapted to look at the pedagogical content knowledge
needed by mathematics teacher educators; see Chick and Beswick (2018).

190

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Table 7.1. Framework for Pedagogical Content Knowledge (PCK)


for teaching school mathematics

PCK category PCK for School Mathematics Teachers (SMTPCK)


Evident when the teacher … Example

Clearly PCK
Beliefs about the Discusses or uses personal or Discusses the use of pedagogies
Nature of Teaching established philosophies and consistent with a constructivist
and Learning approaches to teaching the view of learning as a key facet of
discipline teaching mathematics; requires
students to memorise times tables
because of a belief that students
should have fast recall of key
number facts
Teaching Strategies Discusses or uses general or Uses concrete materials to
specific strategies or approaches demonstrate a concept
for teaching a mathematical
concept or skill
Student Thinking Discusses or addresses student Identifies that a student doesn’t
ways of thinking about a recognise the equivalence of
concept, or recognises typical equivalent fractions
levels of understanding
Student Thinking – Discusses or addresses student Recognises that students often
Misconceptions misconceptions about a concept think “multiplying makes bigger”
Student Affect (in Discusses or addresses students’ Recognises that adolescent
relation to content) affective responses to particular students may have negative
mathematics topics emotional reactions to the
prospect of learning algebra
Cognitive Demand of Identifies aspects of the math Recognises 627 – 359 is more
Task task that affect its complexity difficult to model than 687 – 321
Representations of Describes or demonstrates Uses Multi-base Arithmetic
Concepts ways to model or illustrate a Blocks to model subtraction
concept (can include materials
or diagrams)
Explanations Explains a topic, concept or Explains why we can write a 0 on
procedure the end of a whole number when
multiplying by 10
Knowledge of Uses an example that highlights Uses the 5-12-13 Pythagorean
Examples a concept or procedure triads to model how to solve a
right-angled triangle problem
Knowledge of Discusses/uses resources Identifies and uses a mathematics
Resources available to support teaching website that is useful for students

(cont.)

191

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

Table 7.1. Framework for Pedagogical Content Knowledge (PCK)


for teaching school mathematics (cont.)

PCK category PCK for School Mathematics Teachers (SMTPCK)


Evident when the teacher … Example

Curriculum Discusses how topics fit into Recognises that multiplication


Knowledge the curriculum should be understood by Year 4
Purpose of Content Discusses reasons for Knows that knowledge of
Knowledge content being included in the rounding is needed for money
curriculum or how it might be transactions
used

Content Knowledge in a Pedagogical Context


(Beliefs about) The Expresses an appreciation of the Compares the aesthetic qualities
Nature of Content nature of mathematics that goes of two solution methods;
beyond the school curriculum
and aligns with mathematicians’
view of the discipline
Profound Exhibits deep and thorough Understands why we “invert and
Understanding of conceptual understanding multiply” when dividing fractions
Fundamental Content of identified aspects of
mathematics
Deconstructing Identifies critical mathematical Refers to the importance of
Content to Key components within a concept the distributive law in the long
Components that are fundamental for multiplication algorithm
understanding and applying that
concept
Structure and Makes connections between Links percentages with decimals
Connections mathematical concepts and the base 10 system
and topics, including
interdependence of concepts
Procedural Displays skills for solving Can apply the long division
Knowledge mathematical problems algorithm
(conceptual understanding need
not be evident)
Methods of Solution Demonstrates a method for Adds fractions using both
solving a mathematical problem an algorithm and a visual
representation

Pedagogical Knowledge in a Content Context


Assessment Discusses or designs tasks, Designs a multiple-choice quiz
Approaches activities or interactions that with appropriate distractors
assess learning outcomes

(cont.)

192

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Table 7.1. Framework for Pedagogical Content Knowledge (PCK)


for teaching school mathematics (cont.)

PCK category PCK for School Mathematics Teachers (SMTPCK)


Evident when the teacher … Example

Goals for Learning Describes a goal for students’ Justifies an activity as developing
learning understanding of long term
probability
Getting and Discusses or uses strategies for Designs a puzzle that is solved by
Maintaining Student engaging students answering some routine exercises
Focus
Classroom Techniques Discusses or uses generic Talks about grouping students
classroom practices according to ability levels; sets up
a classroom atmosphere in which
students feel safe discussing their
mathematical solutions
Student Affect Describes how student affect Knows a particular student will
(general) influences pedagogical respond to negatively to being
approach asked for an answer in a large
group session

Sources: Chick (2007) (near-complete presentation of pedagogical content knowledge for


mathematics teaching); Chick, Baker, Pham, and Cheng (2006) (earlier partial presentation
of pedagogical content knowledge framework); Baker (2008) (some illustrative examples
of pedagogical content knowledge); Chick and Beswick (2013, 2018) (presentation of a
pedagogical content knowledge framework specifically for mathematics teacher educators,
including some additional pedagogical content knowledge categories); Dyment et al. (2018)

BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Unlike other teacher knowledge frameworks, the pedagogical content knowledge


framework for teaching school mathematics presented in Table 7.1 explicitly
acknowledges the role of beliefs at the interaction of pedagogical activity and the
content to be taught. Some of this is straightforward; not only would we expect a
pedagogical content knowledge framework to include knowledge of student thinking
about content, but knowledge of students’ affective reactions to content as well as
their affect more generally also should be essential components (Chick & Beswick,
2018). In addition, Chick and Beswick argued that teachers’ beliefs about the nature
of content (i.e., mathematics) also warranted inclusion because understandings of the
nature of the discipline and hence what it means to learn, know, and do mathematics
can be seen to influence pedagogical decisions and activities in the same way as
more conventional “knowledge.” Dyment et al. (2018) added the further category of
beliefs about teaching and learning, arguing that these also impact teachers’ decision
making. As an example, consider two teachers who both identify the Cognitive

193

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

demand of a task, but who respond to this differently in the classroom. One uses
Procedural knowledge and Explanations as a Teaching strategy to provide students
with a correct step-by-step modelled solution; the other uses the Teaching strategy of
allowing students time to grapple with the challenge (Maher, 2018). The reasons for
these different decisions are influenced by beliefs, knowledge of research results, and
acknowledgement of contextual factors that may contribute to the decisions, although
it is conceivable that each teacher would claim to “know” that they have made a better
choice than the other. The inclusion of aspects of teachers’ beliefs in a knowledge
framework is consistent with arguments that beliefs and knowledge are essentially
equivalent, being distinguishable only by the extent of the consensus that they attract.
One arguably helpful consequence of including beliefs in conceptualisations of
pedagogical content knowledge is that it highlights the contentious nature of many
aspects of what is traditionally and uncontroversially called knowledge. For example,
the proponents of each of the pedagogical content knowledge frameworks reviewed
in this chapter would likely agree that they have identified aspects of knowledge
upon which teachers draw but that the frameworks do not distinguish knowledge
that would be deemed “correct” from that which would be widely considered
“incorrect.” Nevertheless, there is an implicit assumption of correctness in the use of
the term knowledge as illustrated by the following hypothetical example. Consider a
teacher who has “knowledge” that “squares are not rectangles” as part of the content
knowledge that informs his teaching. In his teaching about squares and rectangles
he would be drawing upon the relevant parts of his Profound understanding
of fundamental content and/or his knowledge of mathematical Structure and
connections even though it could be argued that this knowledge is incorrect. We
would tend to call such erroneous knowledge a misconception. The categories of
pedagogical content knowledge in Table 7.1 are, of course interconnected and so
the teacher described may also “know” things about, for example, the Cognitive
demand of a task that involves categorising quadrilaterals, the most appropriate
Representations and Explanations of related content, and what constitutes useful
Examples and Resources. This knowledge may be consistent with his content
knowledge and indeed, if it is, would likely be deemed unhelpful and inappropriate
(if not incorrect). We would tend to characterise at least this part of the teacher’s
knowledge as beliefs because it does not align with normative understandings of
the truth about the relationship between squares and rectangles. This is, of course,
precisely the distinction between knowledge and beliefs that Beswick et al. (2012)
and others have argued. Taking this argument to its logical conclusion would mean
that each of the knowledge categories in Table 7.1 would have an analogous beliefs
category (e.g., Knowledge of student thinking and Beliefs about student thinking)
and that we could also talk about pedagogical content beliefs. We are certainly not
arguing that this would be helpful, only that knowledge has an inherent assumption
of correctness within some context and with some imagined audience.
The beliefs categories in the pedagogical content knowledge framework shown
in Table 7.1 – Beliefs about teaching and learning and Beliefs about the nature

194

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

of content – might be argued to be necessary because they pertain to aspects of


knowledge about which there is not, at least in the mathematics education circles in
which the authors move, a single uncontroversial view that a teacher might “know.”
Rather, a range of beliefs about these matters could underpin teaching that is deemed
effective by one or another widely but not universally accepted criterion. Others may
have differing views about which elements of the framework are of that kind and
about what the accepted correct knowledge in a particular category is. Knowledge
frameworks are necessarily contextual in relation to culture, place, and time so it is
entirely appropriate that the current framework (Table 7.1) has evolved over several
iterations as it was tested in differing contexts and for different purposes.
Further evidence of the entailment of beliefs and knowledge is in the difficulty
of defining some of the categories. For example, knowledge of Student thinking, as
defined in Table 7.1, is about the thinking of individual students as well as typical
development of students’ thinking about mathematical concepts. The knowledge in
this category upon which teachers draw is inevitably coloured by their “knowledge”
of particular students and groups of students. We know, for example, that teachers’
beliefs about students’ ability to reason and solve problems are influenced by
their perceptions of whether the student is or is not mathematically capable and
that the latter judgement can be based upon such things as the students’ ability to
memorise basic facts, and behavioural and affective characteristics (e.g., confidence,
motivation) that may not be related to mathematical capability (Beswick, 2017).
Similarly, teachers’ judgements about appropriate Representations of concepts,
Examples, and Explanations are all influenced by their beliefs about the capability
of students, with those deemed less capable experiencing narrower, less challenging
curricula (Beswick, 2007/2008). In effect, teachers may have knowledge about
what the most effective way of explaining a particular concept is, but they may
believe something different about the best way to explain that concept to a particular
type of student. In this last sentence “knowledge” is used to refer to ideas that are
relatively broadly accepted within the teacher’s context (i.e., colleagues are likely to
agree, it would be consistent with messages that might be received in professional
development sessions), and “beliefs” to refer to ideas that might be more controversial
in that context.

THE STUDY

The data used in this chapter are drawn from parts of two focus group discussions
conducted as part of a larger project exploring the knowledge of teachers for teaching
mathematics and English.

Participants

Participants were two groups of secondary school teachers and mathematics


education researchers, one in the Australian state of Tasmania and the other in New

195

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

Zealand. The Tasmanian group comprised two researchers (the authors), two other
mathematics educators, and three teachers. The New Zealand group comprised two
researchers and four teachers. In both cases, teacher participants were experienced
secondary school mathematics teachers purposively selected because they were
known by one or more of the researchers to be interested in examining mathematics
teaching and were reputed to have high levels of relevant expertise.

Group Interview Schedule and Procedure

Discussion was stimulated by the successive presentation of items relating to a range


of content areas, year levels, and pedagogical content knowledge issues. Several of
the items, including the one that is the focus of this chapter, were drawn from an
earlier study and developed as described by Beswick and Callingham (2011). This
chapter draws upon the discussions of the item shown in Figure 7.1. It has been
used with groups of both primary and secondary teachers because although fraction
division appears in the Australian Curriculum: Mathematics (Australian Curriculum
Assessment and Reporting Authority, 2018) in Year 7, the first year of secondary
school, the underpinning concepts are important across the upper primary years as
well. Chick (2015) discussed the responses of groups of primary school teachers to
this item in the context of considering the teachers’ use of language in relation to
content as an aspect of pedagogical content knowledge.

Figure 7.1. The division of fractions item

When Figure 7.1 was presented to a group, the participants were invited to discuss
the student’s work and how a teacher might respond. The discussion was unstructured,
although the researchers helped to keep the focus on the scenario. The intent of the
discussions was not to evaluate the knowledge of any or all of the teachers or mathematics
educators but rather to explore the nature of teachers’ knowledge for teaching secondary
school mathematics. The researchers, therefore, participated in the discussion and
thereby contributed to the data. The discussions were audio-recorded, transcribed, and
coded according to the pedagogical content knowledge framework shown in Figure 7.1.

196

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

RESULTS

Across the transcripts from the Tasmanian group and the New Zealand group we
identified evidence of each of the categories in the framework except for Getting and
maintaining student focus, which is unsurprising, given that this is more likely to be
evidenced in teacher-directed class activity, than in the scenario as described. The
New Zealand group spent longer discussing this item than did the Tasmanian group
and so more of the data presented in this section are from that group. As expected,
given the item’s focus, the knowledge evidenced was weighted toward the clearly
pedagogical content knowledge aspects Teaching strategies and Student thinking,
and the aspects of content in a pedagogical context, Profound understanding of
fundamental content, Structure and connections, Methods of solution, and Procedural
knowledge. While most of the knowledge that was evident was consistent with the
experience and knowledge status of the participants, we also found evidence of
“incorrect” knowledge among our participants that we might conceive of as beliefs
and somewhat controversial knowledge (beliefs) that do not readily fit with the
framework as shown in Table 7.1. We begin with data concerning the beliefs and
affect related aspects of our pedagogical content knowledge framework (Table 7.1).
Participants are identified as belonging to the Tasmanian (TAS) or New Zealand
(NZ) group, and then as a researcher (R) or teacher (T).
Aspects of pedagogical content knowledge that were not included in versions of
the framework prior to Dyment et al. (2018), include knowledge of student affect
(Student affect (in relation to content), and Student affect (general)), Beliefs about the
nature of content, and Beliefs about nature of teaching and learning. The transcripts
included evidence of each of these. We found no explicit reference to Beliefs about
the nature of teaching and learning, possibly because of the high degree of taken-
for-granted understanding in the two groups about these broad ideas and/or because
the stimulus item focussed participants’ attention on a very specific example rather
than on broad approaches to teaching and learning mathematics.
Statements that showed evidence of knowledge of Student affect (in relation to
content) mainly focussed on students’ motivation. For example, one participant
described a student as, “just too lazy to put in any working” (NZ R1), which led to
a discussion of the ubiquity of such students and of the “cost”/”pay-off” calculation
that some students appear to make in relation to the work they are prepared to do.
Later in the same conversation a teacher described how sometimes students need to
“trust the teacher” (NZ T1) when he/she asks them to use a procedure that is longer
and slower for initial examples but that will be more widely useful and robust in
the longer term. One teacher referred to the tension between developing an aspect
of students’ affect, namely perseverance, by allowing them to continue with a “big,
long, complicated way to do it” (NZ T1) or showing them the short way. NZ T3
commented on the potential for the solution method shown in the item to “challenge
their thinking” (NZ T3) with the latter observation leading to the conclusion “I don’t
know that I would show them this method” (NZ T3).

197

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

Other comments that appeared to draw upon knowledge of student affect seemed
to be of a more general nature, although the distinction between these comments
and those related to student affect in relation to content was largely inferred from
the context of the remarks. For example, the New Zealand discussion included
an exchange about students explaining mathematical content to one another that
appeared to turn to the arguably broader issue of students being overly reliant on
their peers rather than thinking for themselves. NZ T4 explained that he had needed
to move students away from a group and tell them that they needed to do their own
thinking. He added that “they initially got a wee bit miffed with that” (NZ T4).
A researcher explained that both the helping and helped students gain from their
interaction. She said, “they get to feel good if they help and the others feel good if
they get the answer” (NZ R1). Reference was also made to some students liking “to
investigate” (NZ T4) and doing more work than was required because “he just liked
doing it” (NZ T4).
Comments that evidenced teachers’ beliefs about the nature of the content
were of two kinds: relating to the aesthetic quality of solutions and to the nature
of mathematics more generally. In the first category was NZ T1’s use of the term
“elegant” to describe solution methods that are “quite short.” She appeared to be
contrasting the solution method depicted in the item with the standard fraction
division algorithm. NZ R2 responded with an alternative criterion for judging a
solution to be elegant. He said:
You talked about elegance of a particular method … do you react to this [the
item] as being less elegant than another procedure for dividing fractions or is it
actually showing, is it more elegant in showing a higher level of understanding.
There were two ways in which teachers used affective language to describe the
specific solution shown in Figure 7.1. NZ T4 described it as “a wonderful answer”
but NZ T2 believed that with inconvenient numbers the method “gets very nasty.”
NZ T3 agreed that it “will eventually get nasty.”
An exchange involving NZ T4 and NZ R1 focussed on the need to take time to
think, either about problems of teaching mathematics, or about solving mathematical
problems not necessarily in the context of teaching. The teacher began by responding
to a question from NZ R2 about how one could respond in a classroom to a novel
solution method by saying:
… you can say ‘I need time to think about this’ … the students have to
understand that some of these things can take time, people want to go away
and work with them … so that you don’t always have instant answers. And you
might say, ‘well that’s a really interesting method’ … If they can’t explain it …
or you just want time to come back the next day and have a look at it … It’s the
understanding that we can take a day or two to work on something. (NZ T4)
NZ R2 agreed saying “I think that’s really important. It’s not just about getting
the answer snap, snap, snap.” Although she expressed agreement, NZ R2 had in fact

198

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

changed the focus from highlighting the need to take time to think about unexpected
student solutions to making a more general comment about the way that mathematics
is done. This shift was not commented upon by any of the participants. Also, at this
broader level concerning the nature of mathematics, a Tasmanian teacher implicitly
expressed a belief that mathematics is not about performing calculations. He said
that he was disappointed with the extent of the emphasis in the curriculum on
fraction operations because, “It’s not that valuable, that knowledge any more unless
you’re manipulating things algebraically, you know, there’s technology and other
strategies” (TAS T2).
There were relatively few statements that contained some evidence of knowledge
for teaching mathematics that could not be readily assigned to one of the categories.
Among these were references to students’ ability, often in connection with other
aspects of pedagogical content knowledge. For example, TAS T1 said “I thought
‘Oh! Clever kid!’” and went on to characterise the solution as “clever” because
it drew upon knowledge of finding a common denominator, a process with which
students would be familiar from primary school in the context of fraction addition.
She appeared to be basing her assessment of the student who produced the solution
on her own Procedural knowledge and Curriculum knowledge. In the New Zealand
discussion there was broad agreement that there could be value in showing the
solution to mathematically capable students such as “a top Year 9 [class]” (NZ
T3), “sort of brighter kids” (NZ R1), or “a Year 12 class” (NZ T4), reflecting an
assessment of the Cognitive demand of a task focussed on understanding the solution
method. It could be that such a task was deemed cognitively demanding because the
participants were themselves challenged by the solution method, and particularly by
whether or not it was mathematically legitimate. At the start of the discussion, for
example, three of the four New Zealand teachers appeared somewhat dismissive of
the solution method, as evidenced by the following exchange:
I’d like to hear from the students why they have changed it to eight over twelve
divided by nine over twelve. (NZ T2)
Because that’s what you do when you add. (NZ T3)
Yeah exactly … and I’d then ask them ‘well are we adding fractions now?’
(NZ T2)
Yeah, they’ve got at least two other process … apart from the one that they
don’t need. They’ve got the adding fractions process and they’ve got some
cross multiplying going on there. (NZ T1)
It was at this point that NZ T4 described the solution as “wonderful” and answered
NZ R2’s subsequent question about whether the answer was right, with “Yes.
Absolutely right. That is a perfectly valid way to do it.” NZ T3 went on to recount
having had her attention drawn, by a researcher with whom she was working at the
time, to the fact that a student had used the same method in her own class. She said:

199

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

One of the students used this method and I never picked up on it as a teacher
but <researcher name> picked up on it when she was interviewing the students
and showed it to me … I can’t remember whether the student actually had the
understanding or whether they just, but they were definitely using their prior
knowledge of adding fractions … I did a little bit more work on it myself … it
is a valid method.
Hereafter the tenor of the conversation changed and, later in the discussion, NZ T1
described how an inexperienced teacher might see such a solution and immediately
conclude “Right answer, wrong working’ and just dismiss it. Whereas it’s made us
think, hasn’t it”? NZ T3 then confessed, “when I first looked at it I said, ‘They’ve
used wrong … you know they’ve taken knowledge and used it incorrectly.” It seems
likely that the social context of the group contributed to a tendency to agree with
one another and a reticence to express a contrary view. NZ T3 was able to agree
with the correctness and generality of the method (asserted by NZ T4) because she
had prior experience with it, had had time to consider it, and had had its validity
confirmed in prior conversations with at least one other “knowledgeable” individual
– the researcher with whom she had worked.
The initial uncertainty about the student’s solution that was described above in
relation to the New Zealand group was evident in both groups. In both there was also
tacit agreement that the value obtained was correct, suggesting that the participants
had the content knowledge – likely Procedural knowledge – to obtain the answer for
themselves via some process. Their uncertainty about the correctness of the solution
appeared to stem from the fact that the written solution did not align with the usual
algorithm. Instead, they acknowledged that it resembled known algorithms for other
fraction operations; as one participant said, “it looks to me like they’re confusing
it with the addition algorithm” (TAS T1) and NZ T1 commented on the multiple
processes that appeared to be involved.
This uncertainty – reflecting the fact that all but one of the teachers did not
immediately seem to know if the solution was valid either for this specific case
or as a more general algorithm – created a contingent moment in the interview.
The participants had been invited to discuss what they might do in response to
the student’s work but seemed not yet to “know-that” it was valid (or otherwise),
and so, obviously, did not yet “know-why.” What this moment provided was an
opportunity to reveal in what way they might “know-how” to respond, or what they
would “know-to” do in response to the uncertainty. For some of the teachers the
first response they suggested they would make was to ask the student to explain the
solution. NZ T2 said:
When I first looked at it, my first thought was, I’d probably ask them why they
put the twelve, and why-, I know why they have, but yeah. I’d like to hear
from the students why they have changed it to eight over twelve divided by
nine over twelve … and I’d then ask them, well, are we adding fractions now?

200

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Similarly, TAS T2 also wanted to hear what the student’s reasons were, with a
view to having the student reconsider the approach (it is not clear if the teacher
believes the present approach is incorrect, inelegant, inefficient, or some combination
of these):
Well, firstly, I would ask them why they did what they did. You know, I think
that’s really important that kids understand how to do something, why it works
and when you use it. And in this case I am intrigued that they’ve got the right
solution by not the standard algorithm so I would ask them why they did it
like that and then I think you could explore with them, and mathematicians do
this all the time, okay, let’s have a look at some other ways that you might do
this and look at why they work and then engage them in a conversation about,
“Let’s weigh up some of the advantages and disadvantages of the different
strategies you’re looking at now.” And see if they would actually change their
mind about what they’ve done there.
These responses seem to prioritise getting access to the student’s thinking, likely
reflecting a preference for solutions that follow the standard “invert and multiply”
algorithm and scepticism about an approach that appears different. There was also
evidence that the degree of scepticism the teachers would feel might be influenced
by their knowledge of (or beliefs about) the student who proposed the solution. In
the context of an exchange about there being more to consider when assessing a
student’s work than just the working that is shown, NZ T4 said, “if you’ve been
working with a student for a while you know” and shortly after, “but you also have
to check that they haven’t done something in the middle, done something crazy.”
By later in the interview, one teacher appeared to have traced the student’s solution
(Figure 7.1) mathematically and wanted to ask the student “how did you get from at
the second line to the third line” (TAS T1).
Given that teachers seemed uncertain about the solution’s validity, it is interesting
that determining the mathematical validity of the student’s method was not expressed
overtly as “the first thing I would do as a teacher” by any of the participants (although
it might be argued that having the student explain the solution allows validity to be
addressed and provides a response that can be made more easily and immediately
in the flow of classroom activity). This possibly reflects a belief in the importance
of attending to student thinking in the classroom; it may also reflect that, in this
case, there was an initial assumption that the method is not correct, and so it is
important to find out what the student is thinking. In both interviews the teachers
moved towards attempting to explain what was going on mathematically but aspects
of their content knowledge were awkwardly held or expressed.
I meant it [putting the fractions over a common denominator] wasn’t needed,
but wasn’t incorrect, but there’s some reason they seem to think somehow
you-, once you put them over the common denominators you can forget the
denominator. (TAS T1)

201

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

NZ T3, who had seen and considered this scenario before, commented about its
validity.
The critical point in this solution, as was pointed out by TAS T1, is dealing with
eight-twelfths divided by nine-twelfths. The teachers’ attempts to explain what is
going on used procedures rather than attempting to make conceptual sense of what
was happening with the fraction division, with at least two teachers suggesting that
from 8/12 ÷ 9/12 we need to consider eight divided by nine, and twelve divided
by twelve, without justifying this mathematically; while another acknowledged that
from 8/12 ÷ 9/12 it is now possible to invert and multiply as per the usual algorithm,
to get 8/9. No teacher attempted to make sense of 8/12 ÷ 9/12 conceptually as, for
example, how many 9 units are there in 8 units.

DISCUSSION

The transcripts of the two discussions provided a rich source from which the
participants’ knowledge could be inferred. From a methodological point of view, it
is important to note that the knowledge that was evidenced was not revealed in the
context of the classroom work of teaching, as has been the case for other pedagogical
content knowledge models (e.g., Ball et al., 2008; Rowland et al., 2005), nor from
any attempt to measure teachers’ knowledge (e.g., Beswick et al., 2012). Rather,
the knowledge that was revealed – and that developed as the discussion progressed
– was uncovered in a social context involving peers. These were teachers, not in
their own classrooms, but looking together at unfamiliar scenarios. It is, therefore,
not surprising that knowledge (or belief) about self as a teacher and mathematician
appeared to influence the confidence with which they participated and the level of
vulnerability that they might have been prepared to expose. A teacher who was not
sure about the validity of the solution may well have been shy about expressing that
uncertainty in front of colleagues, and this may have influenced how specific and
assertive they were prepared to be about what they knew or believed. This point is
worth bearing in mind in teacher education and professional development contexts
where teachers are likely to be sensitive to the perceptions of peers and careful to
protect their beliefs about themselves as competent teachers and mathematicians. It
could be that this issue is more prominent among groups such as those involved in
our study, who are mathematically well qualified and have reputations to protect.
The conversation that unfolded in the New Zealand group provided evidence of one
of the ways in which teachers might protect their beliefs in themselves as competent:
Some time after the discussion had moved from broad agreement that the solution
method was perhaps suspect to the assertion from NZ T3 that it was in fact valid,
NZ T1 attributed the act of dismissing novel solution methods to inexperienced
teachers, despite having contributed to the exchange that had cast doubt on the
method’s validity. NZ T1’s statement at this point could be construed as a reframing
of the conversation in a way that minimised his/her initial position. In addition,
when NZ T3 stated that the solution method was valid she supported the assertion

202

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

with reference to an expert external to the group (the researcher with whom she
had worked) and did so only after NZ T4 had clearly stated that it was correct and
generalisable. This could be construed as NZ T3 protecting her own standing should
the claim be disputed by someone else in the group, as well as suggesting some
lingering uncertainty. Either of these interpretations is consistent with the fact that
NZ T3 felt able to acknowledge her own initially dismissive reaction to the student
solution only after there was consensus that interpreting the solution was challenging
and had “made us think” (NZ T1).
Beliefs research has relied predominantly on Likert-item questionnaires
supplemented with additional sources such as written narratives or interviews
(Ruthven, 2015) or individual interviews (e.g., Beswick, 2018). In this study we
identified evidence of pedagogical content knowledge that could be located in
the beliefs categories in our pedagogical content knowledge framework from
focus group discussions. The methodology employed here makes plain the social,
dynamic and contextual nature of both knowledge and beliefs without turning
to discursive theories of the nature of either construct (e.g., Barwell (2013) in
relation to knowledge and Skott (2015) in relation to beliefs). Inferring teachers’
knowledge from their interactions in a group also further highlights the essential
indistinguishability of knowledge and beliefs. Beliefs research has been criticised
for its reliance on inference from what teachers say and do, and although research on
teachers’ knowledge has less critique in this regard, knowledge is also necessarily
inferred. While there has been broad (although far from complete; e.g., Barwell,
2013) acceptance of the constructivist view that knowledge comprises objective
mental entities that reside “in the heads of persons” (von Glasersfeld, 1995, p. 1)
there has been arguably less willingness to regard beliefs in this way (e.g., Skott,
2015). Objections to acquisitionist conceptualisations of knowledge and beliefs
have been, in part, been founded on recognition of the impossibility of discerning
with any certainty what an individual knows or believes. Proponents of discursive
or participatory approaches avoid this difficulty by focussing on the actions and
utterances of teachers in particular contexts, not with a view to uncovering any
cognitive entities residing in the teachers’ head but rather to understand how the
individual interacts with and in particular contexts. Considering the things that
participants say to be indicative of their knowledge and/or beliefs, as we have done
in this study, allows us to think about the reasons for which they appear to think
as they do, and to suggest ways in which teachers’ beliefs and knowledge might
influenced.
The knowledge that the teachers in this study were able and/or willing to articulate
was dependent upon the context as well as the participants’ previous experiences
and evolved as the conversation progressed. It was expressed with varying degrees
of certainty and confidence, connected with others’ utterances, amended and
reformulated. That is, it appeared to be as subjective and as difficult to infer with any
certainty as beliefs. Whatever knowledge the teachers had or beliefs they held appear
unlikely to have been the same at the beginning and end of the focus group. Rather,

203

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

participation in the group was a learning experience in which the teachers’ beliefs
and knowledge were informed by the interactions in the group as well as the prior
experiences and knowledge/beliefs that each brought with them. It illustrates that
considering beliefs (and knowledge) as dynamic systems, as described by Beswick
(2018), offers a way to think about these entities and, in particular, their fluidity as
they bump up against one another at both the individual level, as that individual
rethinks his/her current understandings, and even more so as the individual beliefs/
knowledge systems of a group of individuals interact. It is not, for example, sensible to
consider teachers’ knowledge or beliefs about teaching independently of their beliefs
about themselves and others. Such a systems perspective retains an acquisitionist
view of beliefs (Skott, 2015) (and knowledge) as something that individuals ‘have’
while emphasising the constant state of flux in which these entities exist particularly
in social situations.
Two aspects of teachers’ beliefs (knowledge) that were prominent in these
discussions have been recognised as among the most important of mathematics
teachers’ beliefs in shaping their practice – beliefs about themselves and particularly
their efficacy as teachers of mathematics (Beswick, 2018), and beliefs about the
capability of students as mathematics learners (Beswick, 2007/2008, 2017, 2018).
Neither of these appear in the framework shown in Table 7.1 but it could be argued
that they should be. Teachers’ beliefs about themselves also appeared to impact their
judgements about the Cognitive demand of tasks based upon the solution shown in
Figure 7.1. For example, the agreement in the New Zealand group that the solution
could be worth showing only to groups of students considered mathematically
capable, arguably because they had found it challenging themselves, can be seen
as a necessary move because to deem it comprehensible by students considered
less capable would mean placing themselves in a similar category. Beswick
(2015) observed a similar phenomenon among prospective teachers who preferred
representations of mathematical ideas that they most readily understood on the
assumption that students would also find these easiest to understand. The prospective
teachers were likely less concerned than the participants in the discussions described
here about their social standing (they were interviewed individually) or protecting
their reputations but, nevertheless, were inclined to believe that others experience
the world, including mathematics, as they do.
In these discussions the teachers’ beliefs about students were implicit in evidence
of their knowledge of student affect. This included attributions of motives to students
such as avoiding showing working because of laziness and doing more work for
enjoyment. The participants may have had particular students in mind when they
made these comments and hence may have been drawing upon a great deal of
contextual and other knowledge in relation to those students. Nevertheless, the
way that these beliefs about students and their affective responses were expressed
suggests that the teachers had generalised these judgements of individuals to broad
categories of students. Even if broadly correct, such generalisations leave open the
possibility that students’ behaviours are miss-judged, including in ways that lead

204

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

them to be considered either mathematically capable or less so on bases that have


little to do with mathematical competence. Beswick (2017) presented evidence of
this phenomenon among teachers who were teaching mathematics out-of-field and/
or much less well-qualified mathematically than those in this study, but there is
at least some evidence here that it may be more widespread among mathematics
teachers. For example, at least one of the teachers (NZ T4) appeared to be more or less
sceptical of novel solutions depending upon the student who produced it. Teachers
can “know-that” students might behave in certain ways, but their “knowing-why”
they behave as they do can include imputed motives that are closer to beliefs than
knowledge, and this knowledge/belief forms the basis of their “knowing-how” and
“knowing-to” respond.
Although the teachers said nothing that we considered mathematically incorrect,
it seemed that they were drawing largely upon procedural knowledge even after
the validity of the solution method had been accepted. Knowledge that was less
mathematical and more pedagogical was more contentious. We have, for example,
in this discussion, questioned the appropriateness of the assessment that it might be
useful to show the solution method shown in Figure 7.1 to mathematically capable
students but, by implication, not others, and have speculated about the validity of
the attributions of student motives that were made. The latter fall into either the
category in Table 7.1, knowledge of Student affect (in relation to content) or the
category, knowledge of Student affect (general). Since these are beliefs categories
and beliefs are by definition somewhat contentious it is not surprising that these
aspects are contentious. Decisions about the student groups for whom the solution
method might be useful, however, fit the category of knowledge of Cognitive
demand of task and may draw also upon other categories such as knowledge of
Student thinking, Examples, or Structure and connections. As discussed earlier
there is an implicit assumption of correctness in relation to knowledge (as opposed
to beliefs) so these examples highlight the fact that there is not a clear delineation
between what is contentious and what is not. In the contexts of the groups in
which they were made, these statements were not challenged and appeared not to
be contentious, but in other contexts, such as when the authors analysed the data,
or among certain groups of mathematics educators, they are contentious. Other
readers may identify other, and perhaps different, examples of knowledge that
they consider contentious thus highlighting the contextual nature of the distinction
between beliefs and knowledge.

CONCLUSION

The inclusion of beliefs categories in the pedagogical content knowledge model


shown in Table 7.1 enabled us to see relationships between teachers’ beliefs and
knowledge more clearly than early versions of this pedagogical content knowledge
framework or other pedagogical content knowledge frameworks, providing a
greater degree of nuance in describing teachers’ knowledge and further insight

205

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

into the complexity of teaching. The framework provided a useful lens through
which to consider the knowledge evidenced by the teachers in our discussion
groups. In particular, we were able to find evidence of the elements added in the
Chick and Beswick (2018) version concerning beliefs about the nature of content
and knowledge of student affect. Our data have highlighted some of the ways in
which these beliefs can interact with other aspects of teachers’ pedagogical content
knowledge to influence teachers’ knowledge and the ways that their knowledge
informs their teaching. We have also highlighted the potentially contentious nature
of many aspects of teachers’ pedagogical content knowledge and hence the blurry
distinction between their beliefs and knowledge. This is, we contend, an important
theoretical contribution to way in which teacher knowledge is conceptualised. It
suggests that studies, ostensibly of teacher knowledge, need to take account of that
knowledge that might conventionally be considered beliefs.
Teachers’ knowledge of students (their thinking, misconceptions and affect)
is a particular part of teachers’ knowledge base that is perhaps most contentious
and most resembles beliefs. Other pedagogical content knowledge framework has
included, drawing from Shulman, knowledge of students, and have specified this in
relation to mathematics (e.g., Ball et al., 2008). Our findings suggest that teachers’
knowledge of student affect in relation to mathematics is an important part of this,
and furthermore, that teachers “knowing-why” students respond to mathematics as
they do, that is their attributions of student affect, are also crucial. Although typically
classified as belief, these are things that teachers “know” about their students and
hence shape the opportunities to learn mathematics that students experience in their
mathematics classrooms. In addition, our data show that teachers’ beliefs about
themselves, including in relation to others, importantly influences the knowledge
and beliefs that they are able and/or willing to express.
In this study we inferred teachers’ knowledge from discussions with peers – a
social context in which the participants’ beliefs about themselves are potentially
threatened and shape the nature and extent of what they share. This methodology
is more akin to discursive or participatory approaches to investigating teachers’
knowledge or beliefs than to typical methods employed in studies of either beliefs
or knowledge. It allowed us to observe teachers’ shifting their understandings and
interpretations as the conversations progressed and to see their knowledge and
beliefs in a degree of flux. Although we have maintained a psychological stance with
its concomitant view of knowledge and beliefs as located within individuals, this
methodology allowed us to see the impacts of the social context on the expression
of those constructs and possibly on their evolution. It afforded a more dynamic view
of teacher knowledge (and beliefs) than is possible using more traditional methods,
including classroom observation in which the teacher acts but without necessarily
articulating his/her thinking. In addition, we retained the cognitivist advantage of
being able to hypothesise about mechanisms that underpinned the behaviour of
individuals as they interacted and that prompted them to modify their views; that is
to learn from their interaction.

206

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Finally, the context of collegial discussion is, or could be, similar to those in
which initial teacher education and professional development of teachers occur. Our
findings suggest that attention to teachers’ affective reactions in such contexts is
warranted in order to create environments in which they can safely confront their
mathematical uncertainties or misconceptions, or their beliefs that do not conform
to normative beliefs of mathematics educators. As suggested by Beswick (2018),
helping teachers to reflect upon the nature and quality of the bases of their knowledge/
beliefs and to challenge one another would help to maximise the impacts of teacher
education and development.

ACKNOWLEDGEMENTS

This research was funded by Australian Research Council grants DP130103144 and
FT140101351. We acknowledge the contributions of other researchers on the first of
these projects: Rosemary Callingham, Ian Hay, Tom Nicholson and Tim Burgess as
well as the teachers who shared their knowledge and insight.

REFERENCES
Ajzen, I., & Fishbein, M. (1980). Understanding attitudes and predicting social behavior. Englewood
Cliffs, NJ: Prentice-Hall.
Australian Curriculum Assessment and Reporting Authority. (2018). The Australian Curriculum:
Mathematics. Retrieved from http://www.australiancurriculum.edu.au/Mathematics/Curriculum/F-10
Baker, M. (2008). An investigation of the effect of teachers’ reflection on their development of pedagogical
content knowledge for teaching primary mathematics (Unpublished Master of Education thesis).
University of Melbourne, Melbourne, Australia.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Barwell, R. (2013). Discursive psychology as an alternative perspective on mathematics teacher
knowledge. ZDM-The International Journal on Mathematics Education, 45(4), 595–606.
Beswick, K. (2005). The beliefs/practice connection in broadly defined contexts. Mathematics Education
Research Journal, 17(2), 39–68.
Beswick, K. (2007). Teachers’ beliefs that matter in secondary mathematics classrooms. Educational
Studies in Mathematics, 65(1), 95–120.
Beswick, K. (2007/2008). Influencing teachers’ beliefs about teaching mathematics for numeracy to
students with mathematics learning difficulties. Mathematics Teacher Education and Development,
9, 3–20.
Beswick, K. (2011). Knowledge/beliefs and their relationship to emotion. In K. Kislenko (Ed.), Current
state of research on mathematical beliefs XVI: Proceedings of the MAVI-16 conference June 26–29,
2010 (pp. 43–59). Tallinn, Estonia: Institute of Mathematics and Natural Sciences, Tallinn University.
Beswick, K. (2015). Inferring pre-service teachers’ beliefs from their commentary on knowledge items. In
K. Beswick, T. Muir, & J. Wells (Eds.), Proceedings of the 39th Annual Conference of the International
Group for the Psychology of mathematics education (Vol. 2, pp. 113–120). Hobart, Australia: IGPME.
Beswick, K. (2017). Raising attainment: What might we learn from teachers’ beliefs about their best and
worst mathematics students? In C. Andrà, D. Brunetto, E. Levenson, & P. Liljedahl (Eds.), Teaching
and learning in math classrooms. Emerging themes in affect-related research: Teachers’ beliefs,
students’ engagement and social interaction (pp. 95–106). New York, NY: Springer.
Beswick, K. (2018). Systems perspectives on mathematics teachers’ beliefs: Illustrations from beliefs
about students. In E. Bergqvist, M. Österholm, C. Granberg, & L. Sumpter (Eds.), Proceedings of the

207

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
KIM BESWICK AND HELEN CHICK

42nd Conference of the International Group for the Psychology of mathematics education (Vol. 1, pp.
3–18). Umeå, Sweden: PME.
Beswick, K., & Callingham, R. (2011). Building the culture of evidence-based practice in teacher
preparation: Instrument development and piloting. In J. Wright (Ed.), Annual Conference of the
Australian Association for Research in Education. Hobart, Australia: AARE. Retrieved from
http://www.aare.edu.au/11pap/papers_pdf/aarefinal00667.pdf
Beswick, K., Callingham, R., & Watson, J. M. (2012). The nature and development of middle school
mathematics teachers’ knowledge. Journal of Mathematics Teacher Education, 15(2), 131–157.
Blömeke, S., & Kaiser, G. (2014). Theoretical framework, study design and main results of TEDS-M. In
S. Blömeke, F.-J. Hsieh, G. Kaiser, & W. H. Schmidt (Eds.), International perspectives on teacher
knowledge, beliefs and opportunities to learn (pp. 19–48). Dordrecht: Springer.
Chick, H. (2007). Teaching and learning by example. In J. M. Watson & K. Beswick (Eds.), Mathematics:
Essential research, essential practice: Proceedings of the 30th Annual Conference of the Mathematics
Education Research Group of Australasia (Vol. 1, pp. 3–21). Sydney, Australia: MERGA.
Chick, H. (2015). The language used to articulate content as an aspect of pedagogical content knowledge.
In M. Marshman, V. Geiger, & A. Bennison (Eds.), Mathematics education in the margins: Proceedings
of the 38th Annual Conference of the Mathematics Education Research Group of Australasia
(pp. 149–156). Sunshine Coast, Australia: MERGA.
Chick, H., & Beswick, K. (2013). Educating Boris: An examination of pedagogical content knowledge
for mathematics teacher educators. In V. Steinle, L. Ball, & C. Bardini (Eds.), Mathematics education:
Yesterday, today and tomorrow: Proceedings of the 36th Annual Conference of the Mathematics
Education Research Group of Australasia (pp. 170–177). Melbourne, Australia: MERGA.
Chick, H., & Beswick, K. (2018). Teaching teachers to teach Boris: A framework for mathematics teacher
educator pedagogical content knowledge. Journal of Mathematics Teacher Education, 21(5), 475–499.
Chick, H. L. (2011). God-like educators in a fallen world. In J. Wrigth (Ed.), Proceedings of the 2011
annual conference of the Australian Association for Research in Education. Hobart, Australia: AARE.
Retrieved from http://www.aare.edu.au/
Chick, H. L., Baker, M., Pham, T., & Cheng, H. (2006). Aspects of teachers’ pedagogical content
knowledge for decimals. In J. Novotná, H. Moraová, M. Krátká, & N. Stehlíková (Eds.), Proceedings
of the 30th Annual Conference of the International Group for the Psychology of Mathematics
Education (Vol. 2, pp. 297–304). Prague, Czech Republic: PME.
Cilliers, P. (2005). Knowledge, limits and boundaries. Futures, 37(7), 605–623.
Dyment, J. E., Chick, H. L., Walker, C. T., & Macqueen, T. P. N. (2018). Pedagogical content knowledge
and the teaching of outdoor education. Journal of Adventure Education and Outdoor Learning, 18(4),
303–322. https://doi.org/10.1080/14729679.2018.1451756
Fantl, J. (2008). Knowing-how and knowing-that. Philosophy Compass, 3(3), 451–470.
Guba, E. G., & Lincoln, Y. S. (1989). Fourth generation evaluation. Newbury Park, CA: Sage.
Hill, H. C., Ball, D. L., & Schilling, S. G. (2008). Unpacking pedagogical content knowledge:
Conceptualizing and measuring teachers’ topic-specific knowledge of students. Journal for Research
in Mathematics Education, 39(4), 372–400.
Krauss, S., Brunner, M., Kunter, M., Baumert, J., Blum, W., Neubrand, M., & Jordan, A. (2008).
Pedagogical content knowledge and content knowledge of secondary mathematics teachers. Journal
of Educational Psychology, 100(3), 716–725.
Leder, G. C. (2015). Forward: From hidden dimensions to dynamic systems in affect research. In B.
Pepin & B. Roesken-Winter (Eds.), From beliefs to dynamic affect systems in mathematics education
(pp. v–x). Dordrecht: Springer.
Lerman, S. (1989). Constructivism, mathematics and mathematics education. Educational Studies in
Mathematics, 20(2), 211–213.
Liljedahl, P. (2008, March). Teachers’ beliefs as teachers’ knowledge. Paper presented at the International
Commission on Mathematical Instruction (ICMI), Centennial Conference, Rome, Italy.
Loughran, J., Berry, A., & Mulhall, P. (Ed.). (2012). Understanding and developing science teachers’
pedagogical content knowledge (2nd ed.). Rotterdam, The Netherlands: Sense Publishers.

208

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BELIEFS AND PEDAGOGICAL CONTENT KNOWLEDGE

Maher, N. (2018). PCK and senior secondary mathematics teaching (Unpublished PhD thesis in
progress). University of Tasmania, Australia.
Mason, J., & Spence, M. (1999). Beyond mere knowledge of mathematics: The importance of knowing-to
act. Educational Studies in Mathematics, 38(1–3), 135–161.
Philipp, R. A. (2007). Mathematics teachers’ beliefs and affect. In F. K. Lester (Ed.), Second handbook
of research on mathematics teaching and learning: A Project of the National Council of Teachers of
mathematics (Vol. 1, pp. 257–315). Charlotte, NC: Information Age Publishing.
Pitfield, M. (2012). Transforming subject knowledge: Drama student-teachers and the pursuit of
pedagogical content knowledge. Research in Drama Education: The Journal of Applied Theatre and
Performance, 17(3), 425–442.
Richardson, V. (1996). The role of attitudes and beliefs in learning to teach. In J. Sikula, T. J. Buttery, &
E. Guyton (Eds.), Handbook of research on teacher education (2 ed., pp. 102–119). New York, NY:
Simon & Schuster Macmillan.
Rowland, T., Huckstep, P., & Thwaites, A. (2005). Elementary teachers’ mathematics subject knowledge:
The Knowledge Quartet and the case of Naomi. Journal of Mathematics Teacher Education, 8(3),
255–281.
Ruthven, K. (2015). Reaction to section 3: Some methodological reflections on studies of mathematical
affect. In B. Pepin & B. Roesken-Winter (Eds.), From beliefs to dynamic affect systems in mathematics
education (pp. 383–393). Dordrecht: Springer.
Ryle, G. (1949). The concept of mind. London: Hutchinson.
Shulman, L. S. (1986). Those who understand: Knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Shulman, L. S. (1987). Knowledge and teaching: Foundations of the new reform. Harvard Educational
Review, 57(1), 1–22.
Skott, J. (2015). Towards a participatory approach to ‘beliefs’ in mathematics education research. In
B. Pepin & B. Roesken-Winter (Eds.), From beliefs to dynamic affect systems in mathematics
education (pp. 3–23). Dordrecht: Springer.
von Glasersfdeld, E. (1995). Radical constructivism: A way of knowing and learning. London: Falmer
Press.

Kim Beswick
Faculty of Education
University of Tasmania

Helen Chick
Faculty of Education
University of Tasmania

209

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS, ANNE BENNISON, STEPHEN QUIRKE,
NIAMH O’MEARA AND COLLEEN VALE

8. DEVELOPING PROFESSIONAL KNOWLEDGE


AND IDENTITIES OF NON-SPECIALIST TEACHERS
OF MATHEMATICS

The professional knowledge needed for teaching mathematics effectively has been
the subject of a great deal of research, mainly focusing on the role of content
knowledge and pedagogical content knowledge and their interactions. Additionally,
there is growing interest in the concept of identity in relation to exploring teachers’
professional roles and development. Our focus in this chapter is on the professional
knowledge and identities of non-specialist teachers of mathematics. Here we are
referring firstly to those “out-of-field” teachers in secondary schools who are
teaching mathematics without formal qualifications in mathematical content
or pedagogy, and secondly to teachers who are recognizing and exploiting the
numeracy demands of subjects other than mathematics. We present a focused review
of international research on non-specialist mathematics teachers’ knowledge and
identities, and then draw on our research conducted in Australia and Ireland to
discuss challenges in teacher development when crossing subject boundaries.

INTRODUCTION

Research in mathematics teacher education has adopted different theoretical lenses


to understand how teachers learn and develop. Many studies invoke cognitive/
constructivist theories that conceptualise teacher learning as the growth of professional
knowledge. Alternatively, studies drawing on sociocultural theories propose that
teachers’ learning can be understood as increasing participation in socially organised
practices that develop their professional identities (Lin & Cooney, 2001). Each of
these theoretical perspectives has provided rich insights into mathematics teacher
development. Similarly, these perspectives have provided complementary lenses for
exploring issues concerning the professional knowledge and identities of teachers of
mathematics who lack formal qualifications in this subject. We refer to such teachers
as “non-specialist teachers of mathematics.”
In this chapter our focus is on two types of non-specialist teachers: first, those
“out-of-field” teachers in secondary schools who are teaching mathematics without
formal qualifications in mathematical content or pedagogy; and secondly, teachers
of subjects other than mathematics who are encouraged by national curriculum

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_009

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

policies to recognise and exploit the numeracy demands of their subjects in order
to develop their students’ curriculum-specific numeracy capabilities. These two
phenomena no doubt manifest in different ways in different parts of the world. Our
exploration aims to illuminate the circumstances that give rise to these phenomena,
ways of characterising out-of-field teachers and teachers who embed numeracy
across the curriculum, and research-based approaches to supporting such teachers in
developing their professional knowledge and professional identities in the domain
of mathematics.
In the next sections of the chapter we offer a brief survey of research on mathematics
teacher knowledge and identity development. The subsequent section explores
issues and problems surrounding non-specialist teaching of mathematics. We then
provide snapshots of emerging research in Australia and Ireland on developing the
knowledge and identities of out-of-field teachers of mathematics and teachers who
are embedding numeracy into subjects other than mathematics.

KNOWLEDGE REQUIRED FOR TEACHING

The first theoretical lens we explore for studying teacher learning involves the
growth of professional knowledge for teaching. Teacher knowledge has been a topic
of interest for researchers in the field of mathematics education since the seminal
work of Shulman was published in 1986. Before the mid-1980s research on effective
teaching had focused on questioning, direct instruction and timing issues instead
of teacher knowledge. Shulman’s (1986) influential article called for a change of
emphasis, citing subject matter knowledge as the missing paradigm in previous
research in the field. In order to address this gap in the literature Shulman proposed a
model of teacher knowledge. This model was not mathematics specific but outlined
three different knowledge types required for teachers of any subject: subject matter
content knowledge, pedagogical content knowledge, and curricular knowledge.
Content knowledge was defined as “going beyond knowledge of the facts or concepts
of a domain [e.g., mathematics]. It requires understanding the structures of the subject”
(p. 9). Shulman argued that teachers should first be concerned with developing their
subject matter content knowledge, and only then begin to concentrate on developing
their pedagogical content knowledge. Pedagogical content knowledge refers to
the transformation of a teacher’s subject matter knowledge into representations,
explanations, analogies, illustrations, examples and demonstrations that students
could easily comprehend. Shulman also highlighted the interrelated nature of these
two knowledge domains, proposing that developing pedagogical content knowledge
would serve to enhance one’s subject matter content knowledge. He argued that more
time should be dedicated to developing pedagogical content knowledge, both during
initial teacher education and throughout one’s professional career because of the
important role this type of knowledge plays in a teacher’s overall knowledge base.
Shulman additionally advocated that teachers should develop curricular knowledge.

212

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

He viewed the curriculum as a tool capable of enhancing pedagogy by providing


teachers with alternative approaches, instructional methods and texts.
Since the publication of Shulman’s influential work, many other researchers
have investigated the nature of knowledge that underpins effective teaching and
developed their own models of this knowledge. From a mathematics perspective,
researchers such as Ernest (1989), Fennema and Franke (1992), Rowland, Huckstep,
and Thwaites. (2005), Ball, Thames, and Phelps (2008), and O’Meara (2011)
have developed models of knowledge for effective teaching of mathematics. We
summarise two of these more recently developed models in the following sections.

Rowland’s Knowledge Quartet

Rowland and his colleagues at the University of Cambridge developed the Knowledge
Quartet in 2005 as a tool for observing and reviewing mathematics teaching with
student teachers (Rowland, in this volume). The four different knowledge domains
in this model are:
‡ Foundation knowledge;
‡ Transformation knowledge;
‡ Connection knowledge;
‡ Contingency knowledge.
The foundation knowledge dimension of this model refers to teachers’
mathematical content knowledge and beliefs. The name foundation indicates how,
as was the case with the Shulman (1986) model, content knowledge forms the basis
from which all other knowledge types can be developed. It is only when teachers
have become proficient in this area that they should concern themselves with the
other knowledge domains required for teaching.
The next component in this model is transformation knowledge, or “knowledge-
in-action.” While foundation knowledge is the core component of this model,
transformation knowledge is seen as the knowledge domain that distinguishes a
mathematics teacher’s knowledge base from that of a mathematician or layperson.
This assertion aligns with the thinking of Shulman (1987), who stated that a teacher’s
knowledge base is characterised by “the capacity of a teacher to transform the
content knowledge he possesses into forms that are pedagogically powerful” (p. 15).
The third dimension of Rowland’s model is connection knowledge: it is this
domain that links teachers’ knowledge of students and their learning. This knowledge
type refers to teachers knowing how to arrange and sequence topics in a way that
students will understand, to make connections between different mathematical
concepts and topics, and to highlight the link between mathematics and students’
everyday lives inside and outside of school. This is the knowledge, when combined
with those previously discussed in the model, which equips teachers with the skills
necessary to teach for understanding.

213

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Finally, Rowland et al. (2005) outline the need for teachers to develop
contingency knowledge, involving flexibility to deal with the unexpected. As the
name contingency suggests, this knowledge may be called upon only irregularly and
requires the teacher to be aware of how to respond to unexpected situations and to
deviate from an intended lesson plan if necessary. Contingency knowledge draws on
the other three knowledge types. Without a strong content (foundation) knowledge
teachers will be unable to deal with any variation in a planned lesson, while
pedagogical (transformation) knowledge will enable teachers to alter lesson plans
with relative ease. An awareness of the applications of mathematics (connection)
will make it easier for teachers to cope with uninterested students who unexpectedly
question the relevance of any topic.

Ball’s Model of Mathematical Knowledge for Teaching

Another well-known model that stemmed from the work of Shulman (1986) was
that proposed by Deborah Ball and her colleagues at Michigan State University in
2008. Ball’s model, as depicted in Figure 8.1, sought to elaborate on two of the three
knowledge dimensions proposed by Shulman.
Content (subject matter) knowledge again underpins this model of teacher
knowledge with Ball et al. (2008) dividing this domain into three inter-related subsets.
Common content knowledge is the mathematical knowledge needed to solve routine,
everyday tasks (similar to foundation knowledge in the model of Rowland et al.,
2005). It is referred to as common since it is a knowledge domain that is not specific

Figure 8.1. Ball’s model of Mathematical Knowledge for Teaching (Ball et al., 2008, p. 403)

214

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

to teaching, but rather the mathematical knowledge possessed by the majority of the
population. Horizon knowledge is the second subset of subject matter knowledge.
This domain refers to a teacher’s knowledge of the relationship between different
mathematical concepts over the span of the mathematics curriculum, thus enabling
teachers to highlight the interrelated and coherent nature of the discipline. Horizon
knowledge is closely linked with the knowledge of connections that is central to
Rowland’s Knowledge Quartet. A further similarity between the models proposed
by Rowland et al. (2005) and Ball et al. is that both models incorporate a knowledge
domain that is unique to the profession of teaching. Ball et al. label this domain as
specialised content knowledge – the third subset of subject matter knowledge and
the one that they saw acting as a hybrid between subject matter knowledge and
pedagogical content knowledge. They explained the need for this knowledge type:
The mathematical demands of teaching require specialised mathematical
knowledge, needed by teachers and not needed by others. Accountants have to
calculate and reconcile numbers and engineers have to mathematically model
properties of materials, but neither group needs to explain why, when you
multiply by 10 you “add a zero.” (p. 401)
Ball et al. outlined an extensive list of teacher-specific tasks that this knowledge
domain serves, including the ability to respond to “why” questions posed by students,
the ability to ask productive questions in the mathematics classroom, and the ability
to find an example to convey their message. The comprehensive nature of the list
outlined in the work of Ball et al. shows how specialized content knowledge is a
fundamental element of a teacher’s package of knowledge and is considered to be
the defining feature of the Ball et al. model (O’Meara, 2011).
As is the case in the models discussed previously, Ball et al. (2008) maintain that
only when teachers have developed deep subject matter knowledge are they in a
position to focus their attention on developing sound pedagogical content knowledge.
This domain, in the right-hand half of Figure 8.1, is also divided into three subsets.
Knowledge of content and teaching refers to “knowing about teaching and knowing
about mathematics” (p. 401). It combines what Shulman would have referred to as
subject matter knowledge with pedagogical knowledge. This type of knowledge is
necessary when teachers have to make key instructional decisions such as how to
best sequence a topic or what representation is most appropriate in a given context.
It therefore shares some characteristics of transformation knowledge in Rowland’s
Knowledge Quartet. The second subset of pedagogical content knowledge is a
knowledge of content and students, which again combines elements of Shulman’s
subject matter knowledge and pedagogical knowledge. This knowledge type allows
teachers to pre-empt students’ misconceptions; to identify, in advance, areas where
students will struggle; and to better understand the thinking of the different students
in the class (cf contingency knowledge in the model of Rowland et al., 2005). The
third subset of pedagogical content knowledge described by Ball et al. refers to
knowledge of content and curriculum. By including this dimension in their model

215

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Ball et al. are highlighting how it is necessary for teachers to develop a knowledge
of the materials and resources available to help students navigate their way through
the curriculum.
While research on mathematics teacher knowledge has highlighted the existence
of knowledge that is unique to the teaching profession, we must look to theories of
identity to understand the challenges of teaching mathematics as a non-specialist.

THE CONCEPT OF IDENTITY

The concept of identity is experiencing a renaissance with the enduring notion


being adopted in the fields of psychology, sociology, cultural studies, anthropology,
history and educational research (Sfard & Prusak, 2005). In the social sciences,
identity studies have become a rapidly growing area of interest, as research with
this orientation seeks to understand people’s psychological experiences and social
behaviour (Côté, 2006). From an educational research perspective, Hoffman (1998)
states that “in many ways identity has become the bread and butter of our educational
diet” (p. 324). This surge in educational researchers’ interest in identity is evidenced
by the large number of studies produced within the past ten years that employ the
concept, in particular to investigate teacher identity (e.g., Akkerman & Meijer, 2011;
Beauchamp & Thomas, 2009; Beijaard, Meijer, & Verloop, 2004; Canrinus, Helms-
Lorenz, Beijaard, Buitink, & Hofman, 2011; Mockler, 2011).
Over the last two decades, there has been a significant increase in identity studies
in mathematics education research (Darragh, 2016; Heyd-Metzuyamin, Lutovac,
& Kaasila, 2016). If the 1980s and 1990s signalled a social turn (Lerman, 2000),
then the 2000s and beyond now signal an identity turn (Chronaki, 2013). However,
the meaning of identity has been interpreted in a variety of ways (e.g., Gee, 2001;
Holland, Lachicotte, & Skiner, 1998; Sfard & Prusak, 2005; Wenger, 1998) and
many researchers in their theoretical and empirical writing on this topic do not
include a definition of identity, assuming that its meaning is self-evident. The lack of
explicit definition of “identity” has been highlighted in reviews of research literature
conducted by Beijaard et al. (2004), and by Darragh (2016). Darragh additionally
claimed that, as a result, authors sometimes have collected inappropriate data and
drawn conclusions that are inconsistent with the view of identity they have espoused.
Hence, Graven and Lerman (2014) argue that it is necessary for mathematics
education researchers to clearly specify their use of the term “identity,” the sources
that have shaped their perspectives on identity, and the pertinence of their work with
identity with regard to the teaching and learning of mathematics.
Darragh (2016) identified two different paradigms at work in identity research
in mathematics education – a psychological paradigm in which identity is an
acquisition, or something we are, and a sociological paradigm that sees identity as
an action, or something we do.
The psychological branch of identity research has mainly stemmed from the
work of Erikson (Beijaard et al., 2004). Erikson regarded identity as an acquisition.

216

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

In Erikson’s view, identity involves a person’s biography unfolding within an


epigenetic process according to a “ground plan” that occurs in ordered stages
and sequences (Weigert & Gecas, 2005). Erikson’s theory characterises identity
formation by the stages through which people progress based on biological and
psychological maturation (Beijaard et al., 2004). This developmental perspective
prioritises the individual’s mental traits, states and dispositions (Côté, 2006). In this
paradigm, identity is the property of persons – that is, something one acquires. The
acquisitionist perspective envisages obtaining a core, fixed and stable identity that is
the answer to the question “Who am I?” (Weigert & Gecas, 2005).
The acquisitionist paradigm has been criticised for producing a dichotomy
between the individual and society that fails to recognise the complex ways in
which identities are formed and operate in different social contexts. The alternative
sociological paradigm views identity as a performative process of becoming, and
recognises that social contexts have the capacity to shape human development. From
the sociological perspective, to understand the “self,” one must also understand
the society in which the “self” is acting, and that the “self” is always acting in a
social context in which other selves exist (Stryker, 1980). From this perspective,
then, the answer to the question “Who am I?” is determined by external social
structures and roles. Mead, a sociologist who studied social behaviourism in the
early 20th century, developed the notion that the self and identity are interlinked.
He suggested that the self is comprised of two components, “I” and “me.” The “I”
is the personal component and the subject-self, reflectively articulating thoughts.
The “me” is the sociological component and the object-self – it is the identity that
the self develops. Mead concluded that the “I” and the “me” must co-exist, which
leads to the notion of multiple and possibly contradictory identities (Darragh,
2016).
Within the sociological frame of identity as action, identity is seen as having
several key features. First, individuals are recognised as participating in a number
of communities, suggesting they will have multiple identities that depend on the
practices and contexts of each community (Wenger, 1998). Part of the negotiation
process of identity development involves reconciliation of multiple identities into
what Gee (2001) has called a core identity that holds across these contexts. A second
important feature of identity is its dynamic nature (Beijaard et al., 2004; Holland et
al., 1998; Wenger, 1998). Wenger (1998) claimed that identity development involves
a learning trajectory that connects the past, the present, and the future. Therefore,
learning provides a mechanism for moving from a current to a future identity and a
context that enables an individual to determine what is important and what is not, in
other words, “what contributes to our identity and what remains marginal” (p. 155).
Finally, this observation suggests that an individual can act as an agent in his or her
own identity development, which opens up possibilities for awareness and choice in
learning trajectories.

217

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Identity in Mathematics Education Research

Cutting across the broad psychological and sociological paradigms in identity


research, Darragh (2016) distinguished five ways in which identity was defined in
the mathematics education research literature: participative, narrative, discursive,
psychoanalytic, and performative. The majority of research appeared to be
represented in the first three categories. Wenger’s (1998) notion of communities
of practice has been highly influential within the participative conceptualisation
of identity. Wenger’s social theory of learning comprises four components:
meaning, practice, identity, and community, which are “deeply interconnected and
mutually defining” (p. 5). In Wenger’s framework identity is a “lived experience of
participation in specific communities” (p. 151) in which meaning is negotiated by
engaging in a number of communities of practice. This conceptualisation of identity
results from an individual’s participation in “a set of distinctive practices” (p. 105,
original emphasis). However, this view contrasts with the definition of narrative
identity offered by Sfard and Prusak (2005) who, while agreeing with Wenger that
identities originate in practice, claimed that identity is about how practices are talked
about. They define identity as “those narratives about individuals that are reifying,
endorsable and significant (p. 16, original emphasis). Interpretations of discursive
identity differ amongst mathematics education researchers. While some draw on
Gee’s (2001) notion of discourse identity, emanating from what is said about an
individual by themselves and other people, others take a post-structuralist approach
to discourse as a wider societal meta-narrative, for example, in analysing the way
girls are subjectively constructed within mathematics.

(Mathematics) Teacher Identity

The notion of identity is seen by many researchers as providing useful insights


into the learning and practices of teachers. Sachs (2005) argued that, in addition
to acquiring the necessary knowledge and skills, learning to be a teacher involves
developing an understanding of what it means to be a teacher, and thus becoming a
teacher involves developing a professional identity.
Teacher professional identity then stands at the core of the teaching profession.
It provides a framework for teachers to construct their own ideas of ‘how to
be’, ‘how to act’, and ‘how to understand’ their work and their place in society.
Importantly, teacher identity is not something that is fixed nor is it imposed;
rather it is negotiated through experience and the sense that is made of that
experience. (p. 15)
This portrayal refers to some of the features of identity highlighted in the previous
section: that teacher identity is an ongoing process of becoming, it involves both the
person and the context, and agency is an important element.

218

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

Despite widespread acknowledgement of the importance of teacher identity as


a means of understanding how teachers learn, there are challenges in analysing
identity development. Studies have typically investigated factors that influence
teacher identity. Most have focused on particular aspects; for example, Enyedy,
Goldberg and Welsh (2005) described science teachers’ identities in terms of beliefs
about the nature of teaching and learning and the nature of science, goals for the
classroom and instruction, and knowledge of science and pedagogy; while Williams
(2011) elicited biographical narratives of two mathematics teachers that drew on
their life history and included information about family and background along with
school, university, and teaching experiences.
Other studies have investigated teacher identity as part of broader frameworks,
revealing additional factors that influence how a teacher’s identity develops.
Graven (2004) used Wenger’s (1998) framework of meaning, practice, identity, and
community but found that teachers in her study increasingly mentioned confidence
as the study progressed. For Graven, confidence was “part of an individual teacher’s
ways of learning through experiencing, doing, being, and belonging. As such, it is
deeply interconnected with learning as changing meaning, practice, identity, and
community” (p. 179). Graven therefore argued that confidence could be added to
the four components of meaning, practice, identity, and community in Wenger’s
framework as a fifth overarching component. The interconnected nature of the
components means that it could be argued that confidence is a factor that influences
identity
Most of the research into mathematics teachers’ professional knowledge and
identity has been conducted with teachers qualified to teach this subject, or without
taking into account teachers’ qualifications, expertise, and experience in teaching
mathematics. However, in this chapter we are interested in non-specialist teachers
of mathematics – how do they develop the necessary knowledge for teaching the
subject, as well as an identity aligned with the discipline of mathematics?

NON-SPECIALIST TEACHERS OF MATHEMATICS: EXPLORING


ISSUES AND PROBLEMS

Grootenboer and Zevenbergen (2008) suggest that a teacher’s well-developed


mathematical identity should be built on significant mathematical knowledge and
skills with a positive attitude towards the subject and a sense of satisfaction from
carrying out mathematical practices. In addition, these authors claim that such
teachers see mathematics as fundamental to their broader identity and integral to
defining their sense of self and vocation. This view connects professional knowledge
and professional identity as crucial influences on mathematics teacher development.
However, in this chapter we consider the case of non-specialist teachers of
mathematics whose primary professional knowledge base and professional identity
are aligned with another discipline.

219

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Teaching Mathematics Out-of-Field

The phenomenon of out-of-field teaching is often defined as “teachers assigned


by school administrators to teach subjects which do not match their training or
education” (Ingersoll, 2002, p. 5). The practice of assigning secondary school
teachers to teach subjects for which they are not qualified or educated to teach is
an ongoing issue in many countries around the world, such as Australia, Germany,
Ireland, South Africa, the United States of America, and England (du Plessis,
Gillies, & Carroll, 2014; Ingersoll, 2002; Marginson, Tytler, Freeman, & Roberts,
2013; McConney & Price, 2009; Mullis, Martin, Foy, & Hooper, 2016; Törner &
Törner, 2012; Vale, McAndrew, & Krishnan, 2011). However, there are differences
around the world in how this phenomenon is manifested and understood, and also
many international variations in what counts as “out-of-field” due to local and
national differences in teacher education, certification, and assignment practices
(Price et al., 2019).
Certification obtained from teacher education programs is commonly utilised as
the measure for teacher qualification and teacher quality. However, some researchers
argue that it is an over-simplification to rely on certification as an indicator of in-
field teaching (Sharplin, 2014), and that the term “out-of-field” unfairly implies a
deficit on the part of the teacher (Törner, 2014). Du Plessis (2017) instead argues that
expertise can develop with teaching experience over time and proposes that out-of-
field teaching should be defined as “Teachers teaching in subject areas or year levels
outside their field of qualifications or expertise” (p. 11). Thus, this issue is much
more complex than a binary distinction: it necessitates examining the degrees of
fit or misfit between a teacher’s subject assignment, qualifications, and experience.

Incidence of out-of-field teaching of mathematics. Only a brief and selective


summary of the incidence of out-of-field teaching of mathematics is provided here.
For a more extensive and nuanced discussion of practical and philosophical issues in
deciding who is teaching “out-of-field,” see Price et al. (2019).

‡ Australia
In Australia, there are national standards for accreditation of initial teacher education
programs (AITSL, 2018). The requirement for a secondary teaching specialisation
is at least a major study in one teaching area (six units studied over three years)
and preferably a minor study in a second teaching area (four units studied over two
years). In all but one Australian state, teachers are registered as teachers rather than
formally certified to teach only the subjects or year levels for which they are qualified.
Responsibility for appointment of teachers to schools varies from being centrally
organised by a state Education Department to allowing principals to advertise and
appoint staff. In all cases, however, principals have discretion in assigning teachers
to teach specific subjects and year levels.

220

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

The scale of out-of-field teaching of mathematics in Australia is greater than


other comparable countries (Marginson et al., 2013; Mullis, Martin, Foy, & Hooper,
2016). Whilst there are higher proportions of teaching out-of-field in non-STEM
subjects such as geography and media, more than one-third of those teaching Years
7–10 mathematics (37%) are teaching out-of-field (Weldon, 2016). In 2011, 34% of
Year 8 mathematics students in Australia were taught by teachers without a major in
tertiary mathematics or mathematics teacher education compared to the international
averages of 12% and 8% respectively (Mullis et al., 2012). This reduced to almost
a quarter (22%) in 2015 (Mullis et al., 2016). Furthermore, the incidence of out-
of-field teaching is much higher in remote locations (41%) than provincial (32%)
and metropolitan locations (24%) and higher in low socio-economic schools (31%)
than high socio-economic schools (22%) (Weldon, 2016). A further concern is the
increasing incidence of beginning teachers teaching out-of-field. Weldon (2016)
reported that 37% of teachers in their first two years post-graduation are teaching
out-of-field compared with 25% of teachers with more than five years’ experience.
Another Australian study reported that up to 23% of graduate teachers with secondary
discipline specialisation qualifications are teaching out-of-field (Mayer et al., 2014)
with secondary graduates with specialisations in humanities, the arts and health and
physical education the most likely to be teaching out-of-field. This means that the
responsibility for teaching mathematics, especially in junior secondary rural and low
socio-economic schools is being charged to inexperienced teachers from non-STEM
specialist disciplines.

‡ South Africa
Research carried out by the Centre for Development and Enterprise (2015) on Teacher
Evaluation in South African schools acknowledges the occurrence of out-of-field
teaching; however, there are no official statistics on the out-of-field phenomenon.
Nevertheless, small-scale surveys of secondary schools have revealed that all of the
teachers who responded had taught out-of-field at some stage (Steyn & du Plessis,
2007).
Research conducted by Graven (2004), Bertram, Mthiyane, and Mukeredzi
(2013), and van Putten, Stols, and Howie (2014) highlights that teachers in South
Africa are being co-opted into teaching mathematics “because there is no one else
to do so in a particular school, or some such circumstance” (van Putten et al., 2014,
p. 371). Each of these authors refers to South Africa’s apartheid history as having
a significant impact on teacher education and teacher qualifications. Adler and
Davis (2006) explain that between 1976 and 1996, apartheid education reinforced
racial and economic inequality, with primarily black teachers having little or no
opportunity to teach mathematics. Due to this practice, curriculum reform in South
Africa has focused on redressing and repairing the damage this caused, mainly
though professional development programmes for teachers.

221

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

‡ England
Most secondary school teachers in England qualify via a one-year Postgraduate
Certificate in Education (PGCE) with a subject specialisation based on their
undergraduate degree. Applicants for teacher education programs whose degree
subject does not link closely to their chosen teaching subject may participate in a
8–12 week subject knowledge enhancement course before enrolling.
In England, data collected through the School Workforce Census has highlighted
the existence of unqualified teaching at secondary level, largely because demand
outstrips supply of qualified teachers. In this context, an unqualified teacher
is defined as not having attained Qualified Teacher Status (QTS) in England
(Ogilvie & O’Brien, 2015). The initial School Workforce Census in 2010 found
that 26.6% of mathematics teachers did not hold a degree in their subject (Loveys,
2011). Findings from the 2016 School Workforce Census indicate that 22.3%
of mathematics teachers are teaching the subject out-of-field (Department for
Education, 2017). The distribution of these non-specialist mathematics teachers is
uneven throughout England with greater numbers estimated to be working in regions
of social deprivation. This is a recurrent finding amongst countries in which out-of-
field teaching is prevalent.

‡ Ireland
Teaching is a high-status occupation in Ireland and attracts the highest academic
achievers when they leave school. Accreditation requirements for initial teacher
education programs set out not only the number of units to be studied in relation
to subject matter knowledge but also prescribed topics to be covered. The Teaching
Council of Ireland (2013, 2017) sets out subject-specific criteria for teacher
registration purposes. In mathematics, fully qualified teachers must have a degree-
level qualification with at least one-third of the degree comprising specific study
of mathematics. There are also minimum credit requirements in analysis, algebra,
geometry, and probability and statistics with additional credits to be obtained in a
variety of optional topics. However, school principals are responsible for advertising
and appointing staff and for assigning teachers to subjects and classes. Thus,
although there are prescriptive requirements for qualifying as a mathematics teacher
in secondary schools, principals are under no obligation to assign only fully qualified
mathematics teachers to teach mathematics classes.
In response to the growing concern over the underperformance of Ireland’s
students in national examinations and international assessments such as the
Programme for International Student Assessment (PISA) 2006, Ní Riordáin and
Hannigan (2009) conducted a study on the prevalence of out-of-field teaching in
mathematics in secondary schools. For this study, these authors adopted Ingersoll’s
(2002) definition of out-of-field teaching cited earlier in this chapter. This study
discovered that 48% of those who teach mathematics in secondary schools did not
possess a mathematics teaching qualification recognised by the Teaching Council.
This research also indicated that mathematics teachers over the age of 35 were more

222

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

likely to be in-field whereas 60% of mathematics teachers under the age of 35 were
out-of-field. It was of some concern to these researchers that, although nearly half
the surveyed teachers were out-of-field, 78% of respondents nevertheless considered
their qualification as sufficient to teach secondary school mathematics.

Causes of out-of-field teaching of mathematics. Many of the causes of out-of-field


teaching appear to be common across different educational settings. However, in
some cases the causes are unique to the specific country. In South Africa, for example,
the history of apartheid has affected teacher qualifications and mathematics teacher
education (Adler & Davis, 2006; Bertram et al., 2013). For teacher candidates of
different races there were different teacher education programmes that differed in
duration (Welch, 2002, cited in Bertram et al., 2013). This resulted in racial and
economic inequality with black teachers receiving poor opportunities to learn
mathematics and teaching (Adler & Davis, 2006). Therefore, the apartheid history is
a factor in the shortage of suitably qualified mathematics teachers in South Africa.
In an overview of the phenomenon of out-of-field teaching, Price (2017) notes
that issues related to teacher supply and demand have been identified as a possible
reason for out-of-field teaching in Australia and Ireland. In Australia, for example,
Hobbs (2013) discovered that teachers were teaching out-of-field because they were
covering another teacher’s load on a short or medium term, filling in on a longer
term to cover a teacher’s long term absence, as a consequence of “load allocation”
as no other teachers were available, or by request as part of a career move to teach a
subject that was new and interested them.
Nevertheless, there exists an alternative argument that the assignment of non-
specialist teachers to teach out-of-field is not a short-term problem of supply of
appropriately qualified teachers to meet increasing population or changes in the
school curriculum, but rather a long-term problem of failure to attract and retain
teachers in the least advantaged schools (Ingersoll, 2011; Masters, 2015; Vale &
Drake, 2019). Of concern is the way in which the incidence, causes and implications
of out-of-field teaching is obscured by governments, education systems and
principals through their policies and procedures.
One way in which these policies play out in many parts of the western world
is illustrated by the provision of government schools with more autonomy and an
increasing focus on school performativity and accountability. Schools are increasingly
given the autonomy to recruit teachers, instead of having teachers assigned by a
national or regional government Department of Education. However, schools in
low-socio economic areas or rural and remote locations have not been able to attract
the most qualified and experienced teachers (Donaldson, 2013; Ingersoll, 2011). In
Australia, special incentives such as funding to support accommodation and travel
to teach in remote locations, and specialist training programs such as Teach for
Australia, have not been sufficient to retain teachers in rural and remote schools after
the expiry of the agreement (Handal, Watson, Patocz, & Maher, 2013; Vale & Drake,
2019). Furthermore, teacher salaries and working conditions are not sufficient to

223

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

attract or retain teachers (Ingersoll, 2011; Vale & Drake, 2019), especially when
teachers who are teaching out-of-field are not supported by a collaborative school
culture (du Plessis, Hobbs, Luft, & Vale, 2019).

Implications of out-of-field teaching for teacher professional knowledge and


identity. Out-of-field teachers have been found to suffer from a lack of confidence,
stress and feelings of inadequacy (du Plessis, 2016; Steyn & du Plessis, 2007). These
issues arise from their experience of working in performance-focused educational
environments in which school leaders focus more on results than caring for human
needs. Du Plessis (2015) found that out-of-field teachers’ lack of pedagogical content
knowledge and subject matter knowledge was fundamental to these concerns, and
that the out-of-field phenomenon hampered development of a professional identity
in an out-of-field subject.
Hobbs (2012) notes that technical measures, such as teacher certification, type
of degree, college ratings and test scores, have often been used to link teacher
characteristics to student outcomes. While these factors may be important indicators
of teacher competence, she argues that there are other aesthetic dimensions of teaching
that can be used to describe teacher effectiveness and understand how teachers relate
to the subject they are teaching. Hobbs proposes an aesthetics framework integrating
the cognitive and affective domains to analyse the relationship between teacher
knowledge, identities and passions. She explains that as an aesthetic experience
maintains its value and wholeness it results in deep meaning and learning that is unified
and coherent. Hobbs conceptualises this coherence in terms of teachers’ knowledge
of content, students and pedagogy. Using the aesthetics framework to analyse the
experiences of out-of-field teachers teaching mathematics and science across
subject boundaries, Hobbs found that their limited subject teaching qualifications
hindered their aesthetic understanding of what the subject could offer their students.
Aesthetic understanding, combining emotion, cognition and action, is claimed to
be transformative and thus contributes to a teacher’s evolving identity. Becoming a
specialist subject teacher is a continuous process shaped and reshaped through the
teacher’s interactions with and reflections on professional and personal experiences
with that subject. Thus out-of-field teachers may find it difficult to develop
… a coherent and unified sense of what the subject is about and how to bring it
to life for students, and to be transformed by what he/she knows and believes
in a way that aligns them to personally and professionally identify with the
subject. (p. 726)
In this section we have explored the characteristics, incidence, causes, and effects
of out-of-field teaching, noting the complexities of the out-of-field phenomenon
and how it is constructed in different contexts. Although teaching out-of-field is
most often defined in terms of the professional knowledge that teachers possess
or lack, there are clear links between knowledge and the development of teachers’
professional identities.

224

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

Embedding Numeracy across the School Curriculum

A second kind of non-specialist teacher of mathematics can be characterised


as those teachers who recognise the numeracy demands of the subjects they
teach. The emergence of international testing regimes such as the Programme for
International Student Assessment (PISA) and the Programme for the International
Assessment of Adult Competencies (PIAAC) has created an increased focus on
mathematical literacy, also referred to as numeracy. Numeracy is a term used to
identify knowledge, skills and practices related to the use of mathematics in non-
mathematical contexts and, in particular, to the use of mathematics in work, home
and civic life. Steen (2001) identified seven dimensions of numeracy (using the term
quantitative literacy): confidence with mathematics; appreciation of the nature and
history of mathematics and its significance for understanding issues in the public
realm; logical thinking and decision-making; use of mathematics to solve practical
everyday problems in different contexts; number sense and symbol sense; reasoning
with data; and the ability to draw on a range of prerequisite mathematical knowledge
and tools. Steen argued that numeracy development is the responsibility of all
teachers, and that opportunities abound for developing students’ numeracy across
the whole school curriculum.

Rationale for numeracy as a cross-curricular commitment. While the idea of


embedding numeracy across the curriculum is challenging, it has been taken up
in education policies in countries such as Australia and Ireland. In Australia, the
rationale for including numeracy in the curriculum has evolved over 30 years and
three national Declarations on the goals of schooling agreed by the State, Territory,
and Australian Ministers for Education (e.g., Ministerial Council on Education,
Employment, Training and Youth Affairs, 2008). Recognition of numeracy as an
important goal for schooling was further confirmed through a national numeracy
review (Council of Australian Governments, 2008), which also promoted the view
that the development of students’ numeracy requires a cross-curricular commitment
by schools and systems. This review recommended that:
… all systems and schools recognise that, while mathematics can be taught
in the context of mathematics lessons, the development of numeracy requires
experience in the use of mathematics beyond the mathematics classroom, and
hence requires an across the curriculum commitment. (p. 7)
Further, numeracy has been identified as one of seven General Capabilities
embedded in the Australian Curriculum. Numeracy is described within each school
subject’s curriculum document via the following statement:
Students become numerate as they develop the knowledge and skills to use
mathematics confidently across other learning areas at school and in their lives
more broadly. Numeracy encompasses the knowledge, skills, behaviours and

225

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

dispositions that students need to use mathematics in a wide range of situations.


It involves students recognising and understanding the role of mathematics
in the world and having the dispositions and capacities to use mathematical
knowledge and skills purposefully. (Australian Curriculum, Assessment and
Reporting Authority, n.d.)
A commitment to developing numeracy across the curriculum is also evident in
the Australian Professional Standards for Teachers, a set of statements developed
by the Australian Institute for Teaching and School Leadership (2017) that specify
the professional knowledge, professional practices, and professional engagement
required of effective teachers. The standards are used as a mandatory framework
for accrediting initial teacher education programs in Australia. Standard 2 states
that teachers should “Know the content and how to teach it,” and one of the focus
areas elaborating on this statement relates to knowledge of literacy and numeracy
strategies. Thus, graduate teachers should “know and understand literacy and
numeracy teaching strategies and their application in teaching areas.” Despite these
developments in Australian curriculum and teacher education policy supporting
a cross-curricular approach to numeracy, no explicit guidance was provided to
teachers on how to achieve this goal and there are few resources available to teachers
to support them in this approach (Goos, Geiger, & Bennison, 2015).
In Ireland the rationale for embedding numeracy across the curriculum is a response
to the results of the Third International Mathematics and Science Study (Beaton et al.,
1996) and Ireland’s substantial decline in PISA mathematical literacy performance in
2009 (Shiel, Kelleher, McKeown, & Denner, 2016). Performance on these international
assessments, together with the national economic crisis of 2010, provided impetus for
development of a national literacy and numeracy strategy (Department of Education
and Skills, 2011). The government has agreed that all young people in Ireland should
leave school with the appropriate numeracy and literacy skills to live and participate
as informed citizens in society. Initial teacher education programs are now required to
“address student teachers’ literacy and numeracy and their competence in promoting
and assessing literacy and numeracy as appropriate to their curricular/subject area(s)”
(Teaching Council, 2017, p. 14). In addition, a revised curriculum framework for
the lower secondary years has introduced a set of nine Key Skills, one of which is
“being numerate” (Department of Education and Skills, 2015). Teachers are meant
to embed these key skills in the learning outcomes of every subject, but there is not
yet any explanation within newly developed subject specifications of how this can be
done. Thus, in both Australia and Ireland, although new curricula and initial teacher
education standards require all teachers to understand the meaning of numeracy and
to address the numeracy demands of the subjects they teach, little support has been
provided to help teachers fulfil these goals.

Implications of numeracy across the curriculum for teacher learning and


development. Studies have demonstrated that opportunities for attending to

226

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

numeracy across the school curriculum arise in a wide range of subjects, such as
science (Quinnell, Thompson, & LeBard, 2013), economics (O’Neill & Flynn,
2013), and the social sciences (Lake, 2002). These subjects not only demand
quantitative skills but also offer opportunities to develop critical thinking and active
citizenship as important elements of numeracy. It is not intended that teachers in
other subjects should be required to be expert teachers of mathematics. However,
there is an expectation that teachers should be familiar with the inherent numeracy
demands of their subject, can recognise a numeracy opportunity when it arises, and
have the disposition and pedagogical skill to take advantage of such opportunities.
Goos and colleagues have conducted a long-term research program investigating
the effectiveness of a teacher professional learning approach aimed at enhancing
numeracy teaching across a range of school subjects, including history, science,
English, health and physical education, and studies of society and environment. This
program was based on a multi-faceted model of numeracy that represents a synthesis
of research related to effective numeracy practice. The model, which was designed
to support teachers’ planning and reflection, incorporates the dimensions of contexts,
mathematical knowledge, tools, and dispositions, embedded in a critical orientation
to using mathematics. These are summarised in Table 8.1. This model has been used
to identify the numeracy demands of non-mathematics subjects in the Australian
Curriculum, investigate teachers’ understanding of numeracy, and support teachers
to recognise and take advantage of numeracy opportunities in the subjects they teach
(Goos, Dole, & Geiger, 2012; Goos, Geiger, & Dole, 2011, 2014).
By directing teachers’ attention to the knowledge, dispositions, and capacities of
a numerate person, Goos and colleagues found that the model could be used to trace
trajectories of teachers’ learning as they became more familiar with the concept of

Table 8.1. Elements of the numeracy model developed by Goos and colleagues

Mathematical Mathematical concepts and skills; problem solving strategies;


knowledge estimation capacities.
Contexts Capacity to use mathematical knowledge in a range of contexts,
both within schools and beyond school settings
Dispositions Confidence and willingness to use mathematical approaches to
engage with life-related tasks; preparedness to make flexible and
adaptive use of mathematical knowledge.
Tools Use of material (models, measuring instruments), representational
(symbol systems, graphs, maps, diagrams, drawings, tables) and
digital (computers, software, calculators, internet) tools to mediate
and shape thinking
Critical orientation Use of mathematical information to: make decisions and
judgements; add support to arguments; challenge an argument or
position.

227

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

numeracy that the model represents. While most teachers reported becoming more
confident in their professional knowledge of numeracy and their capacity to plan
and implement strategies for exploiting the numeracy demands of the subjects they
taught, few found it easy to take a critical orientation to the use of mathematics
in other curriculum contexts or real world situations (Goos, Geiger, & Dole,
2014). These findings highlight the challenges that teachers experience in not only
developing content knowledge and pedagogical content knowledge for embedding
numeracy into subjects other than mathematics, but also in establishing a new
identity as an “embedder-of-numeracy” alongside their primary, discipline-based
identity as a subject teacher.

EMERGING RESEARCH: DEVELOPING THE PROFESSIONAL KNOWLEDGE


AND IDENTITIES OF NON-SPECIALIST TEACHERS OF MATHEMATICS

Having explored some of the issues and challenges faced by non-specialist teachers
of mathematics, in this section we provide brief snapshots of research that is
emerging in Australia, the United Kingdom, and Ireland with the aim of developing
the professional knowledge and identities of non-specialist teachers who are either
teaching mathematics out-of-field or addressing the numeracy demands of subjects
other than mathematics.

From Knowledge to Identity Development in Teaching Mathematics Out-of-Field

Typically, professional development programs for out-of-field teachers of


secondary school mathematics are directed towards developing teachers’ subject
matter and pedagogical content knowledge. For example, in Australia, Vale et al,
(2011) created and implemented a Mathematics Professional Learning Program
that aimed to enable teachers teaching junior secondary mathematics out-of-field
to teach mathematics at senior secondary school level. This program consisted of
31 three-hour seminars that took place fortnightly in the afternoons partly during
school working hours and partly in the teachers’ own time. The program, which
was collaboratively developed by participants, a deputy principal, and teacher
educators and mathematicians at a university in the State of Victoria, focused on the
core topics of the senior secondary mathematics curriculum. The seminars included
mathematics and professional learning tasks and were structured to first discover the
teachers’ knowledge of a mathematics concept and then develop their conceptual
and procedural understanding. Vale et al. found that their approach to developing
teachers’ knowledge of mathematics needed for teaching at the senior secondary
level deepened and broadened the teachers’ understanding of junior secondary
mathematics content and pedagogy. These teachers learned how to connect their
knowledge of mathematics and mathematics teaching practice with knowledge of
mathematics at the horizon (Ball et al., 2008) through enhanced appreciation of
mathematical structure.

228

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

On a much larger scale, in Ireland the Department of Education and Skills has
funded a Professional Diploma in Mathematics for Teaching – a 2-year part-time
blended learning program for out-of-field teachers of mathematics with a combined
emphasis on subject matter and pedagogical content knowledge (Faulkner, Kenny,
Campbell, & Crisan, 2019). The main stimulus for the program was the report
published by Ní Ríordáin and Hannigan (2009) highlighting that 48% of secondary
mathematics teachers in Ireland were not suitably qualified to teach mathematics
according to national teacher registration requirements. The Professional Diploma
in Mathematics for Teaching is coordinated by a national consortium of higher
education institutions led by the National Centre for STEM Education at the
University of Limerick. By 2019, six cohorts comprising more than 1100 teachers
will have graduated from the Professional Diploma in Mathematics for Teaching
with their tuition fees fully funded by the government. The program content and
structure meet the requirements for preparing fully qualified teachers as specified
by the Teaching Council (described earlier), which regulates the teaching profession
in Ireland. Participants complete five university-level mathematics modules and one
mathematics pedagogy module in each of the two years of the program, with the
former being delivered online and at face-to-face tutorials, and the latter delivered
via week-end workshops at multiple venues throughout the country and a week long
summer school. Participants also complete an action research project in the second
year of the Professional Diploma in Mathematics for Teaching.
To date there has been limited research on the impact of the Professional Diploma
in Mathematics for Teaching, beyond evaluation of participant satisfaction with
the program. However, a study by Ní Ríordáin, Paolucci, and O’Dwyer (2017)
established that participants in the second program cohort demonstrated inadequate
cognitive and conceptual proficiency with curriculum-aligned mathematical content
on entry to the program. Also, of concern was the discrepancy between participants’
proficiency levels and their confidence in teaching the content, with most describing
themselves as being either somewhat or very confident. Thus, this study provided
evidence of the need for a professional development program like the Professional
Diploma in Mathematics for Teaching, which prioritises development of deep
subject matter knowledge in mathematics.
Given the professional development emphasis on developing the subject and
pedagogical content knowledge of out-of-field teachers of mathematics in Ireland,
it is not surprising to observe a similar focus in research related to these initiatives.
However, a doctoral study (Quirke) is currently under way that builds on other
studies that have shifted the focus from knowledge to identity (e.g., Bosse & Törner,
2015), inquiring into the professional identities recognised and enacted by out-
of-field teachers participating in the a Professional Diploma in Mathematics for
Teaching, their current teaching practices, and how they experience the process of
becoming an in-field teacher.
Research conducted in England, in conjunction with a similar program to the
Professional Diploma in Mathematics for Teaching, confirms the need to attend to

229

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

out-of-field teachers’ professional identities as well as their professional knowledge.


In 2008 the Teacher Development Agency commissioned and funded the
Mathematics Development Programme for Teachers to serve out-of-field teachers
who were teaching mathematics to students aged 11 to 19 years (Crisan & Rodd,
2011). The course was free for participants and consisted of 30 university-based
sessions and ten school-based sessions. The purpose of the course was to revisit
and teach school mathematics to enhance the teachers’ technical fluency in the most
challenging topics in secondary school mathematics.
By facilitating Mathematics Development Programme for Teachers programs,
Crisan and Rodd (2017) became aware that issues of identity were significant for out-
of-field mathematics teachers. This awareness emanated from concerns regarding
the disconnection between out-of-field mathematics teachers wanting to be thought
of as expert mathematics teachers yet lacking the requisite knowledge to solve
problems in standard topics on the mathematics curriculum that they were teaching.
Crisan and Rodd developed a Modes of Belonging Mathematics Teacher Identity
Framework, drawing on Wenger’s (1998) notions of engagement, imagination, and
alignment. Using this framework, they investigated the teachers’ transition and
learning in terms of identity formation and transformation as they participated on the
Mathematics Development Programme for Teachers program. This process involved
studying the ways in which the participants spoke of themselves as mathematics
teachers and the changes in their mathematics teaching practice.
The growing significance of identity formation within this field of research
is further evidenced in Hobbs’s (2013) theorising of out-of-field teaching as a
boundary crossing event. Akkerman and Bakker (2011) explain that a boundary
exists where sociocultural differences result in discontinuities in an individual’s
actions and interactions. Thus, Hobbs proposes that a boundary exists for out-of-
field teachers “when the differences between the practices and perspectives required
to teach the subject are ‘discontinuous’” (p. 274). She argues that successfully
negotiating the boundary crossing between in-field and out-of-field spaces provides
opportunities for identity development as out-of-field teachers increase their
knowledge and appreciation of the subject outside their primary area of expertise.
Hobbs formulated a Boundary Between Fields theoretical model to account for
factors that influenced the teachers’ identity construction. The model has three
groups of factors, concerning context, support mechanisms, and personal resources.
Contextual factors included the school’s geographical location, size and design, as
well as governance structures, practices and policies. For example, in isolated rural
schools there were limited opportunities to participate in professional development
or to access subject specialists for advice. Support mechanisms could be provided by
the school or sought out by the teacher to help them adapt to teaching an unfamiliar
subject. Personal resources that teachers brought to the out-of-field experience
included adaptive expertise, knowledge, and dispositions such as confidence and
commitment. Teachers in this study spoke about knowledge as a major factor
influencing their experience and identified all of the knowledge types described

230

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

by Shulman as comprising a teacher’s professional knowledge base (e.g., subject


matter knowledge, pedagogical content knowledge, curricular knowledge). Hobbs
proposed that the Boundary Between Fields model can help raise awareness of
where a discontinuity arises for a teacher crossing boundaries between subjects, and
thus could help inform suitable responses to the problem of teaching out-of-field.

Developing an Identity as an Embedder of Numeracy across the Curriculum

Although the idea of numeracy as a cross-curricular commitment is gaining traction


in several countries, there has been little research into the growth of knowledge
or change in professional identity of teachers who explicitly address the numeracy
demands in the subjects they teach. In Australia, Bennison’s (2016) doctoral study
provides one example of such research that used identity as an analytic lens. The study
identified factors that influence how teachers implement learning from professional
development interventions that support them to promote numeracy learning in
secondary schools. It was conducted in two interrelated phases: a theoretical phase
and an empirical phase. One of the outcomes of the theoretical phase of the study
was a framework for identity as an embedder-of-numeracy (Bennison, 2015, 2017a).
Five Domains of Influence were used to organize the framework: Life History,
Knowledge, Affective, Social, and Context. Factors that have previously been shown
to influence a teacher’s identity were included where it could be argued that these
factors were likely to influence how teachers promote numeracy learning through
the subjects they teach.
The empirical phase of the study was conducted over a two-year period from 2014
to 2015 and employed case study methodology. Participants were eight teachers from
two Australian secondary schools who were recruited from a larger professional
development project that was supporting them to embed numeracy into the subjects
they were teaching: English, history, science, and mathematics. Three of the teachers
were specialists in the subjects they were observed teaching. The qualifications and
teaching assignments of those teaching a subject out-of-field varied. For example,
two teachers were qualified to teach mathematics but were also required to teach
science for which they had no formal qualifications. Another teacher was qualified
to teach physical education and history but was teaching English out-of-field. Data
collected during school visits included interviews and lesson observations.

Life history domain. Past experiences contribute to identity; consequently, many


factors that influence how teachers promote numeracy learning are likely to have
been shaped by their past experiences. The Life History Domain included three
factors: past experiences of mathematics, prospective teacher education, and initial
teaching experiences. Findings of the study pointed to the limited opportunities for
participating teachers, even the most recent graduates, to develop the knowledge
for addressing numeracy across the curriculum during their prospective teacher
education program.

231

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Knowledge domain. A teacher’s knowledge is an important part of his or her


identity (Van Zoest & Bohl, 2005). Shulman’s (1987) three knowledge types
needed for teaching are included in the Knowledge Domain: mathematical content
knowledge, pedagogical content knowledge and curriculum knowledge. Each of
these types of knowledge was defined in a nuanced way in order to relate them
directly related to the knowledge base needed for teachers to promote numeracy
learning. Thus mathematical content knowledge encompasses the level of expertise
in the mathematics that is in inherent in the non-mathematics subjects taught,
pedagogical content knowledge refers to the capacity to design effective numeracy
tasks, and curriculum knowledge includes being able to identify numeracy learning
opportunities in the subjects taught and make connections between numeracy and
subject learning.
An additional type of knowledge, termed subject knowledge, was identified as
necessary for addressing numeracy across the curriculum. Subject knowledge was
seen to encompass content, pedagogical and curriculum knowledge of the subject
in which numeracy was to be embedded – such as history, science, or economics.
Subject knowledge for embedding numeracy is especially salient for out-of-field
teachers: not only are they teaching a subject for which they have limited content
and pedagogical content knowledge, but they experience greater challenges than in-
field teachers in recognizing the numeracy demands of this unfamiliar subject. While
subject knowledge might seem to be encompassed by Shulman’s content knowledge
category, we think it is important to distinguish between content knowledge of
this subject and content knowledge of mathematics that is relevant to the non-
mathematics subject in which the teacher is a specialist.

Affective domain. The Affective Domain initially included personal conception


of numeracy, attitudes towards mathematics, and perceived preparation to embed
numeracy into subjects. Two additional factors emerged from the empirical phase of
the study: motivation to embed numeracy and beliefs about pedagogical approaches
that are possible. Teachers who could be described as embedders (Thornton & Hogan,
2004) see numeracy as enriching understanding in the subjects they teach, and so
feel motivated and efficacious in promoting numeracy learning. In contrast, teachers
who find it difficult to make connections between numeracy and subject learning
may see numeracy as something extra to be added (e.g., Carter, Klenowski, &
Chalmers, 2015).

Social domain. Wenger (1998) proposes that identity development involves


participation in communities, and so teachers’ participation in school and
professional communities contributes to how they promote numeracy learning.
Within school communities, interactions with colleagues and school administrators
related to the meaning of numeracy and who is responsible for promoting numeracy
learning contribute to a teacher’s developing identity as an embedder-of-numeracy.

232

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

Professional communities can offer opportunities for teachers to engage in learning,


and therefore contribute to how they promote numeracy learning.

Context domain. Because practice and identity are related (Wenger, 1998),
affordances and constraints on practice within teachers’ professional contexts can
influence the ways in which they promote numeracy learning. The empirical study
confirmed the importance of the school policy environment and access to appropriate
resources for teaching. The policy environment includes curriculum initiatives and
accountability measures related to numeracy, while resources include access to
representational, physical and digital tools needed to support numeracy learning.
The framework for identity as an embedder-of-numeracy suggests ways of
supporting teachers to embed numeracy across the curriculum by giving attention
to the different domains that influence teachers’ identity formation. Bennison’s
(2016) study uncovered common themes (e.g., lack of opportunities for teachers to
develop pedagogical content knowledge for numeracy) as well as differences (e.g.,
individual interpretations of professional context). The findings of this study in
relation to the knowledge and affective domains could inform the design of teacher
education programs for prospective and practicing teachers. For example, the study
contributed several examples of how numeracy can be enhanced in different school
subjects (e.g., via effective use of scaled timelines in history). While all factors
included in the framework seem to contribute to shaping a teacher’s identity as an
embedder-of-numeracy, a rich personal conception of numeracy and motivation to
embed numeracy into the subjects taught appear to be crucial. Bennison (2017b)
also demonstrated how teachers’ identities as embedders-of-numeracy evolved
over time, thus capturing the dynamic nature of identity as an action rather than an
acquisition.

CONCLUSION

In this chapter we have focused on two areas with a long research history relevant to
mathematics teaching: teacher knowledge and teacher identity. The main theoretical
perspectives illuminating these areas might seem contradictory, being respectively
aligned with cognitive and social theories of learning. However, it is clear from the
literature surveyed in this chapter that knowledge and identity are intertwined.
Knowledge matters, and professional development programs supporting out-of-
field teachers of mathematics typically focus on this requirement. But non-specialist
teachers face the additional challenge of developing a new professional identity,
giving them a sense of belonging to the community of mathematics teachers. Wenger
(1998) describes this transformation as a process of engagement, imagination
and alignment as individuals reconcile their multiple identities of participation in
different communities. Thus, developing a new identity as an in-field teacher of
mathematics, or as an embedder-of-numeracy, does not imply that one must abandon

233

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

one’s original teacher identity as a specialist in another subject. Instead, the task is
one of reconciliation of multiple identities rather than replacement of one identity
by another.
The idea that identities are neither fixed nor stable, advanced by many identity
theorists, is particularly relevant to out-of-field teaching. For example, Hobbs (2012,
2013) captures the dynamic and transformative aspects of identity development in
two frameworks that respectively highlight the aesthetic and boundary-crossing
aspects of teaching out-of-field. Teacher knowledge is a key element of both
frameworks, influencing teachers’ understanding of the subject and their sense of
personal efficacy in teaching it. Similarly, Bennison (2015, 2017a) draws attention to
the multiple identities that teachers must develop as embedders-of-numeracy within
subjects other than mathematics and identifies the different types of professional
knowledge that shape these identities.
In this chapter we have aimed to survey literature related to the development
of professional knowledge and professional identities in non-specialist teachers of
mathematics, and to highlight emerging studies that might point to future research
directions. For example, we need a better understanding of how gains in professional
knowledge, developed in programs such as Ireland’s Professional Diploma in
Mathematics for Teaching, influence out-of-field teachers’ identities as teachers
of mathematics. On the other hand, numeracy-related research that has focused on
teacher identity development could be usefully extended by investigating the forms
of professional knowledge needed to be an effective embedder-of-numeracy into
subjects other than mathematics. Thus, in each of these fields there is a need to
move back and forth between investigations of knowledge and identity, giving due
attention to both conceptualisations of teacher learning.
In the context of this Handbook the chapter is distinctive for its focus on non-
specialist teachers who are either teaching mathematics “out-of-field” or addressing
the numeracy demands inherent in subjects other than mathematics. More
fundamentally, the chapter is concerned with the question of who is responsible for
teaching mathematics in secondary schools, and a pragmatic interest in supporting
all such teachers, especially those who lack formal qualifications in mathematics and
mathematics pedagogy.
Different circumstances have led to the two manifestations of non-specialist
teaching of mathematics discussed in this chapter. First, within any country it
could be argued that the ideal situation would be to have a mathematics teaching
work force in secondary schools that is fully qualified in terms of subject matter
and pedagogical content knowledge, and equitably deployed to meet the learning
needs of all students. The difficulty in realising this ideal is evident from data on
the incidence of out-of-field teaching in many countries, and there are different
interpretations of the causes of this phenomenon ranging from workforce planning
failures to criticisms of education systems that are unable to recruit and retain
teachers in disadvantaged schools. The second type of non-specialist teaching has
arisen largely from international assessments of mathematical literacy, leading to the

234

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

notion that it is the responsibility of all teachers, not only those formally prepared
to teach mathematics, to develop their students’ numeracy competencies across
the whole school curriculum. In each of these circumstances the responsibility for
teaching mathematics, whether as a stand-alone school subject or a competency that
underpins all subjects, often falls to teachers who did not choose to make mathematics
their primary area of disciplinary expertise. We suggest that researchers, policy-
makers, and school leaders ought to engage with this phenomenon to improve our
theoretical understanding of teachers’ needs and to enhance teaching practice for the
benefit of all students.

REFERENCES
Adler, J., & Davis, Z. (2006). Opening another black box: Researching mathematics for teaching in
mathematics teacher education. Journal for Research in Mathematics Education, 37(4), 270–296.
Akkerman, S., & Bakker, A. (2011). Boundary crossing and boundary objects. Review of Educational
Research, 81, 132–169.
Akkerman, S. F., & Meijer, P. C. (2011). A dialogical approach to conceptualizing teacher identity.
Teaching and Teacher Education, 27(2), 308–319.
Australian Institute for Teaching and School Leadership [AITSL]. (2017). Australian professional
standards for teachers. Retrieved from https://www.aitsl.edu.au/teach/standards
Australian Institute for Teaching and School Leadership [AITSL]. (2018). Accreditation of initial teacher
education programs in Australia: Standards and procedures. Retrieved from https://www.aitsl.edu.au/
docs/default-source/default-document-library/accreditation-of-initial-teacher-education-programs-
in-australia_2018-_.pdf?sfvrsn=6ccf23c_2
Australian Curriculum, Assessment and Reporting Authority [ACARA]. (n.d.). Numeracy. Retrieved
from https://www.australiancurriculum.edu.au/f-10-curriculum/general-capabilities/numeracy/
Ball, D., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Beaton, A. E., Mullis, V. S., Martin, M. O., Gonzalez, E. J., Kelly, D. L., & Smith, T. A. (1996).
Mathematics achievement in the middle school years: IEA’s third international mathematics and
science study. Boston, MA: TIMSS International Study Centre.
Beauchamp, C., & Thomas, L. (2009). Understanding teacher identity: An overview of issues in the
literature and implications for teacher education. Cambridge Journal of Education, 39(2), 175–189.
Beijaard, D., Meijer, P. C., & Verloop, N. (2004). Reconsidering research on teachers’ professional
identity. Teaching and Teacher Education, 20(2), 107–128.
Bennison, A. (2015). Developing an analytic lens for investigating identity as an embedder-of-numeracy.
Mathematics Education Research Journal, 27(1), 1–19.
Bennison, A. (2016). Teacher identity as an embedder-of-numeracy: Identifying ways to support teachers
to promote numeracy learning across the curriculum (Unpublished doctoral dissertation). The
University of Queensland, Brisbane, Australia.
Bennison, A. (2017a). Re-examining a framework for teacher identity as embedder-of-numeracy. In A.
Downton, S. Livy, & J. Hall (Eds.), 40 years on: We are still learning! Proceedings of the 40th Annual
Conference of the Mathematics Education Research Group of Australasia (pp. 101–108). Melbourne,
Australia: MERGA.
Bennison, A. (2017b). Understanding the trajectory of a teacher’s identity as an embedder-f-numeracy.
In B. Kaur, W. K. Ho, T. L. To, & B. H. Choy (Eds.), Proceedings of the 41st Conference of the
International Group for the psychology of mathematics education (Vol. 2, pp. 145–152). Singapore:
PME.
Bertram, C., Mthiyane, N., & Mukeredzi, T. (2013). ‘It will make me a real teacher’: Learning experiences
of part time PGCE students in South Africa. International Journal of Educational Development,
33(5), 448–456.

235

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Bosse, M., & Törner, G. (2015). Teacher identity as a theoretical framework for researching out-of-field
teaching mathematics teachers. In C. Bernack-Schüler, R. Erens, T. Leuders, & A. Eichler (Eds.), Views
and beliefs in mathematics education. Freiburger Empirische Forschung in der Mathematikdidaktik
(pp. 1–13). Wiesbaden: Springer Fachmedien Wiesbaden.
Canrinus, E. T., Helms-Lorenz, M., Beijaard, D., Buitink, J., & Hofman, A. (2011). Profiling teachers’
sense of professional identity. Educational Studies, 37(5), 593–608.
Carter, M., Klenowski, V., & Chalmers, C. (2015). Challenges in embedding numeracy throughout the
curriculum in three Queensland secondary schools. Australian Educational Researcher, 42, 595–611.
Centre for Development and Enterprise. (2015). Teacher evaluation in South African schools.
Johannesburg: CDE.
Chronaki, A. (2013). Identity work as a political space for change: The case of mathematics teaching
through technology use. In M. Berger, K. Brodie, V. Frith, & K. le Roux (Eds.), Proceedings of the
Seventh International Mathematics Education and Society Conference (Vol. 1, pp. 1–18). Cape Town:
MES.
Côté, J. (2006). Identity studies: How close are we to developing a social science of identity? An appraisal
of the field. Identity, 6(1), 3–25.
Council of Australian Governments. (2008). National numeracy review report. Retrieved from http://
webarchive.nla.gov.au/gov/20080718164654/http://www.coag.gov.au/reports/index.htm#numeracy
Crisan, C., & Rodd, M. (2011). Teachers of mathematics to mathematics teachers: A TDA mathematics
development programme for teachers. In C. Smith (Ed.), Proceedings of the Conference of the British
Society for Research into Learning Mathematics (Vol. 31, Issue. 3, pp. 29–34). Oxford: BSRLM.
Crisan, C., & Rodd, M. (2017). Learning mathematics for teaching mathematics: Non-specialist teachers’
mathematics teacher identity. Mathematics Teacher Education and Development, 19(2), 104–122.
Darragh, L. (2016). Identity research in mathematics education. Educational Studies in Mathematics,
93(1), 19–33.
Department for Education. (2017). School workforce in England: November 2016. London: Department
for Education.
Department of Education and Skills. (2011). Literacy and numeracy learning for life: The national
strategy to improve literacy and numeracy among children and young people 2011–2020. Dublin:
Author. Retrieved from https://www.education.ie/en/Publications/Policy-Reports/lit_num_strategy_
full.pdf
Department of Education and Skills. (2015). Framework for junior cycle 2015. Dublin: Author. Retrieved
from https://www.ncca.ie/media/3249/framework-for-junior-cycle-2015-en.pdf
Donaldson, M. (2013). Principals’ approaches to cultivating teacher effectiveness: Constraints and
opportunities in hiring, assigning, evaluating, and developing teachers. Educational Administration
Quarterly, 49(5), 838–882.
du Plessis, A. E. (2015). Effective education: Conceptualising the meaning of out-of-field teaching
practices for teachers, teacher quality and school leaders. International Journal of Educational
Research, 72(Supplement C), 89–102.
du Plessis, A. E. (2016). Leading teachers through the storm: Looking beyond the numbers and turning
the implications of out-of-field teaching practices into positive challenges. International Journal of
Educational Research, 79, 42–51.
du Plessis, A. E. (2017). Out-of-field teaching practices: What educational leaders need to know.
Rotterdam, The Netherlands: Sense Publishers.
du Plessis, A. E., Gillies, R. M., & Carroll, A. (2014). Out-of-field teaching and professional development:
A transnational investigation across Australia and South Africa. International Journal of Educational
Research, 66, 90–102.
du Plessis, A., Hobbs, L., Luft, J., & Vale, C. (2019). The out-of-field teacher in context: The impact
of the school context and environment. In L. Hobbs & G. Törner (Eds.), International perspectives
on teaching across specialisations: Experiences, practices, and contexts of out-of-field teachers
(pp. 217–242). Cham: Springer.
Enyedy, N., Goldberg, J., & Welsh, K. M. (2005). Complex dilemmas of identity and practice. Science
Education, 90, 68–93.

236

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

Ernest, P. (1989). The knowledge, beliefs and attitudes of the mathematics teacher: A model. Journal of
Education for Teaching, 15(1), 13–33.
Faulkner, F., Kenny, J., Campbell, C., & Crisan, C. (2019). Teacher learning and continuing professional
development. In L. Hobbs & G. Törner (Eds.), Examining the phenomenon of “teaching out-of-field”:
International perspectives on teaching as a non-specialist (pp. 269–308). Singapore: Springer.
Fennema, E., & Franke, M. L. (1992). Teachers’ knowledge and its impact. In D. A. Grouws (Ed.),
Handbook of research on mathematics teaching and learning (pp. 147–164). New York, NY:
Macmillan.
Gee, J. P. (2001). Identity as an analytic lens for research in education. Review of Research in Education,
25, 99–125.
Goos, M., Dole, S., & Geiger, V. (2012). Auditing the numeracy demands of the Australian Curriculum.
In J. Dindyal, L. Chen, & S. F. Ng (Eds.), Mathematics education: Expanding horizons. Proceedings
of the 35th Annual Conference of the Mathematics Education Research Group of Australasia
(pp. 314–321). Singapore: MERGA.
Goos, M., Geiger, V., & Bennison, A. (2015). Conceptualising and enacting numeracy across the
curriculum. In K. Beswick, T. Muir, & J. Wells (Eds.), Proceedings of the 39th Annual Conference
of the International Group for the psychology of mathematics education (Vol. 3, pp. 9–16). Hobart,
Australia: PME.
Goos, M., Geiger, V., & Dole, S. (2011). Teachers’ personal conceptions of numeracy. In B. Ubuz (Ed.),
Proceedings of the 35th Conference of the International Group for the psychology of mathematics
education (Vol. 2, pp. 457–464). Ankara, Turkey: PME.
Goos, M., Geiger, V., & Dole, S. (2014). Transforming professional practice in numeracy teaching. In
Y. Li, E. Silver, & S. Li (Eds.), Transforming mathematics instruction: Multiple approaches and
practices (pp. 81–102). New York, NY: Springer.
Graven, M. (2004). Investigating mathematics teachers’ learning within an in-service community of
practice: The centrality of confidence. Educational Studies in Mathematics, 57(2), 177–211.
Graven, M., & Lerman, S. (2014). Mathematics teacher identity. In S. Lerman (Ed.), Encyclopedia of
mathematics education (pp. 434–438). New York, NY: Springer.
Grootenboer, P. J., & Zevenbergen, R. (2008). Identity as a lens to understand learning mathematics:
Developing a model. In M. Goos, R. Brown, & K. Makar (Eds.), Navigating currents and charting
directions. Proceedings of the 31st Annual Conference of the Mathematics Education Research Group
of Australasia (pp. 243–250). Brisbane, Australia: MERGA.
Handal, B., Watson, K., Petocz, P., & Maher, M. (2013). Retaining mathematics and science teachers in
rural and remote schools. Australian and International Journal of Rural Education, 23(3), 13–27.
Heyd-Metzuyamin, E., Lutovac, S., & Kaasila, R. (2016). Identity. In M. Hannula, P. Di Martino,
M. Pantziara, Q. Zhang, F. Morselli, E. Heyd-Metzuyamin, S. Lutovac, R. Kaasila, J. Middleton,
A. Jansen, & G. Goldin (Eds.), Attitudes, beliefs, motivation and identity in mathematics education:
An overview of the field and future directions (pp. 14–17). Cham: Springer.
Hobbs, L. (2012). Examining the aesthetic dimensions of teaching: Relationships between teacher
knowledge, identity and passion. Teaching and Teacher Education, 28(5), 718–727.
Hobbs, L. (2013). Teaching ‘out-of-field’ as a boundary-crossing event: Factors shaping teacher identity.
International Journal of Science and Mathematics Education, 11(2), 271–297.
Hoffman, D. M. (1998). A therapeutic moment? Identity, self, and culture in the anthropology of
education. Anthropology and Education Quarterly, 29(3), 324–346.
Holland, D., Lachicotte, W. J., Skinner, D., & Cain, C. (1998). Identity and agency in cultural worlds.
Cambridge, MA: Harvard University Press.
Ingersoll, R. M. (2002). Out-of-field teaching, educational inequality and the organization of schools:
An exploratory analysis. Seattle, WA: University of Washington: Center for the Study of Teaching
and Policy.
Ingersoll, R. M. (2011). Do we produce enough mathematics and science teachers? Phi Delta Kappan,
92(6), 37–41.
Lake, D. (2002). Critical social numeracy. The Social Studies, 93(1), 4–10.

237

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Lerman, S. (2000). The social turn in mathematics education research. In J. Boaler (Ed.), Multiple
perspectives on mathematics teaching and learning (pp. 19–44). Westport, CT: Ablex Publishing.
Lin, F.-L., & Cooney, T. (Eds.). (2001). Making sense of mathematics teacher education. Dordrecht:
Kluwer Academic Publishing.
Loveys, K. (2011, April 21). Scandal of the untrained teachers: Thousands don’t have degrees in the
subjects they teach. The Daily Mail. Retrieved from http://www.dailymail.co.uk/news/article-1378908/
Thousands-teachers-dont-degrees-subjects-teach.html
Marginson, S., Tytler, R., Freeman, B., & Roberts, K. (2013). STEM: Country comparisons: International
comparisons of Science, Technology, Engineering and Mathematics (STEM) education (Final Report).
Melbourne: Australian Council of Learned Academies.
Masters, G. (2015). Planning a stronger workforce, Teacher: Evidence + Insight + Action. Camberwell:
Australian Council for Educational Research. Retrieved from http://www.teachermagazine.com.au/
geoff-masters/article/planning-a-stronger-teacher-workforce
Mayer, D., Doecke, B., Ho, P., Kline, J., Kostogriz, A., Moss, J., … Hodder, P. (2014). Longitudinal
teacher education and workforce study (Final Report, November, 2013). Canberra: Department of
Education, Commonwealth of Australia. Retrieved from https://docs.education.gov.au/system/files/
doc/ other/ltews_main_report.pdf
McConney, A., & Price, A. (2009). Teaching out-of-field in Western Australia. Australian Journal of
Teacher Education, 34(6), 86–100.
Ministerial Council on Education, Employment, Training and Youth Affairs (MCEETYA). (2008).
Melbourne declaration on educational goals for young Australians. Retrieved from
http://www.curriculum.edu.au/verve/_resources/National_Declaration_on_the_Educational_Goals_
for_Young_Australians.pdf
Mockler, N. (2011). Beyond ‘what works’: Understanding teacher identity as a practical and political tool.
Teachers and Teaching, 17(5), 517–528.
Mullis, I. V. S., Martin, M. O., Foy, P., & Arora, A., (2012). TIMSS 2011. International results in
mathematics. Chestnut Hill, MA: TIMSS and PIRLS International Study Centre, Boston College.
Mullis, I. V. S., Martin, M. O., Foy, P., & Hooper, M. (2016). TIMSS 2015. International results in
mathematics. Retrieved from http://timssandpirls.bc.edu/timss2015/ international-results
Ní Ríordáin, M., & Hannigan, A. (2009). Out-of-field teaching in post-primary mathematics education:
An analysis of the Irish context. Limerick: NCE-MSTL. Retrieved from http://epistem.ie/wp-content/
uploads/2015/04/Out-of-field-teaching-in-post-primary-Maths-Education.pdf
Ní Ríordáin, M., Paolucci, C., & O’Dwyer, L. (2017). An examination of the professional development
needs of out-of-field mathematics teachers. Teaching and Teacher Education, 64, 162–174.
Ogilvie, C., & O’Brien, L. (2015). Unqualified teachers. Retrieved from https://fullfact.org/education/
unqualified-teachers/
O’Meara, N. (2011). Improving mathematics teaching at second level through the design of a model
of teacher knowledge and an intervention aimed at developing teachers’ knowledge (Unpublished
doctoral dissertation). University of Limerick, Limerick, Ireland.
O’Neill, P. B., & Flynn, D. T. (2013). Another curriculum requirement? Quantitative reasoning in
economics: Some first steps. American Journal of Business Education, 6(3), 339–346.
Price, A. (2017). An overview of the phenomenon of teaching out of field. Unterrichtswissenschaft,
45(2), 102–113.
Price, A., Vale, C., Porsch, R., Esti, R., Faulkner, F., Ni Riordain, M., … Luft, J. (2019). Teaching out-
of-field internationally. In L. Hobbs & G. Törner (Eds.), Examining the phenomenon of “teaching
out-of-field”: International perspectives on teaching as a non-specialist (pp. 53–83). Cham: Springer.
Quinnell, R., Thompson, R., & LeBard, R. J. (2013). It’s not maths; it’s science: Exploring thinking
dispositions, learning thresholds and mindfulness in science learning. International Journal of
Mathematical Education in Science and Technology, 44(6), 808–816.
Rowland, T., Huckstep, P., & Thwaites, A. (2005). Elementary teachers’ mathematics subject knowledge:
The Knowledge Quartet and the case of Naomi. Journal of Mathematics Teacher Education, 8(3),
255–281.

238

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PROFESSIONAL KNOWLEDGE AND IDENTITIES

Sachs, J. (2005). Teacher education and the development of professional identity. In P. Denicolo &
M. Kompf (Eds.), Connecting policy and practice: Challenges for teaching and learning in schools
and universities (pp. 5–21). New York, NY: Routledge Farmer.
Sfard, A., & Prusak, A. (2005). Telling identities: In search of an analytic tool for investigating learning
as a culturally shaped activity. Educational Researcher, 34(4), 14–22.
Sharplin, E. D. (2014). Reconceptualising out-of-field teaching: Experiences of rural teachers in Western
Australia. Educational Research, 56(1), 97–110.
Shiel, G., Kelleher, C., McKeown, C., & Denner, S. (2016). Future ready? The performance of 15-year-
olds in Ireland on science, reading literacy and mathematics in PISA 2015. Dublin: Educational
Research Centre.
Shulman, L. S. (1986). Those who understand: Knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Shulman, L. S. (1987). Knowledge and teaching: Foundations of the new reform. Harvard Educational
Review, 57(1), 1–23.
Steen, L. (2001). The case for quantitative literacy. In L. Steen (Ed.), Mathematics and democracy:
The case for quantitative literacy (pp. 1–22). Princeton, NJ: National Council on Education and the
Disciplines.
Steyn, G. M., & du Plessis, A. E. (2007). The implications of the out-of-field phenomenon for effective
teaching, quality education and school management. Africa Education Review, 4(2), 144–158.
Stryker, S. (1980). Symbolic interactionism: A social structural version. Menlo Park, CA: Benjamin
Cummings Publishing.
Teaching Council. (2013). Teaching council registration: Curricular subject requirements (post-primary)
for persons applying for registration on or after 1 January 2017. Retrieved November 9, 2018,
from https://www.teachingcouncil.ie/en/Publications/ Registration/Documents/Curricular-Subject-
Requirments-after-January-2017.pdf
Teaching Council. (2017). Initial teacher education: Criteria and guidelines for programme providers.
Maynooth: Author. Retrieved 2 November 2018 from https://www.teachingcouncil.ie/en/Publications/
Teacher-Education/Initial-Teacher-Education-Criteria-and-Guidelines-for-Programme-Providers.pdf
Thornton, S., & Hogan, J. (2004). Orientations to numeracy teachers’ confidence and disposition to use
mathematics across the curriculum. In M. J. Høines & A. B. Fugelstad (Eds.), Proceedings of the
28th conference of the International Group for the psychology of mathematics education (Vol. 4, pp.
313–320). Bergen, Norway: PME.
Törner, G. (2014). Underqualified mathematics teachers or out-of-field-teaching in mathematics
in Germany – The cultural framework. In L. Hobbs & M. Bosse (Eds.), Taking an international
perspective on “out-of-field” teaching: Proceedings and agenda for research and action from the
1st Teaching Across Specialisations (TAS) Collective Symposium (pp. 15–16). Porto, Portugal: TAS
Collective.
Törner, G., & Törner, A. (2012). Underqualified math teachers or out-of-field teaching in mathematics –A
neglectable field of action? In W. Blum, R. Borromeo Ferii, & K. Maaß (Eds.), Mathematikunterricht
im kontext von realitat, kultur und lehrerprofessionalitat (pp. 196–206). Wiesbaden, Germany:
Springer Spektrum.
Vale, C., & Drake, P. (2019). Attending to out-of-field teaching: Implications of and for education policy.
In L. Hobbs & G. Törner (Eds.), International perspectives on teaching across specialisations:
Experiences, practices, and contexts of out-of-field teachers (pp. 195–215). Cham: Springer.
Vale, C., McAndrew, A., & Krishnan, S. (2011). Connecting with the horizon: Developing teachers’
appreciation of mathematical structure. Journal of Mathematics Teacher Education, 14(3), 193–212.
van Putten, S., Stols, G., & Howie, S. (2014). Do prospective mathematics teachers teach who they say
they are? Journal of Mathematics Teacher Education, 17(4), 369–392.
Van Zoest, L. R., & Bohl, J. V. (2005). Mathematics teacher identity: A framework for understanding
secondary school mathematics teachers’ learning through practice. Teacher Development, 9(3),
315–345.
Weigert, A. J., & Gecas, V. (2005). Symbolic interactionist reflections on Erikson, identity, and
postmodernism. Identity, 5(2), 161–174.

239

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MERRILYN GOOS ET AL.

Weldon, P. (2016). Out-of-field teaching in secondary schools. Policy Insights, Issue 6. Camberwell:
Australian Council for Educational Research.
Wenger, E. (1998). Communities of practice: Learning, meaning, and identity. Cambridge: Cambridge
University Press.
Williams, J. (2011). Teachers telling tales: The narrative mediation of professional identity. Research in
Mathematics Education, 13(2), 131–142.

Merrilyn Goos
EPI*STEM – National Centre for STEM Education
University of Limerick

Anne Bennison
School of Education
University of the Sunshine Coast

Stephen Quirke
EPI*STEM – National Centre for STEM Education
University of Limerick

Niamh O’Meara
EPI*STEM – National Centre for STEM Education
University of Limerick

Colleen Vale
Faculty of Education
Monash University

240

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PART 4
MATHEMATICS TEACHING AND ITS
DEVELOPMENT

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL, ROBERTA HUNTER AND SANDRA CRESPO

9. MATHEMATICS TEACHERS COMMITTED


TO EQUITY
A Review of Teaching Practices

This chapter provides a selective review of the research literature on mathematics


teaching that attends to and addresses inequities in the learning opportunities offered
to students. First, we trace and discuss the growing focus on educational equity in
the teaching of mathematics. We then describe the challenges and possibilities for
a few prominent equity-based approaches to mathematics teaching (e.g., culturally
relevant mathematics teaching, teaching mathematics for social justice, and complex
instruction). We also discuss the demands of these equity-based approaches on
teachers’ practices.
Everything has been said before, but since nobody listens we have to keep
going back and beginning all over again.
–André Gide (1891)

INTRODUCTION

Since the previous handbook, educational equity has become a more central concern
for research in mathematics education, which has led to an increased emphasis
on equity for teacher education (Cai, 2017; Frade, Acioly-Régnier, & Jun, 2013;
Healy & Powell, 2013; Jablonka, Wagner, & Walshaw, 2013; Meaney & Lange,
2013; Planas & Valero, 2016). To summarize the large body of research on equity
in mathematics education is beyond the scope of this chapter. What we provide is
a brief overview of the “equity journey.” Several handbook chapters that provide
reviews of equity related topics point to a journey that has developed in parallel
to the dominant narrative of mathematics education, from individual/psychological
to constructivism/social constructivism, where the emphasis is on developing
conceptual understanding and students taking ownership of their learning. In this
journey, equity is first understood as all students having access to rich mathematics
and being provided opportunities to construct their understandings. Then we have
the socio-cultural perspective, which pays attention to students’ experiences, their
funds of knowledge, and aims to develop learning that builds on students’ cultural
backgrounds and knowledge. Continuing the journey, the socio-political perspective
takes us into the wider context and brings particular attention to issues of power.

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_010

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

This journey is not to be interpreted as a linear trajectory from, for example, social
constructivism to socio-political views. While there is a sense of timeline, with the
former one having been in place longer than the more recent socio-political approach,
these three main areas co-exist and should not be seen in any kind of hierarchy.
Our review in this chapter follows this journey from an emphasis on issues related
to equalizing learning opportunities to students through emphasis on distributing
more fairly opportunities for conceptual understanding and classroom participation
through collaborative learning, as one view of equity, what we call in this chapter –
Inclusive Mathematics Teaching; to a view where students’ cultural backgrounds/
funds of knowledge become more central, Culturally Sustaining Mathematics
Teaching, and we end with a view that draws on the sociopolitical turn, Social
Justice Mathematics Teaching. Again, this is by no means a linear progression, nor
is there always a clear separation among these three broad perspectives on equity in
mathematics education.
Next, we describe the method for our review; we then elaborate on the three
broad perspectives on equitable mathematics teaching practices that emerged from
our review – Inclusive Mathematics Teaching, Culturally Sustaining Mathematics
Teaching, and Social Justice Mathematics Teaching. We then move to discussing
three examples of how we, the authors, have engaged with teachers around these
broad perspectives on equitable mathematics teaching. We chose to focus on these
long-term examples from our own work because they have a solid research basis
and a strong theoretical and practical emphasis. However, it is important to note the
way in which these examples interweave these broad perspectives as opposed to
representing each of them separately. We close with a discussion on challenges and
new directions for research on mathematics teaching for equity.

METHOD FOR THE REVIEW

We started reading some key chapters in prior handbooks (Forgasz & Leder, 2008;
Frade, Acioly-Régnier, & Jun, 2013; Healy & Powell, 2013; Jablonka et al., 2013;
Meaney & Lange, 2013; Planas & Valero, 2016) to get the lay of the land and used
this information to frame this chapter. The topic of equity in mathematics education
is very broad, so we had to make some decisions that would help us determine the
focus for this chapter. The handbook chapters led us to three general approaches
to equity-focused mathematics teaching, which we have labeled as Inclusive
Mathematics Teaching, Culturally Sustaining Mathematics Teaching, and Social
Justice Mathematics Teaching. We then decided on a main definition of equity that
would guide our search: equity as opportunity to participate. Viewing learning as
participation requires teachers to move away from:
students’ cognitive differences/performances (conveyed by expressions such as
“good student” and “weak student”) towards the students’ semiotic interactions
within mathematical practices or activities, which are situated in a broader

244

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

historical/socio-cultural context. This does not imply that students’ individual


needs are ignored. On the contrary, cognition viewed in terms of symbolic
mediation means that different individuals will interact in semiotically different
ways. In producing different meanings, teachers should be aware that students
will have distinct needs that must be considered by the practice. (Frade et al.,
2013, p. 116)
Equity as opportunity to participate served as a broad lens by which we selected
relevant articles. That is, we looked for articles that focused on teachers working
towards increasing the participation in the mathematics classroom of non-dominant
students. We understand that there are different ways of naming the inequities and
power dynamics that are at play in all social interactions. We will use the terms non-
dominant or minoritized throughout this chapter to refer to students (and communities)
who have traditionally been underrepresented or marginalized. We also looked for
articles that would inform us about practices of teachers as they were engaged in
developing equitable learning environments. We limited our search to studies with
practicing teachers (not with prospective teachers) and published since 2008 (the year
of the previous International Handbook on Mathematics Teacher Education).
Our approach was to read widely from many sources including research journals
such as the Journal of Mathematics Teacher Education, Mathematics Education
Research Journal, Journal for Research in Mathematics Education, Educational
Studies in Mathematics, Journal of Urban Mathematics Education, and Mathematical
Thinking and Learning. We also looked at proceedings of the Mathematics Education
and Society (MES), Mathematics Education Research Group of Australasia
(MERGA), and Psychology of Mathematics Education (PME). Finally, we looked at
recently published books that are specific to the topic of this chapter (e.g., Celedón-
Pattichis, White, & Civil, 2017; Crespo, Celedón-Pattichis, & Civil, 2018; Fernandes,
Crespo, & Civil, 2017; Wager & Stinson, 2012; White, Fernandes, & Civil, 2018).
We tried to get a broad geographic representation beyond the typical Western world
English-dominant countries, by looking at different outlets, including journals
written in Spanish (e.g., RELIME (Revista Latinoamericana de Investigación en
Matemática Educativa); Revista Latinoamericana de Etnomatemática) to reach into
the Latin American countries. We come back to this point in our conclusion.
One further consideration for this review was to refrain from perpetuating the
negative discourse about teachers and students of mathematics. Our focus in this
chapter therefore is to move away from deficit views of non-dominant students
and communities and focus primarily on strength-based socio-cultural and socio-
political approaches to the teaching and learning of mathematics and in particular
on the implications for teaching and practices. To this end, we also wanted to move
away from deficit views of teachers as they try to implement these different equity-
based approaches to teaching mathematics. So, while we acknowledge the tensions
and challenges that teachers may encounter, we have largely focused on studies
where teachers were trying to implement these approaches.

245

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

INCLUSIVE MATHEMATICS TEACHING

A serious concern of mathematics teaching has been to provide equal access to


high quality instruction to a classroom full of students with diverse backgrounds,
experiences and needs. In this section we focus on research studies that have
examined teaching strategies that attend to how the language of instruction and
classroom participation structures exclude and include students in the mathematics
classroom. We focus on studies that discuss approaches and practices of teachers
with an orientation to what we are calling inclusive mathematics teaching. By this
we mean teaching that recognizes that in mathematics classrooms some students
learn more and some students learn less (some might say they are cheated out of a
good education), not simply because of their personal characteristics but because of
unequal access to the learning opportunities offered in the classroom. We are using
“inclusive mathematics teaching” because we found multiple terms that were too
specific to a particular country or region and not used internationally, hence our
decision to use this umbrella term.
Although equity may not have been a dominant lens in the 1960s and 1970s,
attention to gender differences and unfairness for girls in mathematics classrooms
was a major concern during that time period. We can trace interest and attention
of mathematics education researchers and practitioners to inclusive mathematics
teaching as the field began to interrogate disparities in the learning outcomes of girls
as compared to boys, and began searching for possible explanations that challenged
deficit views of women in mathematics (Leder, 1992; Sadker, Sadker, & Zittleman,
2009). We can also connect this early interest on inclusive mathematics teaching to
research with a focus on teachers’ behaviors (rather than their cognition) or what
Stinson and Bullock (2012) call historical moments in mathematics education
research, in particular to studies with a focus on of relations between teacher
expectations and students’ unequal outcomes (Oakes, 1990, 1992).
We can also find traces of equity concerns in one of the most prolific areas of
educational research that still continues today – the research on classroom talk, in
particular, teacher questioning patterns and how these are distributed in the classroom
(e.g., Brophy & Good, 1986). Research on classroom talk has consistently reported that
the quantity and quality of talk matters for student learning – the more students talk, the
more they learn (O’Connor, 1998). Yet classroom talk is often unequally distributed
among students (Sfard & Kieran, 2001; Shah, Lewis, & Caires, 2014). In fact, more
often than not, it is the students who most need opportunities to talk and share their
ideas that are marginalized and excluded in the mathematics classroom (Crespo &
Featherstone, 2012, Featherstone et al., 2011). These patterns of inequitable classroom
interactions happen in both small and large group class discussions. Concerns about
language demands and participation opportunities in mathematics classrooms have
also and more recently been extended to multilingual students and students whose
home language differ from the dominant language of instruction used in schools,
which tends to in many parts of the world favor white middle-class students.

246

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Louie’s (2017) characterization of exclusionary and inclusionary teaching


practices helps to map out important distinction between mathematics teaching
practices that attend to and neglect concerns for inclusive mathematics education
as we have defined it for the purposes of this chapter. Table 9.1 is a reproduction

Table 9.1. Characterization of exclusionary/inclusive mathematics teaching practices


(Louie, 2017, p. 496)
Exclusionary Inclusive

The rote practice frame The sense-making frame


Mathematics is a fixed body of Mathematics is about making sense of
knowledge to be absorbed and ideas and understanding connections.
practiced. Correctness is paramount.
‡ Assigning open-ended, nonroutine
‡ Presenting standard formulas, task
The nature of mathematical activity

algorithms, and so forth ‡ Asking open-ended questions and


‡ Assigning routine task requiring pressing for meaning in conversation
only the application of previously with students
demonstrated algorithms ‡ Explicitly stating the importance of
‡ Asking closed questions in sense making
conversation with students
The multidimensional frame
‡ Explicitly stating the importance of
repetitive practice Mathematics includes activities such
‡ Focusing discussion exclusively on as collaboration, experimentation, and
answers argumentation, not just rote practice.
‡ Assigning open-ended, nonroutine
task
‡ Explicitly naming skills that have
not traditionally been seen as
mathematical as mathematically
important

The hierarchical ability frame The multidimensional ability frame


The nature of mathematical ability

Mathematics ability is distributed Everyone has both intellectual strengths


along a linear continuum. Some people and areas for growth that are relevant to
have a lot; others have very little. mathematics learning.
‡ Explicitly valorizing speed and ‡ Valorizing skills that have
correctness not traditionally been seen as
‡ Positioning some students as helpers mathematical
and others as in need of help ‡ Naming a variety of students as
resources for their peers’ learning
‡ Making statements about mutual
dependence (everyone contributes,
everyone learns together)

247

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

of Louie’s characterization of instructional practices that operate with either an


exclusionary or inclusive frame of mathematics education.
Noteworthy here is the point we made earlier that attention to deep understanding
of mathematics has always been a concern of mathematics education research
and that a sense-making frame (as Louie calls it) is rooted in the field’s collective
commitment to developing students’ self-confidence and ownership of their
mathematical ideas, such as the explicitly naming of mathematical skills that have
not been traditionally seen as mathematically important. Shifting the authority of
who knows and produces mathematical knowledge from the teacher to the student
has been a central concern in much the research in mathematics education. However,
a frame for inclusive mathematics teaching further considers teaching practices
that communicate to students that their peers are valuable resources for learning
mathematics. Teachers committed to inclusive mathematics education organize their
classrooms and make teaching moves that valorize productive learning interactions
among peers let those be during whole class or small group activities.
A survey of the research on inclusive mathematics teaching suggests that much
of the focus is on documenting the work that needs to happen in order to support
teachers to transform their teaching from an exclusionary to an inclusionary form
of mathematics instruction. Meaning that much of the focus of the research has
been on individual teachers’ teaching practices and on how these practices impact
the learning of individual students. However, more contemporary approaches to
inclusive mathematics teaching research focus on better understanding the classroom
culture that teachers committed to equity need to create and to systemic inequities,
such as the practice of tracking (streaming or ability grouping) and accompanying
teaching practices that further disadvantage students from historically marginalized
groups based on race, language, immigration status, (dis)ability, and other markers
of difference.
Although the focus on equity has become more explicitly prominent in the past
five years in the United States and around the world, the number of publications
on this topic published in mainstream international journals, such as the Journal of
Mathematics Teacher Education are still low. A search of articles published between
2008 and 2018 yielded a total of four articles with an explicit focus on supporting
teachers to address issues of exclusion of students’ opportunities to participate in
mathematics classrooms. Staples (2008) discusses the successes and challenges
of using group work as an approach to detracking mathematics classrooms and
supporting students’ mathematical success in more inclusive classrooms. Bobis,
Way, Anderson, and Martin (2016) report on professional development focused
on challenging teachers’ beliefs about student engagement in mathematics and the
shifts in their instructional practice that were made possible by this intervention.
Ross (2014) and Owens (2015) focus on professional learning of teachers who work
with minoritized students. Ross (2014) discusses the role of teacher self-efficacy
and the need for teachers’ continued professional learning to become effective
teachers of English Language Learners. Owens (2015) focuses on improving

248

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

the teaching of mathematics for indigenous students in rural Australia through


professional development and establishing partnerships between the school and
the students’ community. Overall, these four studies provide a general idea of the
ongoing conversation of mathematics education researchers and teacher educators
with a focus on inclusive mathematics teaching, and highlight a common concern
for how to support teachers to change and to sustain a commitment to instructional
practices that are consistent with principle of educational equity in their mathematics
classrooms.
Later in the chapter we discuss further the practices of teachers committed to
inclusive mathematics teaching with the example of complex instruction which is an
approach that considers how to equalize students’ participation. In the next section
we turn our attention to research studies which have examined the teaching of
mathematics by teachers within a culturally responsive and sustaining perspective.
Our specific interest is placed on how the teacher’s practices play out as they enact
classrooms which show key elements of what we have termed Culturally Sustaining
Mathematics Teaching. We begin by looking at the genesis of this term we have
adopted and what the literature outlines as required teacher practices which uphold it.

CULTURALLY SUSTAINING MATHEMATICS TEACHING

Culturally sustaining mathematics teaching fits within what multiculturalists (e.g.,


Banks & Banks, 2009; Nieto & Bode, 2011) over time have advocated for – an
education system that is both equitable and culturally diverse – one which includes
culturally responsive curricula, appropriate pedagogical and assessment practices,
and which is organized to facilitate interactions across racial and ethnic lines, so that
all students have equitable opportunities to learn and achieve (Johnson, 2014). The
term Culturally Sustaining Mathematics Teaching which we use has its beginnings
in the seminal work of Ladson-Billings (1994, 1995) in which she coined the phrase
culturally relevant pedagogy. Gay (2002, 2010) in building on Ladson-Billing’s work
argued that the defining qualities of culturally responsive teaching, are teachers
developing a knowledge base about cultural diversity; using the cultural
characteristics, experiences, and perspectives of ethnically diverse students
in the curriculum; demonstrating culturally sensitive caring and developing
learning communities; using effective cross-cultural communication; and
responding to ethnic diversity in the delivery of instruction. (Johnson, 2014,
p. 146)
Further, Gay and Kirkland (2003) and Villegas and Lucas (2001) draw attention to
teachers’ need to have a sociopolitical consciousness with which they consciously
appraise dominant power structures and support their students to do likewise.
More recently, the term culturally sustaining pedagogy has been adopted to
encompass a more dynamic and changing view of culture (Paris, 2012, 2015). Paris
and Alim (2014) explain that within this asset-based perspective, culturally relevant

249

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

pedagogy is reframed towards one in which its specific goal is to “perpetrate and
foster – to sustain – linguistic, literate and cultural pluralism as part of the democratic
project of schooling and as a needed response to demographic and social change”
(p. 88). This requires an openness to sustain languages and cultures not only in
traditional but also in more recent and evolving ways. We use Culturally Sustaining
Mathematics Teaching akin to how Ladson-Billings (2000) considered culturally
relevant pedagogy, as “a pedagogy of opposition not unlike critical pedagogy”
(p. 208). Inherent in culturally sustaining mathematics teaching are notions of
empowerment not only of individual students but also of the empowerment of the
group as a collective within the mathematics classroom.
As can be seen by what we have written, to build a picture of culturally sustaining
mathematics teaching has necessitated us drawing together landmark studies from
the past three decades. However, these have most often not had an explicit focus
on mathematics teaching and learning. Many of the studies are broadly theoretical
in nature, or they have focused on exceptional teachers within specific settings,
and with particular student populations (Bonner, 2014). Similarly, they are situated
in classrooms and with teachers within classrooms in the developed world rather
than within the developing world. We turn our attention now to explore studies
which are situated within the mathematics field and which illustrate the practices
of teachers using culturally sustaining mathematics teaching. However, given the
paucity of research articles set specifically within culturally sustaining mathematics
teaching, we have extended our gaze to include relevant research which fit within the
many different terms used (e.g., culturally responsive instruction/teaching/relevant
teaching). We recognize the natural cross over to social justice mathematics teaching
and learning given the critical and sociopolitical aspects of culturally sustaining
mathematics teaching. We also acknowledge that culturally sustaining mathematics
teaching is complex and challenging and cannot be simply summarized as a set of
teaching practices and so to summarize, we have looked for research in mathematics
settings which takes an emancipatory stance on teaching and learning and which
consider “the social, emotional, cognitive, political and cultural dimensions of every
student” (Powell, Chambers Cantrell, Malo-Juvera, & Correll, 2016, p. 6) but we
have widened this lens to include not only individuals but also the collective.
The cornerstone of culturally sustaining mathematics teaching is effective high-
quality mathematics teaching and the systematic support given to all students to
access key disciplinary understandings. However, there are other significant
components which must also be considered. The building of relationships and trust
are positioned as of paramount importance in many research articles (e.g., Averill,
2012a; Bonner, 2014; Civil, 2014; Civil & Hunter, 2015; Hodge & Cobb, 2016;
Hunter & Hunter, 2018a; Powell et al., 2016). This includes relationships and trust
developed within the school setting, but also with parents and the wider community.
As an example, Bonner (2014) drew on the local community to deconstruct the
pedagogical practices of three successful mathematics teachers of diverse learners.
These teachers were nominated by members of the local community in meetings

250

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

which resulted from connections the researcher constructed as she participated


in neighborhood events (e.g., church, school events, after school activity). This
situated selection of the teachers within particular neighborhoods allowed these
communities to act as experts on which teachers they considered were responsive
to their children’s needs. In taking this action, Bonner positioned parents and
local community members as a resource, giving them the authority to decide what
constituted a successful mathematics teacher within their local community. Through
doing so, Bonner illustrated the importance of what many researchers (e.g., Civil,
2014; Jorgensen (Zevenbergen), 2016; Nicol, Archibald, & Baker, 2013; Nutti et al.,
2015; Wager & Whyte, 2013) have argued as needed –an asset-based approach and
the development of a shared power base between the home and the school.
Building strong and positive connections with the community, and within which
teachers assume a learner role, sits well within the asset-based approach we take in
culturally sustaining mathematics teaching. From the considerable work over three
decades within the Funds of Knowledge for Teaching project (González, Moll, &
Amanti, 2005) we see the need for teachers to develop deep understandings of their
students as cultural beings. Later in this chapter we provide more elaboration on
the Funds of Knowledge for Teaching project. Within this frame, students, parents,
and community are seen as potential collaborators and sources for instruction in
mathematics teaching and learning. To achieve this, Civil (2014) underlines the
importance of teachers taking time and developing “confianza (trust)” (p. 10). Civil
shares how a student in an interview gave rich detail of her weekend trips to a ranch
in Mexico which from a mathematics funds of knowledge approach provided many
opportunities for the construction of mathematical learning experiences based on her
out of school experiences. But, while Civil acknowledges the importance of teachers
making connections between out-of-school and in-school mathematics activity she
argues that the connections built between the teacher, students and their families
through showing deep and real interest in understanding their lived experiences is
more significant in constructing confianza.
The increased emphasis in recent times on mathematical discourse has placed
focus on who gets opportunities to talk and why. Relationships and trust are central
in ensuring that all students are able to participate in mathematical discourse. For
example, Civil (2014) describes how the development of confianza with a group
of middle school students allowed her to learn from their interactions. These often
included what appeared to be social chatting as well as such cultural elements as
humor and metaphors. She describes how this knowledge supported engagement
of a student who was low participator in mathematical talk when she became aware
that the student liked to argue. Similarly, Civil and Hunter (2015) tell the story
of how the concepts of trust and family supported a group of immigrant students
3ƗVLILNDLQ1HZ=HDODQGDQG0H[LFDQ$PHULFDQVWXGHQWVLQWKH8QLWHG6WDWHV WR
engage in disciplinary forms of mathematical argumentation. Although the styles of
argumentation were different, within the two contexts there were commonalities in
that both groups of students drew on values related to family (as a collective not as a

251

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

nuclear family) and used forms of humor and social talk as a way of face-saving. The
collaborative family-oriented stance both groups took resulted in the construction
of what looked like seamless mathematical arguments. However, when erroneous
concepts emerged the talk shifted to a deeper, richer level, as the students analyzed,
critically reflected on, and questioned their reasoning. In both these settings, the
teachers had created environments which supported the students to communicate in
comfortable ways and be themselves, to “bring in their cultural ways of being and
acting, including their home language(s), their ways of speaking, their use of humor,
and values not necessarily encountered in other areas of their schooling” and this
acted as a resource for engaging them in mathematical argumentation (p. 308).
Real and enduring relationships based on trust in mathematics classrooms are
dependent on teachers recognizing the importance of developing deep understandings
of their students, and their cultural connections to mathematics within their
everyday lived world. Specific patterns of communication are also important.
These include teachers communicating in culturally connected ways and ensuring
that the mathematics is accessible to students. They also need to communicate
deep respect and care to their students and about their students (Bonner, 2014). As
part of communicating care providing students with opportunities to construct a
positive mathematical disposition while retaining a strong racial and cultural identity
is identified as key outcome for culturally sustaining mathematics teaching (and
its other descriptors) in many studies (e.g., Averill, 2012a; Battey, 2013; Bonner,
2014; Civil & Hunter, 2015; Hunter & Hunter, 2018b; Hunter, Hunter, & Bills,
2018; Nicol, Archibald, & Baker, 2013; Shah, 2017). Bonner illustrated how four
highly successful teachers of traditionally underserved students created classroom
environments in which the students achieved mathematical success while also
constructing and maintaining robust racial and cultural identities. The teachers, in
order to build trusting relationships, all used rich knowledge of their students and
communication, but these played out in distinctly different ways. One teacher had
a highly organized and structured program in which she drew on culturally linked
methods in her instruction. To link her students explicitly to the mathematics she
included “chanting, storytelling, singing and movement” (p. 390). Another teacher
placed importance on constructing a non-threatening and safe environment in
which the students could access materials and activity as well as being able to
interact with other students in reasoned discussions in their home language. A third
teacher recognized the need for her students to have time to engage and struggle
with problems and so she organized them to work in pairs to reason and problem-
solve then engaged them in wider class discussions. Although communication was
foundational to each teacher’s work, power-sharing, pedagogy and discipline were
seamlessly intertwined. For these teachers, their communication, respect and care
included them holding high expectations and success for both students’ learning
and behavior. They were what Ware (2006) previously described as warm demander
practitioners. Fluidity of power sharing of the mathematics between the students and

252

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

teachers in the classrooms was also a central feature which they clearly linked to
student empowerment.
A key part of power sharing between teachers and students is their explicit
acknowledgement of students’ out of school mathematical activity and knowledge;
that is recognizing what González et al. (2005) described as their funds of knowledge
for teaching. The concept of funds of knowledge has been applied widely in recent
times (Civil & Andrade, 2002) and is further expanded on in one of the cases in
this chapter. Essentially, the basis of funds of knowledge is premised on teachers
becoming learners as they engage with and actively listen and learn from parents
and community members (Civil, 2007, 2014, 2016). Other studies (e.g., Foote,
2010; Jorgensen, 2016; Nicol et al., 2013; Owens, 2015; Pinxten & François,
2013; Sullivan, Jorgensen, Boaler, & Lerman, 2013) particularly those related to
indigenous communities support the need for teachers to recognize the importance of
the funds of knowledge of indigenous learners. These many studies outline the need
for teachers to be able to know and understand the culture and values of the students
and embed mathematical activity within these. For example, Hunter and Anthony
(2011) explored the pedagogical practices which best aligned with Pacific Nations’
cultural practices to support student engagement in inquiry and argumentation.
They illustrated how a teacher working with Pacific Nations students, drew on
their core values and beliefs around the concepts of reciprocity, collectivism, and
communalism to support classroom grouping arrangements and development of
mathematical inquiry and argumentation dialogue. We elaborate further on this work
later in the chapter. In another New Zealand study Averill (2012b) illustrated how
specific dimensions of teacher care need to be linked to the cultural perspectives of
LQGLJHQRXV0ƗRULDQG3DFLILFSHRSOHV$YHULOO  FRQWHQGVWKDWWHDFKHUV¶EHOLHIV
and practices need to be deeply embedded in the heritage cultures of indigenous
students to understand how to engage them appropriately in mathematics.
Recognizing and drawing on students out of school mathematics practices is
an important aspect of culturally sustaining mathematics teaching. D’Ambrosio
(1985) coined the term ethnomathematics to describe the study and presentation
of mathematical ideas of traditional peoples. The practices of teachers are clearly
evident around the concept of ethnomathematics however cognizance needs to
be taken around intent and actual practice. Pais (2011) cautioned the need to
ensure that best intentions of teachers do not have an opposite result. For example,
Wager (2012) reports the inherent tensions teachers experience when wanting to
understand students’ out-of-school cultural practices. She describes how a teacher
used a context without realizing that it was a newly developed practice and not a
traditional practice. Nutti (2013) also identifies tensions teachers may encounter
when engaging in Sámi culture-based mathematical teaching. Nutti describes
how a group of indigenous teachers designed and implemented a culturally-
based mathematical activity that included ethnomathematical content. These
teachers showed others how applying culture-based mathematics teaching can be
incorporated into an indigenous school setting. Although Nutti explains how the

253

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

teachers began to be self-empowered agents of change there were tensions which


included their concerns around their “own teacher role, teaching methods, and lack
of knowledge as an obstacle to the achievement of Sámi culture-based teaching”
(p. 70). Furthermore, the teachers voiced a dilemma they faced which resulted in
them choosing not to focus on culture-based teaching but rather using the national
teaching curriculum.
One aspect of the understandings of teachers towards the cultural heritage of
the students includes the role of language. Language is identified as an important
component for teachers to consider when working with both indigenous and other
diverse learners because of the key role it plays in students constructing rich
conceptual understandings. Studies (e.g., Anhalt & Rodríguez-Pérez, 2013; Civil
& Hunter, 2015; Clarkson, 2009; Jorgensen, 2016; Salehmohamed & Rowland,
2014; Phakeng, Planas, Bose, & Njurai, 2018; Xenofontos, 2016) suggest a range
of strategies including drawing on and encouraging students’ home language and
informal forms of talk to scaffold more formal mathematical discourse. In their
study, within the developing world, Phakeng et al. also suggest the need for teachers
to consider students who are multilingual. Jorgensen extends the view of language
to include patterns of interactions and illustrates the importance of this for teachers
working with Aboriginal learners. For example, she explains:
In Pitjantjatjara, speakers use directional language to explain positions – the
house is to the north of the store. So, in coming into the school context, not
only is a language introduced that is foreign, but the patterns of signification
are markedly different. (p. 325)
Jorgensen suggests the need for language differences to be approached within a
strength-based approach and for teachers to constructively support students to bridge
the two worlds “linguistically, culturally and conceptually” (p. 325).
In drawing together, the different themes which support teachers’ practices related
to culturally sustaining mathematics teaching there are many commonalities across
and within the different studies. The importance of relationships is pivotal to the
development of communication and power sharing. Relationships are underpinned
by high expectations of students and a belief that they can all succeed. The teachers
ensure students not only develop essential skills and strong understandings, they
are also able to draw on students’ languages and lived realities (Bills & Hunter,
2015). However, although there are commonalities across the different teachers’
practices in the studies what is interesting is how the different studies, do not
appear to take a comprehensive approach to include all aspects (e.g., Culturally
responsive/sustaining, indigeneity, ethnomathematics, funds of knowledge and
language).
Consistent with a sociocultural view of equity but shifting our lens towards a
more socio-political frame we turn our attention in the next section towards the
practices teachers draw on when teaching mathematics from a social justice position.

254

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

SOCIAL JUSTICE MATHEMATICS TEACHING

Working within the perspective of Social Justice Mathematics Teaching teachers


need to have knowledge and understanding of sociopolitical contexts that can be
studied through mathematics and that are student driven and of interest to students.
Rubel (2017) explains that it is the context that drives the mathematics rather than
the other way around. As a result, power relationships and also tensions between
emerging social justice goals and mathematical goals are prevalent themes in
the literature as teachers work towards enacting the practices of social justice
mathematics teaching (Atweh & Ala’i, 2012; Bartell, 2013). In a study with eight
secondary teachers learning to teach mathematics for social justice, Bartell argues that
“Teachers learning to teach for social justice need an understanding of themselves,
both personally and in relation to others” (p. 132). They also need an understanding
of different power structures (e.g., economic, social) and their role in education and
they need to have a deep understanding of their students. Bartell traced teachers’
changing ideas from focusing on reaching all students to holding more critical views
of how mathematics could be used to discuss inequities. This study underscores the
need for teachers to have opportunities to engage in discussions about what it means
to teach mathematics for social justice given that their understanding “guided and
constrained their development, negotiation, and instantiation of mathematical and
social justice goals in practice” (p. 159).
Providing opportunities to develop equity through using the pedagogical practices
inherent in social justice mathematics teaching is recognized in the literature
as challenging (Bartell, 2013; Felton, 2017). There are examples of teaching
mathematics for social justice (e.g., Gutstein, 2012; Skovsmose, 2012), however
our review of the literature showed that there are fewer examples focusing on how
practicing teachers learn to do this (e.g., Bartell, 2013; Felton-Koestler, 2017;
Leonard & Evans, 2012). One example is Planas and Civil’s (2009) description of a
professional development research project with a group of secondary mathematics
teachers working primarily with immigrant students. Key to teachers’ transformation
of their teaching was the implementation of a Teacher Study Group approach where
teachers first explored their notions of teaching mathematics to immigrant students.
Through these explorations, the idea of student participation became central and that
led the teachers to examine how to increase their immigrant students’ participation
in the mathematics class. Through dialogue in the Teacher Study Group, and
collaboration, teachers moved from a pessimistic and somewhat resistant attitude,
as if feeling that they could not do anything to change things, to an exploration of
what they were actually doing in class and what they could change (e.g., the kinds
of mathematics tasks they were using; the classroom dynamics). This refocused the
Teacher Study Group sessions towards developing tasks that would both promote
student communication about mathematics and engage and interest them. Planas and
Civil explain that through focusing on implementation of one such task – students
designing their ideal flat – teachers examined their own ideas about an ideal flat and

255

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

realized how their cultural and social backgrounds influenced their views. More
importantly, the teachers realized how these were different from the students’ views
of an ideal flat: “teachers realized that in special [e.g., with immigrant students]
classes they were likely to encounter other ways of interpreting the world” (p. 402).
An important aspect of this project was to have a space where teachers had the time
to discuss their ideas and plan activities.
In fact, evident in the literature is the significant time and space needed for
teachers to grapple with and engage with the tensions as they learn to enact the
practices of social justice mathematics teaching. Felton-Koestler (2017) presents the
case of an elementary teacher who went from direct instruction to student-centered
instruction, and eventually (within less than three years of her participation in
professional development program) to begin exploring social justice mathematics
teaching. Similar to the journey through equity described in the introduction
of this chapter, and also present in Bartell’s (2013) study, the teacher in Felton-
Koestler’s case perceived her role in enacting practices that supported equity as
somehow combining student-centered teaching practices and paying attention to
cultural backgrounds. But she saw equity in social justice mathematics teaching as
something different and suggested “that the social justice issue will overshadow the
mathematics and with the limited time we have in our classrooms the math will get
lost” (p. 11). It took until the third year of the study for her to begin considering social
justice mathematics teaching practices. Felton-Koestler argues that the reasons for
this shift lay in her reflectiveness and in her trust in Felton-Koestler (as the facilitator
of the long-term professional development program) and in the other teachers in the
project. Bartell (2013) argues for the need of sustained support for teachers who
want to implement practices aligned with social justice mathematics teaching as this
is a complex undertaking, but other factors also need consideration.
These different researchers illustrate that collaboration, dialogue, and openness
to trying different classroom arrangements and different types of tasks are very
important in teachers’ journeys towards developing the practices of social justice
mathematics teaching. However, tensions around mathematical content and coverage,
versus context, tasks, appropriate level and nature of the activity are recognized in
the literature (Bartell, 2013, Felton-Koestler, 2017; Planas & Civil, 2009). Atweh
and Ala’i (2012) describe some of the challenges that teachers faced as they tried to
teach from a socially response-able approach (social justice is part of this approach;
see Atweh & Brady, 2009 for a discussion of this approach). Some of these challenges
involved the tension between coverage of the required curriculum and the time
involved in engaging students in socially response-able projects as well as finding a
balance between the mathematics topics and the social justice issues. Felton-Koestler
described how, as part of her gradual journey, the teacher implemented a task on
income inequality in what she considered a safer environment of an after-school
setting, rather than in her regular classroom. However, she said that she would never
try an activity around immigration and border crossing which they had discussed in
the professional development sessions, as she found it too controversial. A study by

256

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Gonzalez (2009) adds important elements to consider as teachers start exploring the
practices related to teaching mathematics for social justice; two concerns emerged:
feeling overwhelmed by the demands that this may create in environments where
they do not have much flexibility or support and making students upset as they start
discussing social issues and they realize the inequities.
Attention is also drawn in the literature to the need to consider students’ reactions
to topics related to social justice (e.g., Nolan, 2009; Takker, 2017). Planas and Civil
(2009) refer to this issue as they discuss teachers’ concerns about whether the task
of designing an ideal flat is one that would be of interest to their immigrant students.
Takker notes that while a lot of attention has been devoted to mathematics teacher
knowledge, issues of social justice are missing from those accounts. This researcher
studied the practice of four elementary teachers in Mumbai as they implemented
real-life tasks from the 2005 national curriculum in India. One task presented a
scenario in which a husband and a wife, both farm workers, get paid differently
(the man more than the woman) and furthermore less than the minimum wage. The
mathematical question was a multiplication situation but also included questions
to prompt discussion on the differential pay and the minimum wage issue. In
implementation of this task neither teacher in the study discussed with the students
the questions and focused only on the multiplication. When asked about this, one
of the teachers seemed to dismiss the relevance of the story and focused on the
mathematical content. The other teacher said “this is outside knowledge. How much
of it to bring in class, I don’t know. I mean I don’t know how to handle this” (p.
941). As Takker writes, “what kind of knowledge does a teacher need to discuss and
handle social conflicts in a classroom? And what constitutes important mathematics
in such situations” (p. 942). These questions get to the heart of teachers’ practices in
social justice mathematics teaching.
In the context of South Africa, Vithal (2012) describes the challenges around
teaching mathematics through engaging students in real and relevant problems.
Vithal refers the need for a pedagogy of conflict and dialogue, and then adds the
need for also a pedagogy of forgiveness:
A mathematics that reveals inequities and injustices of the past or present is
likely to produce feelings of resentment and hate….A pedagogy of conflict and
dialogue for a mathematics education for equity and social justice invariably
opens wounds so that the ‘truth’ can be known, even relived, and understood.
Each learns by being in the place and experience of the ‘Other’. But if such a
pedagogy is not to run the risk of deepening divides and difference then it must
provide a means to heal. A pedagogy of forgiveness integrates into conflict and
dialogue, a point of hope and creative action. (p. 9)
A point in common to the work of Takker (2017) and Vithal (2012) is that while
the contexts are different (India and South Africa), they both refer to new curricula
that put emphasis on real-life problems. And in both cases, the need for teachers
who can make those mathematical connections is raised. As Vithal writes, this is

257

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

particularly an issue in schools in poor areas, which in the case of South Africa, are
schools that are largely attended by Black students. An important point from these
studies is the need for teachers’ mathematical knowledge to include knowledge of
social justice. This seems to be also what Panthi, Luitel, and Belbase (2018) are
suggesting for teachers in Nepal if they are to teach from a social justice approach.
As they write, “the mathematics curricula designed by experts and implemented by
the government to all grade levels do not commensurate with the minority and local
culture” (p. 23). Teachers need to make mathematical connections to their students’
everyday activities.
In summary, from looking at research studies with teachers learning to implement
social justice mathematics teaching, we see several points in common that affect
teaching practices:
‡ A tension between the mathematical goals and the social justice goals. Thus, it is
important to provide experiences where the mathematical goals are made explicit.
This may not only create more buy in among the teachers as learners, but they
may be more likely to then try to implement the activity in their own teaching. But
this tension is also an opportunity to discuss what should or could be the goals of
mathematics education in school (Wright, 2017).
‡ A need for a sustained, long term professional development program, preferably
in the form of a teacher study group that emphasizes dialogue and reflection.
Some studies point to the importance of teachers engaging in self-reflection on
their own life experiences as part of their envisioning what teaching for social
justice may look like (Carlson-Lisham & Esmonde, 2015; de Freitas, 2008).
‡ The importance of developing trusting relationships with the facilitators and with
the other teachers in the project.
‡ The opportunity to try out these ideas in safe environments such as an after-school
project, before attempting them in a regular classroom.

THREE EXAMPLES OF EQUITY FOCUSED WORK

In this section we highlight three equity-focused approaches that embody a


commitment to educational equity as we have discussed in this review and that
illustrate how these commitments are implemented in the mathematics classroom.
These long-term and recognized examples show the way in which the three
perspectives of teaching mathematics, Inclusive Mathematics Teaching, Culturally
Sustaining Mathematics Teaching, and Social Justice Mathematics Teaching, can be
woven together in the instructional practices teachers enact.

Funds of Knowledge for Mathematics Teaching

The original term funds of knowledge was coined by anthropologists Vélez-Ibañez


and Greenberg (1992) as they referred to the “strategic and cultural resources, which

258

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

we have termed funds of knowledge, that households contain” (p. 313). From that
work and in collaboration with educational researchers, the Funds of Knowledge for
Teaching project (González et al., 2005) was developed in the United States in the
1980s. Since then, the concept of funds of knowledge has spread to other research
projects across the world. In the 1990s, the Funds of Knowledge for Teaching project
was still active in its place of origin (Tucson, AZ) and led to a project that was
primarily focused on exploring the idea of funds of knowledge for the teaching and
learning of mathematics (Civil & Andrade, 2002). In this brief account, we draw on
over 25 years of this work to describe the main characteristics of teachers’ practices
when engaged in teaching mathematics from a funds of knowledge perspective.
At the heart of this work is the idea of teachers as learners from their students’
households. To this end, teachers conduct ethnographic home visits with a focus on
learning from the resources, experiences, and skills. Then through a Teacher Study
Group format, teachers-researchers and university-based researchers meet to work
on the transformation of these funds of knowledge into learning opportunities for
students in school. It is beyond the scope of this chapter to give a full description
of this large research project (e.g., see Civil, 2007, 2016; Civil & Andrade, 2002;
González, Andrade, Civil, & Moll, 2001). Here we only give the key findings from
the mathematics focused project as they relate to the theme of this chapter.
‡ While some teachers expressed initial apprehension to the idea of making home
visits to learn, it was one of the most powerful experiences towards building
relationships and developing or strengthening an assets-view of their students
and their families. As one teacher said, “realizing that the home is a real learning
place, … I didn’t think it was so much of a learning environment as it is” (Civil
& Andrade, 2002, p. 156).
‡ Learning about their students’ experiences and resourcefulness in out-of-school
settings allows teachers to develop a broader view of their students and redefine
what it means to be a competent learner (Civil, 2016).
‡ The pedagogical transformation of these funds of knowledge towards their use
in school mathematics can prove to be quite challenging (González et al., 2001).
This is in large part due to our views about what counts as mathematics (Civil,
2007). Explorations of different forms of mathematics supported by discussion of
readings in areas such as ethnomathematics were key elements in the Topic Study
Group sessions.
‡ Similar to the tension reported in the research on Social Justice Mathematics
Teaching between the social justice goals and the mathematical goals, there is
likely to be a tension between staying faithful to the funds of knowledge and
promoting a mathematical agenda. For example, in a module around construction,
the mathematics educator proposed a task that while mathematically rich was not
necessarily grounded on the funds of knowledge, which for the teacher was of
most importance (Civil, 2007; Sandoval-Taylor, 2005).

259

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

‡ The successful implementation of mathematical learning experiences grounded


on funds of knowledge (e.g., Ayers, Fonseca, Andrade, & Civil, 2001; Civil, 2002,
2007; Civil & Kahn, 2001; Kahn & Civil, 2001; Sandoval-Taylor, 2005) relied on
teachers’ determination to listen to their students and families and build on their
knowledge and experiences; in some cases, they involved the parents in the co-
development of the activities. It also relied on teachers’ interest and willingness
to expand their views and understanding of mathematics.
‡ By involving the parents in the activities either directly or through conversations,
teachers can gain an understanding of how students learn in the home environment,
which may be differently from the school setting. For example, at home they may
be learning through observation, interaction with other family members, and
scaffolded participation in the activity (Civil, 2016; Rogoff, 2012). In some of the
implementations, the teachers tried to bring in teaching approaches that emphasized
collaboration and participation (Civil, 2002, 2007; Kahn & Civil, 2001).
The funds of knowledge for teaching approach to equity is most clearly an example
of culturally sustaining mathematics teaching. Its pedagogical implementation also
reflects principles of inclusive mathematics teaching. The approach of doing home
visits to learn from and about the families and the community can serve as a step
towards social justice mathematics teaching.

Complex Instruction

Complex instruction is a theoretical and practical framework from which to


understand and enact a set of pedagogical strategies that explicitly attend to and
address inequitable participation in the classroom, especially when heterogeneous
groups of students work in cooperative groups. Founded in the research of sociologist
Elizabeth Cohen in the 1990s, complex instruction offers a sociological lens rather
than a psychological interpretation to peer interactions in the classroom (Cohen,
1994a, 1994b; Cohen & Lotan, 1997). When observing unequal participation of
students during cooperative groups, rather than interpreting those observations
as indexing a hierarchy of students’ ability, she saw something different. Cohen
explained those disparities by applying status-generalization theory, a theory
that explains how social hierarchies of who is expected to be competent and not
competent based on generalized desirable social characteristics that then come into
play when students interact with one another in school classrooms.
In the field of sociology, the theory of status generalization has been used to
research social inequities in various contexts such as the playground, jury room, work
groups (Webster & Foschi, 1988). The theory contends that status generalizations
are made in relation to characteristics associated with social advantage and cultural
preference. Gender and skin color, for example, are status characteristics that are
generally associated with social advantages and disadvantages in a myriad of
contexts. Status generalization theory explains patterns of inequitable interaction

260

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

by proposing that certain status characteristics create hierarchies of higher and


lower expectations for individuals. This, in turn, affects how people participate and
contribute, thereby reinforcing status hierarchies (Cohen, 1994; Webster & Foschi,
1988).
Students who are “good at” reading, or are highly verbal, assertive, or more of
whatever quality the dominant culture values tend to be those students who are
conferred higher status in a group and are given the most opportunities for learning.
The students who are the newcomers and have cultural markers that are rendered
less valuable based on stereotyped and generalized expectations are then provided
much less access and opportunities for learning. Cohen and colleagues did not simply
create a framework to understand and read classroom interactions through the lens
of status generalization theory, they also created a set of pedagogical strategies that
teachers could use to address the inequitable patterns this new framework would
allow them to begin to notice and see. Based at Stanford, Cohen and Lotan began
working with teachers and created a network of teachers who began to work with
their ideas. In mathematics education, the teachers from Railside, a school that
became the focus of Jo Boaler’s research documenting the success of teachers who
were teaching detracked and heterogenous mathematics classrooms using complex
instruction (e.g., Boaler, 2002, 2006; Boaler & Staples, 2008; Staples, 2014).
Complex instruction has become more widely known within mathematics
education as teachers from Railside have become instructional coaches, professional
developers, and university based mathematics educators who are sharing their
experiences and teaching others about the opportunities and challenges of this
approach to teaching mathematics for equity (e.g., Jilk, 2016; Nasir, Cabana,
Shreve, Woodbury, & Louie, 2014). This renewed interest in complex instruction
was propelled in 2011 when the largest organization of teachers of mathematics in
the United States and Canada, the National Council of Teachers of Mathematics,
published two books focused on complex instruction, one for elementary school
teachers – Smarter together: Collaboration and equity in the elementary
mathematics classroom (Featherstone et al., 2011), and the other for secondary
mathematics teachers – Strength in numbers: Collaborative learning in secondary
mathematics (Horn, 2012). These books are part of a growing collection of resources
that are now available for teachers committed to equitable participation in their
mathematics classrooms. These resources unpack core principles and practices of
complex instruction in their elementary and secondary mathematics classrooms by
working with practicing teachers and prospective teachers in a variety of contexts
and settings. This work entails: (a) disrupting what it means to be “smart” in the
mathematics classroom, (b) redefining “good” mathematics tasks as tasks that are
multidimensional, meaning require multiple abilities (sometimes called group-
worthy tasks), and (c) developing instructional strategies that address issues of
inequitable participation in group work, such as making teaching moves that “assign
competence” to low status students and by making teaching moves that create and
bolster interdependence among the students, such as setting up the expectation that

261

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

everyone within the group will have important intellectual contributions to make for
the success of the group and the benefit of everyone.
Disrupting what it means to be smart in mathematics is important because
students are often constructed as “high, medium and low” achievers in mathematics
and these rankings often correlate with how they are then tracked into the low,
medium, high-level mathematics classes. When definitions about what it means to
be smart in mathematics are based on a narrow set of skills (such as computational
skill) only a select few students will appear to be smart in the mathematics
classroom, and only a few students will be given the opportunity to study more
advanced mathematics content. Hence the importance of paying close attention to
how teachers and students think and talk about what it means to be smart in the
mathematics classroom.
Similarly, when considerations about what makes a mathematics task worthwhile
only focuses on the intellectual demand of the task and neglect to pay much attention
to social dimensions of the task, teachers continue to perpetuate normative and
widely held views about mathematics as a solitary activity that is best accomplished
by and in one’s mind. When teachers consider whether a task is rich, engaging and
cognitively demanding enough for a group to take on – is it group-worthy? – then
they begin to recognize how few school mathematics tasks are deliberately designed
to support students in working and getting smarter together in the mathematics
classroom.
In addition to these reconceptualizations, complex instruction also offers a
collection of pedagogical strategies to reinforce and support students’ access,
participation and contributions to their group task. In a nutshell, these strategies
focus on making public and deliberate teaching moves to establish classroom norms
and group roles that serve to support students’ individual and group participation and
sensemaking. Additionally, mathematics educators working within the framework of
complex instruction recognize that simply talking about these issues and becoming
aware of them is not enough. This work entails inviting teachers to work with
these ideas in the context of learning together with and from students and with and
from colleagues, which is unsurprisingly also a collaborative approach to teachers’
professional development. Teachers committed to the principles of complex
instruction work together not simply to improve their own mathematics teaching
and learning but also work with colleagues to design together complex instruction
math lessons and investigate together questions about students’ access, participation,
and learning in diverse classroom settings.
The complex instruction approach to equity is most clearly an example of
Inclusive Mathematics Teaching. However, its pedagogical implementation also
reflects principles of Culturally Sustaining Mathematics Teaching. Practices
of Social Justice Mathematics Teaching are also infused within this approach
through its attention to issues of authority, power, and social status in classroom
interactions.

262

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Developing Mathematical Inquiry Communities

Developing Mathematical Inquiry Communities began as a PhD study (see,


Hunter, 2007a, 2007b), the foundations of which were derived from a broad
base of international research (e.g., Kazemi, Franke, & Lampert, 2009; Wood,
Williams, & McNeal, 2006) around teaching and learning practices which provide
opportunities for all students to participate in mathematical practices and inquiry.
This provided a starting point, and in collaboration with a wider group of both
national and international educational researchers and teacher educators, Developing
Mathematical Inquiry Communities was established. Over the past decade and a
half Developing Mathematical Inquiry Communities, as a proven research based
professional learning and development program with a clear focus on equitable
outcomes for all students, has been very influential in how mathematics is taught
DQGOHDUQHGLQ1HZ=HDODQGDVZHOODVLQVPDOO3ƗFLILFQDWLRQVDQGLQVPDOOSRFNHWV
internationally.
In this short description, the findings of the past fifteen years are used to outline
and illustrate the key aspects of teachers’ practices within Developing Mathematical
Inquiry Communities. The role of teachers within various communities is central.
These communities not only include the classroom and school context but also
the local community. Within a multi-tiered collaborative approach, school leaders,
teachers, and teacher educators are located within a shared learning space where
focus is placed on co-constructing classroom environments in which all students
construct a positive mathematical disposition while maintaining a strong cultural
identity. In looking to scale-up, spread, is not limited to the implementation of
Developing Mathematical Inquiry Communities in new classrooms, but also concerns
the extent to which grounding principles and norms become embedded in school
and school system policies and routines. This requires that key school and system
leaders be encouraged and supported to develop a relatively deep understanding of
the underlying principles (Coburn, 2003).
Leadership is a critical mediator of fidelity and sustainability of reform.
Specifically, transformative pedagogical leadership is shown – one in which the
principal participates as a co-learner with teachers in moving learning and the school
forward – combined with policy stewardship creates conditions for continuous
organizational and system improvement (Robinson, Hohepa, & Lloyd, 2009).
Tailored to address the New Zealand context of systemic underachievement
of priority learners in mathematics classrooms, the professional learning model
for Developing Mathematical Inquiry Communities incorporates leading-edge
internationally recognized research frameworks (e.g., professional learning
(Timperley & Alton-Lee, 2008); practice-based teacher education (Lampert et al.,
2013); culturally sustaining pedagogies (Paris, 2012); and complex pedagogies
(Cohen, 1994a, 1994b). The pedagogy drawn on is ambitious (Kazemi, Franke, &
Lampert, 2009; Lampert, Beasley, Ghousseini, Kazemi, & Franke, 2010) and deeply
embedded within the core values and beliefs which underpin culturally sustaining

263

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

mathematics teaching through developing strong links to the community. This


requires significant professional learning. The school-based professional learning
and development is designed to increase the fidelity and quality of implementation
of Developing Mathematical Inquiry Communities.
The professional development and learning model establishes explicit goals
and expectations for student learning, building on participants’ shared images
and understanding about culturally responsive and equitable practices. It utilizes
instructional resources and associated smart tools (e.g., Communication and
participation framework – Hunter, 2008a, 2008b) to support mentors and teachers
to identify their own learning needs. At the same time, it focuses on building
professional norms and practices associated with adaptive expertise, deprivatizes
practice to build powerful professional learning communities and strengthens
leadership teams. This provides an infrastructure that supports generative learning
within a communities of learning framework. Studies indicate that resulting
trust, mutual accountability for students learning, and access to others’ expertise
and practice are at least as important as teachers’ perceptions of the value of the
professional development initiative.
Within Developing Mathematical Inquiry Communities commitment is made to
a student-centered mathematics approach where all students are enabled to succeed
by leveraging their strengths and abilities. The professionals are positioned as
learners so that teachers understand the strengths students bring to school through
parent-teacher meetings in which the students and their whanau (families and
wider community) describe the mathematics they do in their everyday lived lives.
Similarly, Developing Mathematical Inquiry Communities mentors work alongside
teachers and orientate their approach to strengths and abilities of mentees. The term
differentiated mentoring is used because all teachers enter and work within the
Developing Mathematical Inquiry Communities space at different and shifting points
on a continuum. Entry points and continued shifts align to many factors, including
how they view mathematics and who can do mathematics, their pedagogical
knowledge and skills and their own mathematical knowledge and relationship with
mathematics.
Listed below is a brief summary of the many findings which have emerged from
this professional learning and development project and their impact on teacher
practices.
‡ Constructing communities of mathematical inquiry and argumentation supports
student collaboration. Without explicit attention given to the mathematical
talk students were less able to engage in the discursive dialogue of inquiry and
argumentation (Anthony & Hunter, 2017; Hunter, 2006, 2008a, 2008b, 2012).
‡ Reconstruction of teacher practices is a lengthy journey. Teachers face many
tensions as they reconstruct their practices. Many of these emerge as a result
of teachers not having experience as learners in such classrooms (See, Anthony,

264

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Hunter, & Thompson, 2014; Hunter, Hunter, Bills, & Thompson, 2016; Hunter,
2005, 2010).
‡ The cultural capital, language and beliefs and values of diverse learners provide a
rich context for teachers to use to construct communities of mathematical inquiry.
Teachers and students experience dissonance as they learn to use and inquire
into a range of mathematical practices (Hunter & Anthony, 2011; Bills & Hunter,
2015; Civil & Hunter, 2015; Hunter, Hunter, & Bills, 2018; Hunter & Hunter,
2018; Hunter, 2007a, 2007b, 2013).
The Developing Mathematical Inquiry approach to equity is an example of both
Inclusive Mathematics Teaching and Culturally Sustaining Mathematics Teaching.
Like the other two examples it also has potential to connect with Social Justice
Mathematics Teaching practices through its attention to issues of school-wide
injustice such as detracking mathematics courses and rethinking ability grouping of
students.

CONCLUSION

This review of equity-based mathematics teaching practices is by no means


exhaustive. Our goal was to provide an overview of some key themes that we saw
in the literature we reviewed in the time frame 2008–2018. We chose a broad lens
of equity as opportunity to participate as our guide to select readings. We structured
the chapter along three main domains in equity in mathematics education: Inclusive
Mathematics Teaching; Culturally Sustaining Mathematics Teaching; and Social
Justice Mathematics Teaching. These three approaches place different demands on
teaching practices. For Inclusive Mathematics Teaching, the emphasis is on teachers
moving away from exclusionary forms of teaching practices and seeing that all
students bring mathematical strengths to the classroom. For Culturally Sustaining
Mathematics Teaching, the emphasis is on teachers seeing their students’ lived
experiences (home/community) as mathematically rich and developing instruction
that builds on these experiences. For Social Justice Mathematics Teaching, the
emphasis is on teachers engaging students in using mathematics to critically analyze
relevant issues, particularly responding to the sociopolitical turn in mathematics
education.
Common to these three approaches is the importance of teachers engaging in
teacher inquiry as they learn to design equity focused mathematical learning
experiences for their students. Some form of teacher inquiry group was present
in most research we reviewed. The need to develop trusting relationships among
teachers, students, university-based researchers, and community members was noted
in several studies. While perhaps more prominent in Social Justice Mathematics
Teaching, the three approaches point to a tension that equity-minded teachers often
experience between teaching to “required” mathematical goals and the need and

265

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

desire to focus on engaging their students in inclusive/culturally sustaining/social


justice-oriented mathematics learning environments.
While much progress has been made since 2008 in terms of studies that pay
attention to teachers committed to educational equity in mathematics teaching
and learning, much of the three approaches we discuss in this chapter have been
developed in parallel rather than in conversation with one another. Additionally, it
is important to acknowledge that in this review there is an overrepresentation of
research produced in the Western World and English dominant countries. Moving
forward we need more access to the research on equity in mathematics teaching in
other parts of the world. We also need to consider collaboration and cross pollination
of diverse approaches and that more voices are included in the next generation of
research focused on teachers committed to educational equity. To close, we want
to remind the reader of the opening quote by André Gide, “Everything has been
said before, but since nobody listens, we have to keep going back and beginning
all over again.” As we think of the pressing need that many teachers express in
wanting to develop more equitable learning environments and practices, we share a
sense of urgency and we hope that we do not have to be constantly “going back and
beginning all over again.”

REFERENCES
Anhalt, C. O., & Rodríguez-Pérez, M. E. (2013). K–8 teachers’ concerns about teaching Latino/a students.
Journal of Urban Mathematics Education, 6(2), 42–61.
Anthony, G., & Hunter, R. (2017). Developing student voice in the mathematics classroom. In B. Kaur
(Ed.), Empowering mathematics learners (pp. 99–115). London: World Scientific.
Anthony, G. J., Hunter, R., & Thompson, Z. (2014). Expansive learning: Lessons from one teacher’s
learning journey. ZDM-The International Journal on Mathematics Education, 46(2), 279–291.
https://doi.org/10.1007/s11858-013-0553-z
Atweh, B., & Ala’i, K. (2012). Socially response-able mathematics education: Lessons from three
teachers. In J. Dindyal, L. P. Cheng, & S. F. Ng (Eds.), Mathematics education: Expanding horizons.
Proceedings of the 35th Annual Conference of the Mathematics Education Research Group of
Australasia (pp. 98–105). Singapore: MERGA.
Atweh, B., & Brady, K. (2009). Socially response-able mathematics education: Implications of an ethical
approach. Eurasia Journal of Mathematics, Science and Technology Education, 5(3), 135–143.
Averill, R. (2012a). Reflecting heritage cultures in mathematics learning: The views of teachers and
students. Journal of Urban Mathematics Education, 5(2), 157–181.
Averill, R. (2012b). Caring teaching practices in multiethnic mathematics classrooms: Attending to health
and well-being. Mathematics Education Research Journal, 24(2), 105–128.
Averill, R. (2018). Using rich investigative mathematics activities towards embedding culturally
responsive mathematics teaching: Helpful, but sufficient? In R. Hunter, M. Civil, B. Herbel-
Eisenmann, N. Planas, & D. Wagner (Eds.), Mathematical discourse that breaks barriers and creates
space for marginalized learners (pp. 195–212). Rotterdam, The Netherlands: Sense Publishers.
Ayers, M., Fonseca, J. D., Andrade, R., & Civil, M. (2001). Creating learning communities: The “build
your own dream house” unit. In E. McIntyre, A. Rosebery & N. González (Eds.), Classroom diversity:
Connecting school to students’ lives (pp. 92–99). Portsmouth, NH: Heinemann.
Banks, J. A., & Banks, C. A. M. (Eds.). (2009). Multicultural education: Issues and perspectives (7th
ed.). Hoboken, NJ: Wiley.

266

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Bartell, T. G. (2013). Learning to teach mathematics for social justice: Negotiating social justice and
mathematical goals. Journal for Research in Mathematics Education, 44(1), 129–163.
Battey, D. (2013). “Good” mathematics teaching for students of color and those in poverty: The importance
of relational interactions within instruction. Educational Studies in Mathematics, 82(1), 125–144.
%LOOV 7  +XQWHU 5   7KH UROH RI FXOWXUDO FDSLWDO LQ FUHDWLQJ HTXLW\ IRU 3ƗVLILND OHDUQHUV LQ
mathematics. In M. Marshman, V. Geiger, & A. Bennison (Eds.), Mathematics education in the
margins. Proceedings of the 38th Annual Conference of the Mathematics Education Research Group
of Australasia (pp. 109–116). Sunshine Coast, Australia: MERGA.
Boaler, J. (2002). Experiencing school mathematics: Traditional and reform approaches to teaching and
their impact on student learning. Mahwah, NJ: Lawrence Erlbaum Associates.
Boaler, J. (2006). Opening their ideas: How a de-tracked math approach promoted respect, responsibility
and high achievement. Theory into Practice, 45(1), 40–46.
Boaler, J., & Staples, M. (2008). Creating mathematical futures through an equitable teaching approach:
The case of Railside school. Teachers College Record, 110(3), 608–645.
Bobis, J., Way, J., Anderson, J., & Martin, A. J. (2016) Challenging teacher beliefs about student
engagement in mathematics. Journal of Mathematics Teacher Education, 19(1), 33–55.
Bonner, E. P. (2014). Investigating practices of highly successful mathematics teachers of traditionally
underserved students. Educational Studies in Mathematics, 86(3), 377–399.
Brophy, J. E., & Good, T. L. (1986). Teacher behavior and student achievement. In M. C. Wittrock (Ed.),
Handbook of research on teaching (3rd ed., pp. 376–391). New York, NY: Macmillan.
Cai, J. (2017). Compendium for research in mathematics education. Reston, VA: National Council of
Teachers of Mathematics.
Carlson-Lishman, S. M., & Esmonde, I. (2015). Teaching mathematics for social justice: Linking life
history and social justice pedagogy. In S. Mukhopadhyay & B. Greer (Eds.), Proceedings of the
Eighth International Mathematics Education and Society Conference (Vol. 2, pp. 398–412). Portland,
OR: MES.
Celedón-Pattichis, S., White, D. Y., & Civil, M. (Eds.). (2017). Access and equity: Promoting high-quality
mathematics in grades preK-2. Reston, VA: National Council of Teachers of Mathematics.
Civil, M. (2002). Culture and mathematics: A community approach. Journal of Intercultural Studies,
23(2), 133–148.
Civil, M. (2007). Building on community knowledge: An avenue to equity in mathematics education. In
N. Nasir & P. Cobb (Eds.), Improving access to mathematics: Diversity and equity in the classroom
(pp. 105–117). New York, NY: Teachers College Press.
Civil, M. (2014). Why should mathematics educators learn from and about Latina/o students’ in-school
and out-of-school experiences? Journal of Urban Mathematics Education, 7(2), 9–20.
Civil, M. (2016). STEM learning research through a funds of knowledge lens. Cultural Studies of Science
Education, 11(1), 41–59. doi:10.1007/s11422-014-9648-2
Civil, M., & Andrade, R. (2002). Transitions between home and school mathematics: Rays of hope
amidst the passing clouds. In G. de Abreu, A. J. Bishop, & N. C. Presmeg (Eds.), Transitions between
contexts of mathematical practices (pp. 149–169). Boston, MA: Kluwer.
Civil, M., & Hunter, R. (2015). Participation of non-dominant students in argumentation in the
mathematics classroom. Intercultural Journal, 26(4), 296–312.
Civil, M., & Kahn, L. (2001). Mathematics instruction developed from a garden theme. Teaching Children
Mathematics, 7, 400–405.
Clarkson, P. C. (2009). Potential lessons for teaching in multilingual mathematics classrooms in Australia
and Southeast Asia. Journal of Science and Mathematics Education in Southeast Asia, 32(1), 1–17.
Coburn, C. E. (2003). Rethinking scale: Moving beyond numbers to deep and lasting change. Educational
Researcher, 32(6), 3–12.
Cohen, E. G. (1994a). Restructuring the classroom: Conditions for productive small groups. Review of
Educational Research, 64(1), 1–35.
Cohen, E. G. (1994b). Designing groupwork: Strategies for the heterogeneous classroom. New York, NY:
Teachers College Press.

267

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

Cohen, E. G., & Lotan, R. (1997). Working for equity in heterogeneous classrooms: Sociological theory
in practice. New York, NY: Teachers College Press.
Crespo, S., Celedón-Pattichis, S., & Civil, M. (Eds.). (2018). Access and equity: Promoting high-quality
mathematics in grades 3–5. Reston, VA: National Council of Teachers of Mathematics.
Crespo, S., & Featherstone, H. (2012). Counteracting the language of math ability: Preservice teachers
explore the role of status in elementary classrooms. In L. J. Jacobsen, J. Mistele, & B. Sriraman (Eds.),
Mathematics teacher education in the public interest (pp. 159–179). Charlotte, NC: Information Age
Publishing.
D’Ambrosio, U. (1985). Ethnomathematics and its place in the history and pedagogy of mathematics. For
the Learning of Mathematics, 5, 44–48.
de Freitas, E. (2008). Troubling teacher identity: Preparing mathematics teachers to teach for diversity.
Teaching Education, 19(1), 43–55. doi:10.1080/10476210701860024
Featherstone, H., Crespo, S., Jilk, L., Oslund, J., Parks, A., & Wood, M. (2011). Smarter together:
Collaboration and equity in the elementary math classroom. Reston, VA: National Council of
Teachers of Mathematics.
Felton-Koestler, M. (2017). “Children know more than I think they do”: The evolution of one teacher’s
views about equitable mathematics teaching. Journal of Mathematics Teacher Education. (Advance
online publication) doi:10.1007/s10857-017-9384-0
Fernandes, A., Crespo, S., & Civil, M. (Eds.). (2017). Access and equity: Promoting high-quality
mathematics in grades 6-8. Reston, VA: National Council of Teachers of Mathematics.
Foote, M. Q. (2010). The power of one: Teachers examine their mathematics practice by studying a single
child. In M. Q. Foote (Ed.), Mathematics teaching and learning in K–12: Equity and professional
development (pp. 41–58). New York, NY: Palgrave Macmillan.
Forgasz, H. J., & Leder, G. C. (2008). Beliefs about mathematics and mathematics teaching. In P.
Sullivan & T. Wood (Eds.), Knowledge and beliefs in mathematics teaching and teaching development
(pp. 173–192). Rotterdam, The Netherlands: Sense Publishers.
Frade, C., Acioly-Régnier, N., & Jun, L. (2013). Beyond deficit models of learning mathematics: Socio-
cultural directions for change and research. In M. A. (Ken) Clements, A. J. Bishop, C. Keitel, J.
Kilpatrick, & F. K. S. Leung (Eds.), Third international handbook of mathematics education
(pp. 101–144). New York, NY: Springer.
Gay, G. (2002). Preparing for culturally responsive teaching. Journal of Teacher Education, 53(2),
106–116.
Gay, G. (2010). Culturally responsive teaching: Theory, research and practice (2nd ed.). New York, NY:
Teachers College Press.
Gay, G., & Kirkland, K. (2003). Developing cultural critical consciousness and self-reflection in
preservice teacher education. Theory into Practice, 42(3), 181–187.
Gonzalez, L. (2009). Teaching mathematics for social justice: Reflections on a community of practice
for urban high school mathematics teachers. Journal of Urban Mathematics Education, 2(1), 22–51.
González, N., Andrade, R., Civil, M., & Moll, L. C. (2001). Bridging funds of distributed knowledge:
Creating zones of practices in mathematics. Journal of Education for Students Placed at Risk, 6,
115–132.
González, N., Moll, L. C., & Amanti, C. (Eds.). (2005). Funds of knowledge: Theorizing practice in
households, communities, and classrooms. Mahwah, NJ: Erlbaum.
Gutstein, E. (2012). Reflections on teaching and learning mathematics for social justice in urban schools.
In A. A. Wager & D. W. Stinson (Eds.), Teaching mathematics for social justice: Conversations with
educators (pp. 63–78). Reston, VA: National Council of Teachers of Mathematics.
Healy, L., & Powell, A. B. (2013). Understanding and overcoming “disadvantage” in learning
mathematics. In M. A. (Ken) Clements, A. J. Bishop, C. Keitel, J. Kilpatrick, & F. K. S. Leung (Eds.),
Third international handbook of mathematics education (pp.69–100). New York, NY: Springer.
Hodge, L., & Cobb, P. (2016). Two views of culture and their implications for mathematics teaching and
learning. Urban Education, 1–25. https://doi.org/10.1177/0042085916641173

268

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Horn, I. (2012). Strength in numbers: Collaborative learning in secondary mathematics. Reston, VA:
National Council of Teachers of Mathematics.
Hunter, R. (2005). Reforming communication in the classroom: One teacher’s journey of change. In
P. Clarkson, A. Downton, D. Gronn, M. Horne, A. McDonough, R. Pierce, & A. Roche (Eds.),
Building connections: Research, theory and practice. Proceedings of the 28th Annual Conference
of the Mathematics Education Research Group of Australasia (pp. 451–458). Sydney, Australia:
MERGA.
Hunter, R. (2006). Structuring the talk towards mathematical inquiry. In P. Grootenboer, R. Zevenbergen,
& M. Chinnappan (Eds.), Identities, cultures and learning spaces. Proceedings of the 29th Annual
Conference of the Mathematics Education Research Group of Australasia (Vol. 1, pp. 309–317).
Sydney, Australia. MERGA.
Hunter, R. (2007a). Scaffolding small group interactions. In J. Watson & K. Beswick (Eds.), Mathematics:
Essential research, essential practice. Proceedings of the 30th Annual Conference of the Mathematics
Education Research Group of Australasia (Vol. 2, pp. 430–439). Adelaide: MERGA.
Hunter, R. (2007b). Can you convince me: Learning to use mathematical argumentation. In J. Woo,
H. Lew, K. Park, & D. Seo (Eds.), Proceedings of the 31st annual conference of the International
Group for the psychology of mathematics education (Vol. 3, pp. 381–389). Seul, Korea: The Korea
Society of Educational Studies in Mathematics.
Hunter, R. (2008a). Facilitating communities of mathematical inquiry. In M. Goos, R. Brown, & K. Makar
(Eds.), Navigating currents and charting directions. Proceedings of the 31st Annual Conference of
the Mathematics Education Research Group of Australasia (Vol. 1, pp. 31–39). Brisbane, Australia:
MERGA.
Hunter, R. (2008b). Do they know what to ask and why? Teachers shifting student questioning from
explaining to justifying and generalising reasoning. In O. Figueras, J. Cortina, S. Alatorre, T. Rojano,
& A. Sepulveda (Eds.), Proceedings of the 32nd Annual Conference of the International group for the
psychology of mathematics education (Vol. 3, pp. 201–208). Morelia, Mexico: Cinvestav-UMSNH.
Hunter, R. (2010). Changing roles and identities in the construction of a community of mathematical
inquiry. Journal of Mathematics Teacher Education, 13(5), 397–409.
Hunter, R. (2012). Coming to know mathematics through being scaffolded to ‘talk and do’ mathematics.
International Journal of Mathematics Teaching and Learning. http://www.cimt.org.uk/journal/
hunter2.pdf
Hunter, R. (2013). Developing equitable opportunities for Pasifika students to engage in mathematical
practices. In A. M. Lindmeier & A. Heinze (Eds.), Proceedings of the 37th International Group for the
psychology of mathematics education (Vol. 3, pp. 397–406). Kiel, Germany: PME.
Hunter, R. K., & Anthony, G. (2011). Forging mathematical relationships in inquiry-based classrooms
with Pasifika students. Journal of Urban Mathematics Education, 4(1), 98–119.
Hunter, R., & Hunter, J. (2018a). Opening the space for all students to engage in mathematical talk within
collaborative inquiry and argumentation. In R. Hunter, M. Civil, B. Herbel-Eisenmann, N. Planas, &
D. Wagner (Eds.), Mathematical discourse that breaks barriers and creates space for marginalized
learners (pp. 1–22). Rotterdam, The Netherlands: Sense Publishers.
Hunter, R., & Hunter, J. (2018b). Maintaining a cultural identity while constructing a mathematical
GLVSRVLWLRQ DV D 3ƗVLILND OHDUQHU ,Q ( $ 0F.LQOH\  / 7XKLZDL 6PLWK (GV  Handbook of
Indigenous education (pp. 1–19). Singapore: Springer. https://doi.org/10.1007/978-981-10-1839-
8_14-1
Hunter, R., Hunter, J., & Bills, T. (2018). Enacting culturally responsive or socially-response-able
mathematics education. In C. Nicol, S. Dawson, J. Archibald, & F. Glanfield (Eds.), Living culturally
responsive mathematics curriculum and pedagogy: Making a difference with/in indigenous
communities (pp. 1–20). Rotterdam, The Netherlands: Springer.
Hunter, R., Hunter, J., Bills, T., & Thompson, Z. (2016). Learning by leading: Dynamic mentoring to
support culturally responsive mathematical inquiry communities. In J. Clark (Ed.), Opening up
mathematics research. Proceedings of the 39th Annual Conference of the Mathematics Education
Research Group of Australasia (pp. 109–116). Adelaide, Australia: MERGA.

269

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

Jablonka, E., Wagner, D., & Walshaw, M. (2013). Theories for studying social, political and cultural
dimensions of mathematics education. In M. A. (Ken) Clements, A. J. Bishop, C. Keitel, J. Kilpatrick,
& F. K. S. Leung (Eds.), Third international handbook of mathematics education (pp. 41–68).
New York, NY: Springer.
Jilk, L. M. (2016). Supporting teacher noticing of students’ mathematical strengths. Mathematics Teacher
Educator, 4(2), 188–199.
Johnson, L. (2014). Culturally responsive leadership for community empowerment. Multicultural
Education Review, 6(2), 145–170.
Jorgensen (Zevenbergen), R. (2016). Playing the game of school mathematics: Being explicit for
Indigenous learners and access to learning. Intercultural Journal, 27(4), 321–336.
Kahn, L., & Civil, M. (2001). Unearthing the mathematics of a classroom garden. In E. McIntyre,
A. Rosebery, & N. González (Eds.), Classroom diversity: Connecting school to students’ lives (pp. 37–
50). Portsmouth, NH: Heinemann.
Kazemi, E., Franke, M., & Lampert, M. (2009). Developing pedagogies in teacher education to support
novice teachers’ ability to enact ambitious instruction. In R. Hunter, B. Bicknell, & T. Burgess (Eds.),
Crossing divides: Proceedings of the 32nd Annual Conference of the Mathematics Education Research
Group of Australasia (Vol. 1, pp. 12–30). Palmerston North, NZ: MERGA.
Ladson-Billings, G. (1994). The dreamkeepers: Successful teachers of African American children. San
Francisco, CA: Jossey-Bass.
Ladson-Billings, G. (1995). Toward a theory of culturally relevant pedagogy. American Education
Research Journal, 32(3), 465–491.
Ladson-Billings, G. (2000). But that’s just good teaching! In J. Noel (Ed.), Notable selections in
multicultural education (pp. 206–216). Guildford, CT: Dushkin/McGraw-Hill.
Lampert, M., Beasley, H., Ghousseini, H., Kazemi, E., & Franke, M. (2010). Using designed instructional
activities to enable novices to manage ambitious mathematics teaching. In M. K. Stein & L. Kucan
(Eds.), Instructional explanations in the disciplines (pp. 129–141). New York, NY: Springer.
Lampert, M., Franke, M. L., Kazemi, E., Ghousseini, H., Turrou, A. C., Beasley, H., … Crowe, K. (2013).
Keeping it complex: Using rehearsals to support novice teacher learning of ambitious teaching.
Journal of Teacher Education, 64(3), 226–243.
Leder, G. C. (1992). Mathematics and gender: Changing perspectives. In D. A. Grouws (Ed.), Handbook
of research on mathematics teaching and learning: A project of the National Council of Teachers of
Mathematics (pp. 597–622). New York, NY: Macmillan Publishing.
Leonard, J., & Evans, B. R. (2012). Challenging beliefs and dispositions: Learning to teach mathematics
for social justice. In A. A. Wager & D. W. Stinson (Eds.), Teaching mathematics for social justice:
Conversations with educators (pp. 99–111). Reston, VA: National Council of Teachers of Mathematics.
Louie, N. L. (2017). The culture of exclusion in mathematics education and its persistence in equity-
oriented teaching. Journal for Research in Mathematics Education, 48(5), 488–519.
Maney, T., & Lange, T. (2013). Learners in transition between contexts. In M. A. (Ken) Clements,
A. J. Bishop, C. Keitel, J. Kilpatrick, & F. K. S. Leung (Eds.), Third international handbook of
mathematics education (pp.169–202). New York, NY: Springer.
Nasir, N. S., Cabana, C., Shreve, B., Woodbury, E., & Louie, N. (Eds.). (2014). Mathematics for equity:
A framework for successful practice. New York, NY: Teachers College Press.
Nicol, C., Archibald, J., & Baker, J. (2013). Designing a model of culturally responsive mathematics
education: Place, relationships and storywork. Mathematics Education Research Journal, 25(1),
73–89.
Nieto, S., & Bode, P. (2011). Affirming diversity: The sociopolitical context of multicultural education
(6th ed.). New York, NY: Pearson/Allyn & Bacon.
Nolan, K. (2009). Mathematics in and through social justice: Another misunderstood marriage? Journal
of Mathematics Teacher Education, 12, 205–216, doi:10.1007/s10857-009-9111-6
Nutti, Y. J. (2013). Indigenous teachers’ experiences of the implementation of culture-based mathematics
activities in Sámi school. Mathematics Education Research Journal, 25(1), 57–72.

270

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Nutti, Y. J., Fyhn, A., Eira, S. E. J., Sandvik, O., Borresen, T., Haetta, E., Somby, J., & Gaup, K. M.
(2015). Call your mothers! Sámi culture-based curriculum development based on mathematics
teachers, students and mothers in joint research actions. International Journal about Parents in
Education, 9(1), 10–23.
Oakes, J. (1990). Multiplying inequalities: The effects of race, social class, and tracking on opportunities
to learn mathematics and science (Report No. NSF-R-3928) [Research/Technical Report]. Retrieved
from Rand Corp. & National Science Foundation https://files.eric.ed.gov/fulltext/ED329615.pdf
Oakes, J. (1992). Can tracking research inform practice? Technical, normative, and political considerations.
Educational Researcher, 21(4), 12–21.
O’Connor, M. C. (1998). Language socialization in the mathematics classroom: Discourse practices and
mathematical thinking. In M. Lampert & M. Blunk (Eds.), Talking mathematics: Studies of teaching
and learning in school (pp. 17–55). New York, NY: Cambridge University Press.
Owens, K. (2015). Changing the teaching of mathematics for improved Indigenous education in a rural
Australian City Journal of Mathematics Teacher Education, 18, 53–78.
Panthi, R. K., Luitel, B. C., & Belbase, S. (2018). Teachers’ perceptions of social justice in mathematics
classrooms. REDIMAT, 7(1), 7–37. http://dx.doi.org/10.17583/redimat.2018.2707
Paris, J. (2011). Socioeconomic background and racial earnings inequality: A propensity score analysis.
Social Science Research, 40(1), 37–49.
Paris, D. (2012). Culturally sustaining pedagogy: A needed change in stance, terminology and practice.
Educational Researcher, 41(3), 93–97.
Paris, D. (2015). The right to culturally sustaining language education for the new American mainstream:
An introduction. International Multilingual Research Journal, 9(4), 221–226.
Paris, D., & Alim, H. S. (2014). What are we seeking to sustain through culturally sustaining pedagogy?
A loving critique forward. Harvard Educational Review, 84(1), 85–100.
Phakeng, M., Planas, N., Bose, A., & Njurai, E. (2018). Teaching and learning mathematics in trilingual
classrooms: Learning from three continents. In R. Hunter, M. Civil, B. Herbel-Eisenmann, N. Planas,
& D. Wagner (Eds.), Mathematical discourse that breaks barriers and creates space for marginalized
learners (pp. 277–293). Rotterdam, The Netherlands: Sense Publishers.
Planas, N., & Civil, M. (2009). Working with mathematics teachers and immigrant students: An
empowerment perspective. Journal of Mathematics Teacher Education, 12, 391–409. doi:10.1007/
s10857-009-9116-1
Planas, N., & Valero, P. (2016). Tracing the socio-cultural-political axis in understanding mathematics
education. In A. Gutiérrez, G. C. Leder, & P. Boero (Eds.), The second handbook of research on the
psychology of mathematics education (pp. 447–479). Boston, MA: Sense Publishers.
Pinxten, R., & François, K. (2011). Politics in an Indian canyon? Some thoughts on the implications of
ethnomathematics. Educational Studies in Mathematics, 78(2), 261–273.
Powell, R., Cantrell, S. C., Malo-juvera, V., & Correll, P. (2016). Operationalizing culturally responsive
instruction: Preliminary findings of CRIOP research. Teachers College Record, 118, 1–46.
Robinson, V., & Hohepa, M. L., & Lloyd, C. (2009). School leadership and student outcomes: Identifying
what works and why–Best Evidence Synthesis Iteration (BES). Wellington: Ministry of Education.
Rogoff, B. (2012). Learning without lessons: Opportunities to expand knowledge. Infancia y
Aprendizaje. Journal for the Study of Education and Development, 35(2), 233–252. http://dx.doi.
org/10.1174/021037012800217970
Ross, K. E. L. (2014). Professional development for practicing mathematics teachers: A critical connection
to English language learner students in mainstream USA classrooms. Journal of Mathematics Teacher
Education, 17(1), 85–100.
Rubel, L. (2017). Equity-directed instructional practices: Beyond the dominant perspective. Journal of
Urban Mathematics Education, 10(2), 66–105.
Sadker, D., Sadker, M., & Zittleman, K. R. (2009). Still failing at fairness: How gender bias cheats girls
and boys in school and what we can do about it. New York, NY: Simon and Schuster.
Salehmohamed, A., & Rowland, T. (2014). Whole-class interactions and code-switching in secondary
mathematics teaching in Mauritius. Mathematics Education Research Journal, 26(3), 555–577.

271

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MARTA CIVIL ET AL.

Sandoval-Taylor, P. (2005). Home is where the heart is: A funds of knowledge-based curriculum module.
In N. González, L. Moll, & C. Amanti (Eds.), Funds of knowledge: Theorizing practice in households,
communities, and classrooms (pp. 153–165). Mahwah, NJ: Lawrence Erlbaum.
Sfard, A., & Kieran, C. (2001). Cognition as communication: Rethinking learning-by-talking through
multi-faceted analysis of students’ mathematical interactions. Mind, Culture, and Activity, 8(1),
42–76.
Shah, N. (2017). Race, ideology, and academic ability: A relational analysis of racial narratives in
Mathematics. Teachers College Record, 119, 1–42.
Shah, N., Lewis, C. M., & Caires, R. (2014). Analyzing equity in collaborative learning situations:
A comparative case study in elementary computer science. In J. L. Polman, E. A. Kyza, D. K. O’Neill,
T. Iris, W. R. Penuel, A. S. Jurow, K. O’Connor, T. Lee, & L. D’Amico (Eds.), 11th International
Conference of the learning sciences (pp. 495–502). Boulder, CO: ISLS.
Skovsmose, O. (2012). Critical mathematics education: A dialogical journey. In A. A. Wager & D. W. Stinson
(Eds.), Teaching mathematics for social justice: Conversations with educators (pp. 35–47). Reston,
VA: National Council of Teachers of Mathematics.
Staples, M. (2008). Promoting student collaboration in a detracked, heterogeneous secondary mathematics
classroom. Journal of Mathematics Teacher Education, 11(5), 349–371.
Staples, M. E. (2014). Promoting student collaboration in a detracked, heterogeneous secondary
mathematics classrooms. In N. S. Nasir, C. Cabana, B. Shreve, E. Woodbury, & N. Louie (Eds.),
Mathematics for equity: A framework for successful practice (pp. 53–74). New York, NY: Teachers
College Press.
Stinson, D. W., & Bullock, E. C. (2012). Critical postmodern theory in mathematics education research:
A praxis of uncertainty. Educational Studies in Mathematics, 80(1–2), 41–55.
Sullivan, P., Jorgensen, R., Boaler, J., & Lerman, S. (2013). Transposing reform pedagogy into new
contexts: Complex instruction in remote Australia. Mathematics Education Research Journal, 25,
173–184.
Takker, S. (2017). Challenges in dealing with social justice concerns in mathematics classrooms. In
A. Chronaki (Ed.), Proceedings of the Ninth International Mathematics Education and Society
Conference (Vol. 2, pp. 936–945). Volos, Greece: MES9.
Timperley, H., & Alton-Lee, A. (2008). Reframing teacher professional learning: An alternative policy
approach to strengthening valued outcomes for diverse learners. Review of Research in Education,
32(1), 328–369.
Vélez-Ibañez, C. G., & Greenberg, J. B. (1992). Formation and transformation of funds of knowledge of
among U.S.-Mexican households. Anthropology and Education Quarterly, 23(4), 313–335.
Villegas, A. M., & Lucas, T. (2001). Educating culturally responsive teachers: A coherent approach.
Albany, NY: State University of New York Press.
Vithal, R. (2012). Mathematics education, democracy and development: Exploring connections.
Pythagoras, 33(2), 1–14. doi:10.4102/pythagoras.v33i2.200
Wager, A. A. (2012). Incorporating out-of-school mathematics: From cultural context to embedded
practice. Journal of Mathematics Teacher Education, 15(1), 9–23.
Wager, A. A., & Stinson, D. W. (Eds.). (2012). Teaching mathematics for social justice: Conversations
with educators. Reston, VA: National Council of Teachers of Mathematics.
Wager, A. A., & Whyte, K. (2013). Young children’s mathematics: Whose home practices are privileged?
Journal of Urban Mathematics Education, 6(1), 81–95.
Ware, F. (2006). Warm demander pedagogy. Urban Education, 41(4), 427–456.
Webster, M., & Foschi, M. (1988). Status generalization: New theory and research. Stanford, CA:
Stanford University.
White, D. Y., Fernandes, A., & Civil, M. (Eds.). (2018). Access and equity: Promoting high-quality
mathematics in grades 9–12. Reston, VA: National Council of Teachers of Mathematics.
Wood, T., Williams, G., & McNeal, B. (2006). Children’s mathematical thinking in different classroom
cultures. Journal for Research in Mathematics Education, 37(3), 222–255.

272

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHERS COMMITTED TO EQUITY

Wright, P. (2017). Teaching mathematics for social justice: Transforming classroom practice. In
A. Chronaki (Ed.), Proceedings of the Ninth International Mathematics Education and Society
Conference (Vol. 2, pp. 999–1010). Volos, Greece: MES9.
Xenofontos, C. (2016). Teaching mathematics in culturally and linguistically diverse classrooms: Greek-
Cypriot elementary teachers’ reported practices and professional needs. Journal of Urban Mathematics
Education, 9(1), 94–116.

Marta Civil
Department of Mathematics
University of Arizona

Roberta Hunter
Institute of Education
College of Humanities and Social Science
Massey University

Sandra Crespo
Department of Teacher Education
Michigan State University

273

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

10. INQUIRY-BASED PRACTICE IN UNIVERSITY


MATHEMATICS TEACHING DEVELOPMENT

This chapter addresses the use of inquiry-based practice in developing teaching


of mathematics at university level. It presents a three-layer model linked to theory
of Community of Inquiry and Critical Alignment. Community of Inquiry is seen to
be a transformation of Community of Practice. A brief review of relevant literature
both recognises its limited nature so far, and points to particular studies using
theory of inquiry. A three-layer model acknowledges development that is local
and related to developments in the practice of specific practitioners and extends
explicitly to developmental research in which systematic inquiry and publication
in research outlets communicates research findings to the wider community. These
theoretical ideas are exemplified through three case studies from projects which
have explored teaching development for university teachers and their students. A
theoretical construct, the Teaching Triad, is used for analysing teaching development
within an inquiry-based approach. A codicil to the chapter suggests a new way of
conceptualising scale in inquiry-based research.

INTRODUCTION

This chapter is focused in mathematics teaching and learning at university level


and their development though inquiry-based practice. Inquiry-based practice here
includes all aspects of the practices of teaching and learning in which research/
inquiry is a developmental component. It builds on research in the field of school
mathematics learning and teaching to extend what has been learned about the
development of mathematics teaching and learning through inquiry approaches to
activity at university level.
The extension to university mathematics brings with it the theories and
methodologies that underpin the research at school level, but also transcend the level
of education to be applicable to practices at the higher level. Of course, I appreciate
that many factors at university level are different from school level and these factors
have to be taken into account. These include the mathematical topics being taught,
the degrees of abstraction and formalism of university mathematics that differ from
school mathematics, and the systemic differences such as large grouping of students
and predominance of a lecturing mode of mathematical communication (Alsina,
2000; Nardi, 2008; Winsløw, Gueudet, Hochmut, & Nardi, 2018).

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_011

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

Nevertheless, in a critical review of teaching at university level (in general),


Kane, Sandretto, and Heath (2002) suggested that while having its own distinctive
characteristics, we can learn from research at the earlier educational levels (primary/
secondary), but, unlike these earlier levels, “many academics have had little or no
formal teacher education to prepare them for the teaching role.” Moreover, “we need
to understand how adults, and in particular university academics, learn to teach”
(pp. 181–182).
This suggests that an agenda for research at university level may have three
dimensions: (1) a dimension of theory and research carrying over from school level
education as appropriate; (2) a special consideration of the substantive, practical
and systemic details at the higher level; and (3) a developmental aspect through
which academics in university mathematics can experience and develop alternative
practices. These dimensions are addressed in what follows below.
In a very recent review of research in undergraduate mathematics education
seen through the Congress of the European Society for Research in Mathematics
Education (CERME) conferences, Winsløw et al. (2018) write that much less
research exists when it comes to university teachers’ pedagogical knowledge and
its development through formalised education. While the research community can
point to studies of teaching methods and practices, such as innovative uses of digital
technologies, organised, deliberate development of undergraduate mathematics
education teachers, based on research in undergraduate mathematics education, is
still rare.
In this article I build on theory of inquiry processes in mathematics teaching
and learning in three layers through which teaching and learning develop. Such
development of teaching and learning relates strongly to the development of those
involved in the practices – the students and the teachers. I will present a theoretical
perspective that has been found extremely pertinent in developing inquiry-based
learning and teaching at school level. This involves a three-layer model through
which inquiry in mathematics, in teaching and in research leads both to enhanced
knowledge in practice for both students and teachers and justified new knowledge
shareable in the wider community. I will follow this by a short theoretical section
introducing the idea of Community of Inquiry, relating this to Wenger’s theory of
Community of Practice and the important concept of Critical Alignment in which
I see teaching and learning development to be rooted. This will be followed by a
brief review of literature related to the development of mathematics learning and
teaching at university level – especially sources related to the use of inquiry-based
practice.
In the second part of the chapter I will discuss three cases of developmental inquiry
in practice at university level to demonstrate how teaching and learning can develop
to reveal the learning of all participants with associated issues in practice and to create
new knowledge of practical applicability. This will include a brief introduction to
the Teaching Triad, a tool for analysing and/or developing teaching practice. It will

276

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

include a perspective on developmental research as a legitimate form of knowledge


development for teachers in practice based on the three-layer model of inquiry. I will
end with a critical perspective on the growth of teaching and learning practice in
mathematics rooted in communities of inquiry and developmental research.

INQUIRY AS A DEVELOPMENTAL TOOL – A MODEL IN THREE LAYERS

Inquiry, as I have used it with colleagues, means to ask questions and seek answers,
tackle problems and seek solutions, explore, investigate, and overall look critically
at what we are doing and finding.1 We have drawn on Gordon Wells work (1999)
in a discussion of “dialogic inquiry” rooted in the work of Vygotsky. Wells draws
on notions of inquiry as “a willingness to wonder, to ask questions, and to seek to
understand by collaborating with others in the attempt to make answers to them,” and
as a means to emphasise “the essential continuity of education (Dewey, 1938, 1956)”
(Wells, 1999, p. 122). This continuity is seen through the use of inquiry by students in
classrooms, teachers responsible for their education, and those who are responsible
for teachers’ initial preparation and continuing professional development. Wells’
research focuses on teachers who are “attempting to develop such communities of
inquiry and simultaneously making their attempts the objects of their own inquiries”
(p. 124). These ideas can be seen to have relevance at university level as well as for
school-based studies.
In our use of inquiry with teachers in schools (e.g., Goodchild, Fuglestad, &
Jaworski, 2013) we have conceptualised a model involving three ‘layers’ of inquiry
(Figure 10.1): the central layer involves inquiry in mathematics as experienced by
teachers and students in the classroom; the second layer involves inquiry into the
processes of teaching mathematics as experienced by teachers who inquire into
the practices of themselves and their colleagues; the third layer is a ‘meta’ layer,
of research inquiry into the practices and processes in the two inner layers. The
research inquiry is conducted by both insider and outsider researchers (respectively,
those investigating their own practice and those investigating the practice of others)
and constitutes an inseparable amalgam of research and development which we refer
to as “Developmental Research” (addressed further below).
In the central layer we see participants engaging with inquiry-based processes to
address challenging mathematical questions appropriate to the level of education
of the students. Around this central domain, teachers (and possibly didacticians or
mathematics educators) inquire into the practices of teaching and learning and seek
to innovate and improve them for the benefit of students’ learning/understanding
of mathematics. The third or outer layer involves a research/inquiry process in
which observations and reflections of all participants lead to data which is analysed
rigorously, from well-defined theoretical perspectives, to provide insights into the
entire developmental process. (Jaworski, 2006; Jaworski & Potari, 2009; Goodchild
et al., 2013).

277

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

Figure 10.1. Inquiry in three layers

In a detailed review of Post-Calculus Research in Undergraduate Mathematics


Education (PC-RUME), Rasmussen and Wawro (2017) suggest that “inquiry
oriented instruction is founded on three cornerstones:
‡ deep student engagement in the mathematics;
‡ peer-to-peer collaboration;
‡ instructor inquiry into student thinking.
It is clear that Rasmussen and Wawro’s ‘take’ on inquiry-based practices in
university mathematics learning and teaching are very much in line with the
perspectives above deriving from school-based research. The third of these
‘cornerstones’ requires teacher learning from student activity (in line with the second
of the three layers above). “Peer to peer collaboration” reflects Wells’ emphasis on
the importance of discourse within inquiry-based practice.
Rasmussen and Wawro point to several examples of research showing inquiry at
the post-calculus level (e.g., in differential equations and abstract algebra) and hence
development of teaching practice. With reference to Marrongelle and Rasmussen
(2008) they claim that teachers working in inquiry ways need to “navigate a
continuum between all student discovery and all teacher telling” (p. 566). Their
reference to “deep engagement with mathematics” reminds us that learning and
teaching are of mathematics, and emphasis on mathematics is central to education

278

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

Figure 10.2. Inquiry in three layers at university level

in mathematics, whether from inquiry perspectives or other. Mention of “student


discovery” and “teacher telling” takes us deeply into the practices involved, and,
concomitantly, the philosophies of learning and teaching espoused by teachers at
the university level. These relate centrally to the meanings of “inquiry” espoused
by practitioners and their relationships with the culture and landscapes of practice
(Wenger-Trayner & Wenger-Trayner, 2018 (explored further below) at the particular
educational level. They relate also to pedagogical sophistication: the authors suggest
that such teaching practice demonstrates pedagogical insight which practitioners
develop through engagement in the research project. We can see this as activity
within the third layer of our model above. Figure 10.2 shows another version of this
model recast in language that refers to university mathematics directly.

DEVELOPMENTAL RESEARCH AND INQUIRY-BASED


TEACHING AND LEARNING

As said above, developmental research is an inseparable amalgam of research and


development. A developmental perspective, foundational to the three-layer model
presented above, is rooted in the idea of Communities of Inquiry between researchers
and teachers (and often students) growing together through their inquiry in practice.
Here practice can be seen as what we do, think and learn as we engage in the three

279

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

layers of inquiry. Communities of Inquiry have been seen (e.g., Biza, Jaworski, &
Hemmi (2014) to develop from the concept of Community of Practice initiated by
Jean Lave and Etienne Wenger (1991) and Wenger (1998).
Community of Practice, develops notions of community – “a way of talking about
the social configurations in which our enterprises are defined as worth pursuing and
our participation is recognizable as competence” and practice – “a way of talking
about the shared historical and social resources, frameworks, and perspectives that
can sustain mutual engagement in action” (Wenger, 1998, p. 5). Wenger offers three
dimensions through which “practice is the source of coherence of a community” –
these are mutual engagement, joint enterprise and shared repertoire (p. 72).
Engagement encompasses the myriad ways of participating in the practice; our
enterprises “include the instrumental, the personal, and the interpersonal aspects of
our lives,” while repertoire “can be very heterogeneous, gaining coherences from the
fact that they belong to the practice of a community pursuing an enterprise” (p. 82).
Considering what it means to ‘belong’ to a community of practice, Wenger offers
three further elements, engagement, imagination and alignment (p. 173).2 We engage
together in our joint enterprise, use imagination to trace individual trajectories
through our practice., and align with the norms and expectations of this practice.
For example, a teacher of university mathematics has to engage with all aspects
of teaching within the practices of the university: lecturing, tutoring, examining,
working with different sized groups of students, following designated curricula,
fitting into the expectations of academics, administrators, deans and so on, and
developing a personal repertoire that aligns with the ways of the community.
It may be seen that a community of inquiry, based in educational practice, may
be constituted through similar constructs to those designated by Wenger as being
fundamental to a community of practice: engaging with the enterprise, developing
a repertoire, using imagination to develop a personal trajectory and aligning with
norms and expectations. However, alignment is more problematic. Aligning with
the expectations and goals of the community is clearly necessary for community
engagement, but this does not have to be uncritical alignment. It is possible to
question the norms and expectations while aligning with them.
In fact, a community of inquiry can be constituted within a community of
practice. Inquiry is not the practice, it is a way of acting within the practice, a way
of conceptualising the practice, a way of being that judges the outcomes of practice
and seeks to modify, expand or develop them to achieve the goals of practice. So,
a community of inquiry expands from a community of practice and transforms
the practice in consequence. Here alignment becomes “critical alignment” and
this is exercised through inquiry-based processes. The effect of critical alignment
in university mathematics education, for example, is to open up possibilities
for development and change which are designed to tackle issues in students’
understandings of mathematics (Jaworski, 2006, 2014).
Critical alignment requires members of the community to engage critically with the
practices in which they participate. For example, the practice of “telling” (Rasmussen

280

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

and Wawro, above) gains a critical dimension when teachers seek to moderate their
telling through offering inquiry-based tasks to students. Concomitantly, students need
to develop their agency related to inquiry-based tasks, rather than being satisfied
merely to memorise what is offered by the teacher. Necessarily, both of these
actions raise issues which challenge engagement and imagination (community of
practice) resulting in the tensions and contradictions described in Biza et al. (2014) as
mentioned in the review below. Goodchild’s (2014) criticism, also mentioned below
is that community of practice specifically does not theorise elements of agency,
goal-directed actions, contradictions and tensions that are seen in Biza et al.’s (2014)
example of community of inquiry. My claim is that these elements pertain explicitly
to the transformation of community of practice to community of inquiry through the
construct of “critical alignment.” I elucidate these ideas further in the cases below.
Within the theoretical frame of community of inquiry, and centrally related to
the construct of critical alignment, I use a theoretical tool, the “Teaching Triad”
to characterise teaching. This was developed and extended through research in
secondary classrooms. Three domains were identified, Management of Learning,
Sensitivity to Students and Mathematical Challenge, together characterising the
teaching observed. Sensitivity and Challenge were found to be closely related in
so far as the sensitivities were appropriate to the mathematical challenges offered
to students. When challenge was offered with appropriate sensitivity, the outcome
was seen to be harmonious – fruitful for learning and development. Sensitivities
were classified as affective, cognitive or social, each enabling students to address
challenge. Through relationships between its three domains, it is possible to recognise
tensions in teaching/learning activity which lead to our exploring and questioning
the practices in which we engage (Jaworski, 1994; Potari & Jaworski, 2002). We see
below some uses of the triad in characterising teaching at university level.
The theoretical perspectives discussed above have been used in many studies
at secondary school level which draw on theoretical relationships, synergies and
tensions all relevant to my exposition here but hard to include in their entirety in a
suitably critical way within the space of this chapter.
Next, I offer a short review of research into teaching at university level, some
of which uses concepts of inquiry-based practice. Following this, I present three
examples from developmental research into mathematics teaching and learning at
university level through which I exemplify use of theory and extend it to research at
university level.

WHAT WE LEARN FROM THE LITERATURE

The paper from Rasmussen and Wawro (2017), discussed above, takes us into
the literature on university mathematics education. As indicated by many of the
researchers writing in this area, such literature is scarce, but growing.
We might start from the International Commission on Mathematical Instruction
(ICMI) study into The Teaching and Learning of Mathematics at University

281

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

Level, in which Alsina (2000) criticises a range of perceptions about established


mathematics teaching and associated teacher beliefs such as: “[teaching is] just a
matter of accumulated experience, clear presentation skills and a sound knowledge
of the subject,” a “context-free universal content,” a “top down approach,” a “formal
lecture paradigm” (pp. 4–5). He suggests instead, “a new paradigm of teaching
mathematics at university level” to include for example, reform/renewal of course
content to address context, historical backgrounds, modelling processes; innovative
technological tools, pedagogical strategies in mathematics education at the university
level (pp. 7–9). These aspirations fit with ideas of inquiry-based practice and relate
to research which has been conducted since 2000.
In a review into innovative approaches to teaching mathematics in higher
education, Abdulwahed, Jaworski, and Crawford (2012) draw attention to a number
of reports charting inquiry-based-learning practices, largely conceptualised from
constructivist foundations. These practices are usually focused on students’ learning
of mathematics suggesting that inquiry-based learning stimulates thinking, increases
students’ interest in the subject and their success rate, increases students’ appreciation
of the role of mathematics in life and their motivation to learn mathematics and
realises its applicability. Students tend to follow a conceptual approach in solving
problems and in searching for information. Little is said in these studies about
teaching development.
Concommitantly, a paper by Speer, Smith, and Horvath (2010) had claimed that
“very little research has focused directly on teaching practice [at university level]
– what teachers do and think daily, in class and out, as they perform their teaching
work” (p. 99). Focusing directly on teaching practice that is inquiry-based can be
seen to fit into this wider perspective of teaching practices and particularly the
development of teaching. It therefore seems important to seek out the ‘little’ that has
been done before 2010 and research since then that is relevant to an inquiry-based
practice approach.
An important consideration, for mathematics learning and teaching, concerns the
changing nature of the student population. More students are going to university
than ever before, from a wide range of backgrounds, who might not have had prior
mathematics teaching that enables them to tackle the more abstract and formal
modes of thinking required in university mathematics (Nardi, 1996; Hawkes &
Savage, 2000). Teachers have to acknowledge that large audience lectures in early
university courses and the amount of material which is presented in a typical
university course create problems for students. In addition, the so-called ‘service
teaching’, for students of science, engineering, economics, and so on, needs
alternative teaching practices related to the needs and interests of these students.
This raises questions about the nature of teaching that can be effective for this
wide range of experience of incoming students and we might look to inquiry-based
practice to address issues involved.
A literature review relating to undergraduate mathematics teaching (Treffert-
Thomas & Jaworski, 2015) made a distinction between three sorts of literature

282

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

in the area of undergraduate mathematics education: research, professional, and


pedagogical literature. The professional literature offers accounts from the personal
experiences of mathematicians or mathematics educators reporting from and
reflecting on their teaching practice. The pedagogic literature goes further to draw
out and distil aspects of pedagogy on which a teaching approach is based. While
not underestimating the value of such accounts in providing insights to practices
and their use of pedagogy, our main focus here is on the research literature which
accounts for what is observed, offering theoretical perspectives to explain and justify
outcomes. To date we find just a few papers which fit this category.
From a cognitive analytical perspective, Wagner, Speer, and Rossa (2007) and Speer
and Wagner (2009), in the United States, studied the teaching practices of professional
mathematicians teaching an undergraduate course in differential equations. For the
first time, the mathematicians used an inquiry-oriented curriculum which posed
considerable challenges in formulating and addressing particular instructional goals,
previously unfamiliar to the mathematicians. Of one mathematician, the authors
write: “His perceptions [i.e. reports of the challenges he faced in implementing
such a non-traditional instructional approach] are corroborated by observations
of his teaching and post-class interviews carried out by a mathematics education
researcher” (2007, p. 248). The authors claim that by studying the teaching practices
of professional mathematicians, it is possible to identify forms of knowledge (other
than mathematical knowledge) that are essential to inquiry-oriented teaching.
Focusing on theory related to teachers’ listening, Johnson and Larsen (2012)
report on their study of teaching of three teachers of abstract algebra in an inquiry-
based curriculum development project in a United States university. Researchers
videotaped all regular class sessions, took notes during each class session and held
regular videotaped debriefings with teachers. Although the three teachers were
generally successful at implementing the curriculum, analysis revealed a number of
instances in which students expressed confusion or difficulties in understanding the
mathematics, of which their teachers were unable to make sense. Further analysis
of data from one teacher led to a categorisation of interactions involving student
difficulties. While the teacher was particularly sensitive to dealing with students’
difficulties, it was seen that at times she was unable to enter into the students’
conceptions of the particular mathematics involved. We see here a recognition of
differing perspectives (or cultures) between teachers and students which challenge
both groups in the teaching-learning process in university mathematics. Such
tensions can be seen again in the first of the three cases below.
Another study focusing on the teaching of one professional mathematician was
conducted in Greece by Petropoulou, Potari, and Zachariades (2011) who investigated
the mathematics teaching of one of their team in the context of first year Calculus
lectures in a programme of study leading to a mathematics degree. The focus was on
the lecturer’s teaching decisions, actions and reflections and on the way that these
were linked to his different sources of experience. Different teaching practices that
the lecturer adopted to challenge the students to develop high level mathematical

283

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

reasoning, being at the same time sensitive to their needs, were identified. It was
found that he brought experiences from his practice as a research mathematician; his
involvement in mathematics education research; his participation in the group where
the focus was to inquire into teaching at the university level; and his actual teaching
practice. The researchers used self-reflection (the lecturer), critical questioning (the
other two researchers), and coordinated different interpretations of specific teaching
actions and decisions developing a deeper understanding of what characterizes
university mathematics teaching. Here we recognize the valuable partnership
between mathematician and mathematics education researchers working together
within the third layer of the inquiry model above.
From a sociocultural perspective, Jaworski, Robinson, Matthews and Croft
(2012) studied teaching in the implementation of an inquiry-based innovation into
the teaching of mathematics to first year engineering students in a United Kingdom
university. The teacher was one of a teaching-research team (of 4); data consisted of
observations of teaching practice, oral and written reflections of the teacher, surveys
of student perceptions, and focus group interviews with students. Findings showed
the development of teaching throughout the course as the teacher reflected on and
modified her teaching in discussion with the team. The study (discussed further in
Case 1 below) revealed tensions between the perspectives of the team in creating
and teaching the course and those of students taking the course, which were analysed
through an Activity Theory frame.
Initiatives are starting to become more visible in which university teachers of
mathematics explore ways in which they can develop their own teaching locally
and report on outcomes. In New Zealand, the DATUM project (Development and
Analysis of Teaching in Undergraduate Mathematics), including both mathematicians
and mathematics education researchers, began as a longitudinal project to develop a
model for professional development, theoretically grounded in Schoenfeld’s (2010)
resources, orientations, and goals model of teacher action. Each member of the group
had one of their lectures recorded and selected a short (3- to 4-minute) segment
for discussion, along with a brief written reflection of their sources, orientations,
and goals. Participants were encouraged to reflect on their teaching episodes, to
stimulate discussion of both mathematical and pedagogical knowledge and thereby
develop their practice organically. The study has had an enduring impact on teaching
practice (Barton, Paterson, Oates, & Thomas, 2014).
All of the above studies focused on the teaching of one or a very small number
of teachers, using qualitative approaches to study teaching in depth. We notice
that while four of these studies focus on teachers’ interpretation of inquiry-based
teaching (Jaworski et al., 2012; Johnson & Larsen, 2012; Rasmussen & Wawro,
2017; Wagner et al., 2007), other studies observe teaching in its ‘normal’ state, i.e.
teachers not trying to innovate or explore new approaches. Johnson and Larsen
focused on teachers listening to students and their responses to student difficulties.
Jaworski et al., in a sociocultural frame, noticed tensions between student perceptions
of teaching and those of the teaching team. Several papers pointed to the value

284

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

for practitioners of research into their practice promoting deeper reflections and
potential teaching development. All make clear the value of such in-depth studies of
teaching, for mathematics education research into higher education practices, with
several acknowledging their scarcity to date as pointed out by Speer et al. (2010).
I end this review with reference to a special issue of the journal Research in
Mathematics Education “Institutional, sociocultural and discursive approaches
to research in university mathematics education (2014)” which reports research
that was presented at the CERME conference in 2013. While all the papers add
substantially to the area of research in university mathematics education, two
papers should be mentioned particularly here. The first (Biza et al., 2014) considers
the theory of Community of Practice (e.g., Wenger, 1998) as a basis for research
into mathematics teaching and learning at university level with links to theory of
Community of Inquiry. The paper looks at a relationship between the two theories
based on the extension of community of practice to community of inquiry relying on
the inclusion of the construct of “critical alignment” discussed above. The second
of the two papers (Goodchild, 2014) is a critical review of the first which claims
that community of inquiry theory cannot derive from community of practice due to
incompatibilities in their foundations. As stated above, seeing community pf inquiry
as a transformation of community of practice through critical alignment, allows
questions of agency, tension and contradiction to be addressed.

SYSTEMATIC INQUIRY AS A BASIS FOR DEVELOPMENTAL


RESEARCH – THREE STUDIES

“Research is systematic inquiry made public” (Stenhouse, 1984). It can be said that
all research involves inquiry, but that not all inquiry involves research. The word
“systematic,” in Stenhouse’s succinct definition, offers a key. Engaging with inquiry
within a practice offers the possibility for (members of) the community to learn from
the inquiry processes involved. This is local or personal learning, which benefits
the teacher(s) inquiring and also their students. However, it does not (usually – for
example, in the professional and pedagogic papers mentioned earlier) go beyond
this. When these inquiry processes are organised systematically and the results
disseminated (made public) the result can be considered to be research. This means
that findings can be shared more widely and evidence given for findings from the
systematic inquiry.
Thus, developmental research within mathematics education can be seen as the
use of inquiry-based practices, systematically conducted and analytically processed
to judge the outcomes of inquiry relative to its initial goals. Developmental research
can be seen as the whole complex operation in the three layers of the inquiry model,
with the requirements for systematic inquiry being addressed within the third layer.
The learning of students and teachers within layers 1 and 2 does not depend on the
inquiry being systematic. However, this learning is localised, pertaining to the people
and situations involved, and shareable only anecdotally. The third layer, usually

285

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

involving experienced researchers (who might also be teachers in the practices


studied), ensures that analysis of data from layers 1 and 2 addresses systematically
the achievement of inquiry goals for layers 1 and 2.
The three cases which follow are examples of developmental research projects in
which inquiry-based practice leads to learning and development.3 We see examples
of learning at a number of levels within the three layers of the inquiry model. In all
cases we see critical alignment influencing learning in differing ways in the projects.
In all three studies we see teacher learning (and associated developments in practice)
to be an important outcome alongside the mathematical learning of their students.
Methodology within the three projects is similar, while attending to the specific
research questions in each case. Data, collected with respect to the research
questions was largely qualitative. For example, in Case 1, data was collected through
observation of teaching events, notes from teaching team meetings, post-teaching
interviews with students, and assessment marks from group projects and exam. In
Case 2, qualitative data was collected from interviews with lecturers and interns,
meetings of interns and staff were recorded, as were peer tutorials. Quantitative
data was obtained from attendance at peer tutorials and finally-achieved grades by
students. In Case 3, design meetings were recorded, student partners were interviewed
at several stages of the research, tutorials with foundation studies students and screen
data were recorded. The tasks designed were themselves data. Grounded analyses
of qualitative data were undertaken in all three cases to address research questions.

Case 1 – Engineering Students Understanding Mathematics

The Engineering Students Understanding Mathematics (ESUM) project, supported


financially by the United Kingdom HE-STEM programme, was designed to
improve teaching of mathematics in a mathematics module for first year materials-
engineering students in a United Kingdom university (see also Biza et al., 2014;
Jaworski, 2014). Students (n=50) came to university having learned mathematics
(often very procedurally) in their schools and traditional lecturing had seemed to
encourage a procedural approach. It was an aim of the project to promote more
in-depth learning of mathematical concepts though an inquiry-based approach.
Thus, three experienced teacher-researchers designed an innovation to teaching,
employing inquiry-based questions, small group problem solving, a computer-based
learning environment and an assessed group project (Jaworski & Matthews, 2011).
The community of practice here involved the research team (3 teacher-researchers
plus a research officer) and the cohort of engineering students, all engaged in a
practice which we will call teaching-learning mathematics. We could alternatively
see this as two communities (teaching team and students) and two modes of
practice (teaching and learning): it then becomes important to consider how these
communities interact, boundaries between them and ways in which learning crosses
these boundaries. In more recent work than that described above, Wenger-Trayner
and Wenger Trayner (2018) write about “landscapes of practice” and the concept of

286

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

“knowledgeability” in a landscape of practice (2018). I will return to these concepts


after the three cases.
In this first case, one member of the teaching team was the lecturer: the team
designed the teaching activity and the lecturer interpreted the design in practice. For
example, a goal to use more questioning in lectures, in order to involve students in
thinking mathematically, was interpreted in terms of specific questions to (named)
students during the course of the lecture. Two important aspects of learning (for
the teaching team) were (1) that many students responded to the questions and this
showed evidence of student engagement with their responses providing feedback
to the lecturer who could then ask a further question if necessary; (2) the questions
themselves were rarely of an inquiry nature: rather they involved opportunity for
students to interject responses relevant to the mathematics under consideration.
While it had been an aim to include inquiry-based questions, the value of short, more
factually based questions in getting students involved was seen to be an important
insight. This insight emerged over time as the teaching progressed. It was therefore
a local learning experience relating to layers 1 and 2 of the model. However, a
grounded systematic analysis of lecture observations confirmed it as an important
learning from the project as a whole.
The use of inquiry-based questions was almost entirely reserved for tutorial work
where students were asked to participate in groups, addressing questions related to
a computer-based environment. The lecturer and graduate assistant circulated the
groups engaging students in discussion on the inquiry-based questions and promoting
an inquiry attitude to the mathematics in focus. For example, on the topic of inverses
of functions, students were given certain functions, asked to determine whether or
not each one had an inverse (with justification) and if not, to explore intervals on
the domain where an inverse could exist. They were asked to use GeoGebra to draw
graphs of the functions and use the graphs to argue the existence (or otherwise) of the
inverses. Where students were trying to engage in these suggested ways, a conversation
between the group and the lecturer was seen to explore student perceptions, promote
specific approaches, and encourage students’ own rationalisations. However, some
students resisted the demands of thoughtful mathematical engagement: some played
with GeoGebra, filling the screen with a multitude of functions; others spent time
chatting on other topics or accessing social media. Addressing such problems of
practice was of local concern from one week to the next: it involved lecturer and
teaching team in inquiry into ways in which students might see value in participation
in inquiry-based tasks (thus, related to layers 1 and 2 of the model).
Observation showed evidence of all these modes of action. Ongoing analysis
within the team led to considerations of alternative possibilities for teaching action,
some of which could be incorporated as the weeks progressed. Others were noted for
future cohorts. While it was clear that both teacher-researchers and students learned
from the activity, the project was significant for the teachers, revealing issues relating
to innovation and insights into students’ perceptions of learning and teaching, both
of which influenced future practice (Jaworski et al., 2012; Jaworski, 2015).

287

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

A Teaching Triad skeleton analysis of some aspects of teaching in the Engineering


Students Understanding Mathematics project can be seen in Table 10.3. The Teaching
Triad aims to show the coherence of teaching with Mathematical Challenge and
Sensitivity to Students operating within Management of Learning to draw students
into mathematical thinking with Sensitivity to Students providing a supportive
structure to encourage mathematical engagement (with mathematical challenge). In
the Engineering Students Understanding Mathematics project, the teaching-learning
practice in the joint community is characterized by interactions captured by the
Teaching Triad. For example, use of names in questioning in lectures encouraged
some students to respond frequently to questions. The lecturer had to ensure that
her feedback to students was appropriately sensitive, especially when student
responses were not correct. Although initiative in use of questioning in lectures came
from the teaching team (design) and lecturer (interpretation), the Teaching Triad
characterisation and data from observation suggest a joint community interacting in
practice to support student engagement with (and learning of) mathematics resulting
in corresponding teacher learning of teaching design and pedagogic process.
The inquiry community here can be seen in the overall inquiry activity of the
teaching team. The team designed an inquiry-based interpretation of the curriculum
topics. The Teaching Triad analysis draws attention to teaching goals and the extent
of their achievement. Particularly, inquiry-based tasks in tutorials provided for the
inner layer of the model, and observational data showed the extent to which this
design achieved its goals. Thus, teacher-researchers inquired into teaching and
learned from the associated research.
Critical alignment can be seen in the interpretation and re-interpretation of the
design of inquiry-based practice. When the design appeared to achieve its goals,
further confirmation was sought from analysis of accumulated data. When it was
clear that there were tensions and issues, some level of redesign was necessary
to enable goals to be reached. We see here the meaning of critical alignment for
teaching within this project.
For students who engaged with inquiry-based tasks, observation showed that
their mathematical concepts developed. That not all students engaged in the ways
designed, was a learning experience for the design team; lack of success for some
students suggested that alternative activity and redesign was necessary. Critical
alignment for students can be seen in their need to adjust their perspectives on
learning mathematics from a procedural practice, provided by the lecturer, to one
of mathematical inquiry. Only some of the students achieved this and it remained a
challenge for the teaching team to address this issue.
As we report in Jaworski et al. (2012), students’ perspectives on inquiry-
based practice, as stated in post-teaching interviews, contrasted significantly with
the perspectives of the teaching team. University procedures and departmental
philosophy had required assessment through an end of module exam: despite
being aware of the reasoning behind inquiry-based questions and use of GeoGebra,

288

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

Table 10.1. Teaching Triad analysis of teaching-learning in the ESUM project

Triad Elements of Teaching Practice Relation to Students

Management of ‡ Designed and managed the ‡ More students took part than
Learning innovation. Encouraged more previously. Some discussion
student engagement in lectures between students in the lecture
‡ Managed a computer laboratory ‡ Some groups worked conceptually,
with small group activity to others not
encourage conceptual work ‡ Groups presented good projects,
‡ Changed assessment to but not clear that all students were
include group project work, involved
although retaining a traditional ‡ Traditional examination
examination encouraged students’ procedural
view of learning
Sensitivity to ‡ Used students’ names in ‡ Hard to recall all names with 50+
Students addressing questions in lectures cohort
(Affective) ‡ Effective visual perception
‡ Encouraged visual perception required concentrated engagement
of mathematics using GeoGebra – not all students could sustain it
(Cognitive) ‡ Some groups were able to work
‡ Encouraged communicative effectively – others not
learning through group activity ‡ Teaching needed to foster
(Social) sustained engagement – how to
‡ Encouraged students’ enable students to learn to work
mathematical engagement and effectively in a group and sustain
investigation through questions concentration
in lectures and tutorials
(Cognitive)
Mathematical ‡ Asked more demanding (inquiry- ‡ Some students engaged, others not
Challenge based) questions in tutorials ‡ Some groups achieved greater
‡ Sought a deeper level of engagement and mathematical
mathematical engagement depth, others not
through investigative tasks ‡ Exam proved to be a barrier to
‡ Drew attention to more conceptual engagement.
conceptual understanding and ‡ How to achieve engagement with
challenged procedural learning more students?

students saw the final exam as reason for preferring a more procedural approach to
learning the mathematics of the module.
The Teaching Triad analysis draws attention to tensions and contradictions faced
by the teaching team in designing the innovation and interpreting it in practice.
Teacher learning can be seen in dealing with day to day issues arising in practice
(layers 1 and 2) as well as in the outcomes from analysis of data and theoretical
rationalization (layer 3) (Jaworski & Matthews, 2011; Jaworski et al., 2012).

289

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

Case 2 – Students’ Design of Resources and Peer Support in Module Development

In one United Kingdom university, two second-year mathematics modules (Complex


Analysis and Vector Spaces) had a record over several years of students’ low
performance. The SYMBoL project (Second Year Mathematics Beyond Lectures),
supported by HE-STEM funding in the United Kingdom, set out to improve the
student experience by designing new resources to support students. Four former
students of the two modules (interns) were employed for 6 weeks in the summer to
work with teachers to design resources which would help future students to improve
conceptual understanding.
Through the resulting teacher-student community, interns and teachers together
learned about the needs of teaching and the design of resources, and teachers gained
insights to student perspectives (Duah, Croft, & Inglis, 2014; Duah, 2017). The
following points were notable:
‡ Neither interns nor the module lecturers were initially familiar with talking with
each other about mathematics, so there were concerns on both sides about the
possibilities of interaction.
‡ One staff member, a lecturer in other modules, offered tea in her office each
afternoon at 4pm, to which the interns and all interested mathematics staff were
invited. Ensuing, wide-ranging discussion about mathematics, learning and
teaching broke the ice and enabled interns and staff to talk as colleagues about
module development.
‡ Interns acknowledged that, in reviewing the module in order to design resource
material, they learned mathematics in deeper and more effective ways than had
been possible during the modules.
‡ Interns’ learning experiences and the clear value of interns’ contributions to
the material of the modules led to a vision of including students in the ongoing
module teaching.
As well as the design of new resources, an important consequence of the four
points above was an addition to the modules, in the next year of teaching, of peer
tutoring for second-year students. Pairs of third-year students, the ‘peer-leaders’
taught their peers in specially organized tutorials. These tutorials were an addition
to the regular module tutorials led by mathematics staff and graduate assistants and
were optional for the undergraduates. The peer-leaders were ‘trained’ by a small
group of mathematics staff (including mathematics educators) to work with their
peers in supportive ways encouraging dialogue and questioning. The peer tutorials
were attended by more than half of the student cohort; research showed that those
who attended, as a whole, achieved more highly on the formal assessments of the
modules than their non-attending peers (Duah et al., 2014).
Unlike in the Engineering Students Understanding Mathematics project where
a Teaching Triad analysis focused on the teaching implemented through the
innovation, here the Teaching Triad analysis is of the project itself and the ways

290

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

Table 10.2. Teaching Triad analysis of activity in the SYMBoL Project

Triad Elements of Teaching Practice Relations with Students

Management ‡ Interns designing tasks in an area ‡ Interns learned mathematics more


of Learning – of mathematics that they have not conceptually than earlier
Project understood well before ‡ Activity reinforced peer leaders’
Management ‡ Peer leaders handling the conceptual understanding
mathematics needed in being a ‡ Experience of talking
tutor for their peers mathematics with interns
‡ (Lecturers) talking mathematics revealed, for lecturers, the
with student interns who do importance of findings ways to
not have their mathematical talk mathematics with students
sophistication ‡ New activity and practices were
‡ Organising peer tutorials for a new thought valuable enough to be
cohort included in future practice
Sensitivity to ‡ Interns and staff having tea ‡ The importance of social
Students together each day (affective and considerations in an academic
social) innovation
‡ Talking about mathematics at an ‡ Both lecturers and students found
‘equal’ level (staff and interns) value in talking mathematics
(affective and cognitive) together
‡ Students working in tutorials ‡ Both students and peer leaders
with more experienced peers found value in learning in peer-
(new cohort with peer leaders) led tutorials
(affective, cognitive and social)
Mathematical ‡ Interns designing new resources in ‡ The new approach to engagement
Challenge an area of mathematics that they with mathematics led to enhanced
have not understood well before understanding for interns
‡ Peer leaders handling the ‡ Being a tutor demanded much
mathematics needed in being a more of students in terms of
tutor for their peers engaging with mathematics than
‡ (Lecturers) talking mathematics they experienced as a student in
with student interns who do the past
not have their mathematical ‡ Lecturers developed awareness
sophistication of student understanding through
‡ Lecturers needed to find language use and alternative
alternative ways of expressing modes of expression
mathematics, when formality did
not seem to work well

it supported teaching and learning of all concerned and ultimately the second-year
students of mathematics whose mathematical understanding was the object of the
project as a whole.
The inquiry community here was situated, initially, within an overtly
developmental practice involving the design of resources for future learning and

291

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

teaching in the two modules. The community consisted of the two lecturers, each
responsible for one module, the four interns, and other lecturers (mathematics staff
not graduate students) showing an interest in the project.
Also included in the project community was a PhD student, employed to collect
and analyze data. His activity operated at layer 3 in the inquiry model, providing a
systematic layer of inquiry. Data were collected from observations of practice and
interviews with lecturers, interns and peer leaders.
Critical alignment was visible for both lecturers and interns. There were no
precedents for their joint practice. Through talking together they agreed a basis
for the resources, interns experimented with a variety of sources and drew on
their own awareness of areas of mathematical difficulty, several drafts of material
were discussed and redrafted. In interview, interns and lecturers independently
acknowledged their initial concerns for the collaboration: lecturers as to whether
they could talk mathematics at a level understandable by the interns, and interns
concerned with their mathematical capability and whether they would be able to
communicate adequately. The meetings over tea provided an unexpected forum with
lecturers and interns all agreeing their enjoyment in participation in the mathematical
discussion, and this providing a catalyst for the resource development. We can see
here a (possibly implicit) critical alignment in which lecturers and interns, despite
their initial fears, participated in dialogue, supported by the informal atmosphere
over tea.
The subsequent peer teaching in tutorials emerged from the practices above and
formed a new layer of practice. This time the community included peer leaders,
second-year students, module lecturers, the PhD researcher and other lecturers who
worked with peer leaders to develop teaching approaches. Critical alignment here
was most evident for the peer leaders who had to prepare for tutorials and find ways
to work with their student peers. Students who chose to attend their peers’ tutorials
had to critically align in finding new ways of interacting with each other and with
their peer leaders.
The practices that developed through the SYMBoL activity showed important
learning experiences for both staff and interns. Staff became better able to talk
mathematics with students and interns responded well to talking mathematics with
staff. Evidence of learning at multiple levels is communicated through systematic
inquiry across these levels presented in the resulting PhD thesis (Duah, 2017).

Case 3 – Student Partners in the Design of Computer-Based Tasks

Funding from the United Kingdom HEFCE (Higher Education Funding Council)
Catalyst Programme (http://www.hefce.ac.uk/funding/catalyst/), enabled a project
building on findings from the two projects above. The project involved the
employment of Student Partners to design tasks for students in the university’s
Foundation Studies programme (preparing students for future undergraduate
studies in science and engineering). The student partners employed were former

292

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

foundation studies students who had achieved high scores in their foundation studies
mathematics module and were now successfully engaged in first year studies (in
Physics, Chemistry, Mechanical Engineering and Chemical Engineering). The
project community here included three teacher-researchers (staff in the Mathematics
Education Centre) four student partners (selected through application and interview),
two analytical assistants (PhD students employed to gather and help analyse data
from the project) and one expert in the computer software (Autograph) used for task
design. One of the teacher-researchers was the lecturer in a mathematics module for
foundation studies students.
The object of this project was threefold
‡ To design inquiry-based tasks using computer software to support the learning of
foundationl studies students in two mathematical topics, Complex Numbers and
Matrices.
‡ For teacher-researchers to learn from the engagement of student partners in the
design process to gain access to student thinking and culture in preparing for their
peers.
‡ To study the use of tasks with foundation studies students and the contribution of
the tasks to their mathematical understanding.
Here the community included all the people mentioned above and the practice,
in initial stages, was task design in the two topics using the Autograph software.
Like the SYMBoL project above, Catalyst (not an acronym) provided insights to the
engagement of students (here the student partners) in the teaching process, showing
the increased mathematical understanding this generated for these students and the
associated insights provided for teachers in the project.
A series of whole community meetings guided the task design process. Coffee
and cake enabled an informality which, like the tea in SYMBoL, set the scene for
inclusion of all in a genuine dialogue about task design. Our expert inducted the
rest of the community into the software; the fact that we were all beginners with the
software allowed us to be learners together. The student partners had been sent the
lecturer’s notes in the two topics and had brought themselves up to date with the
mathematics. After the first meeting, student partners went away in two pairs, each
to one of the topics, to start their task design. (Details can be found in Jaworski et
al., 2018.)
Two examples demonstrate issues and tensions in the project. The teacher-
researchers envisaged interactive tasks through which foundation studies students
would engage with mathematical inquiry in the topics. The first iteration on the
tasks, from the student partners, consisted of tasks in which a question was asked
and software engagement led to an expected answer. Student partners told us later
that their expectation of tasks was related to tasks they had experienced in their
own learning of mathematics. The tension that arose for teacher-researchers was
how to signal expectations of more dynamic tasks while valuing what had been
produced. One teacher researcher used the word “static” in contrast with a hope for

293

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

more dynamic tasks. This word, as we later learned, proved to be a catalyst in the
design process. Student partners could now visualize better what was required. The
next tranche of tasks were indeed more dynamic and were accepted as a basis for
foundation studies students’ activity.
These more dynamic tasks were raw, in the sense that they involved Autograph
animations without associated instructions for the foundation studies students. So,
two teacher-researchers wrote a set of instructions. One question asked students
to identify the ‘relationship’ between three complex numbers (z1, z2, z3) seen in
animation in the Argand plane. The word ‘relationship’ caused problems because,
while the foundation studies students could identify that z2=z1+z3, they had no idea
what a ‘relationship’ was. For the teacher-researchers, the language around dynamic
tasks and their use with students became a learning issue.
Both of these examples involve issues of communication. For the teacher-
researchers learning how to talk with students, whether student partners or foundation
studies students, about mathematics and the pedagogy of tasks, created challenges.
Our knowledge and understanding developed in relation to these challenges. For
the student partners there was a similar developmental process. For a conference
presentation in which they were asked to talk about what they had learned through
the project, they commented on deeper learning of mathematics; insights into what
it means to learn and quality of relationships with teachers. We ask the question of
how the project’s practices, which generated such learning, might become part of the
teaching-learning process more generally.
Here again is a Teaching Triad analysis is of the project as a whole in contributing
to the learning of all its participants and ultimately the understanding of the
mathematical topics by Foundation Students.
As with SYMBoL, the community of inquiry included both teachers and
students working together for a common purpose and with important learning for
both leading to developments in understanding and engagement. Both experienced
critical alignment in adapting to new forms of practice and related communication.
The nature of the practice, which was new for both teacher-researchers and student
partners meant that there were not clear expectations of alignment. Both groups were
exploring what communication and collaboration could look like. In this, they were
generating the practice as it progressed. Both groups demonstrated a willingness to
engage with each other that was in itself an inquiry process.

THE ROLE OF THE TEACHING TRIAD

In all three cases, use of the Teaching Triad had a dual role. First, it imposed the
three domains Management of Learning, Sensitivity to Students and Mathematical
Challenges onto the analysis of data. Previous research had studied and analysed the
relevance of these domains in characterising teaching which was not only designed
to teach mathematical topics, but also to teach in such ways that students could be
simultaneously supported and challenged to engage. Second, it allowed researchers

294

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

Table 10.3. Teaching Triad analysis of learning in the Catalyst Project

Triad Elements of Teaching Practice Relations with Students

Management ‡ Employing student partners ‡ Aiming to gain insight into


of Learning to design computer-based students’ perceptions of
(Project mathematical tasks mathematical learning
Managment) ‡ Providing a friendly respectful ‡ Creating a community of inquiry
atmosphere for iterative task in which participants could talk to
design by student partners for each other in informal ways and
foundation studies students not be afraid to express ideas
‡ Organising raw tasks for ‡ Encouraging the student partners
foundation studies students, creative thinking about the tasks
computer labs, software and ‡ Providing (minimal) instructions
screen capture to foundation studies students so
‡ Organising foundation studies that they could engage with the
students for use of prepared tasks computer-based tasks
Sensitivity to ‡ Creating friendly respectful ‡ Building on SYMBoL
Students relations between student partners experience – talking together over
and staff (social) tea and cake
‡ Talking about mathematics, ‡ Drawing student partners into
software and task design at an discussion using Autograph
‘equal’ level (teacher-researchers software as a catalyst to express
and student partners) (affective ideas.
and cognitive) ‡ Encouraging foundation studies
‡ Use of computer-based tasks students to visualize, and thus
in tutorials to offer visual engage with mathematical
and animated mathematical situations through use of the
experience (cognitive) computer-based tasks.
Mathematical ‡ Student partners designing tasks ‡ In order to produce dynamic
Challenge in an area of mathematics that tasks, student partners had to
they have not understood well understand the mathematical
before concepts better than they had
‡ Student partners using new earlier.
software with mathematics to ‡ Both student partners and
create tasks foundation studies students need
‡ foundation studies students to engage with mathematics more
engaging with new software conceptually: through designing
to learn new concepts in or working with the tasks
mathematics. ‡ Teacher-researchers learn
‡ Teacher-researchers finding to express instructions and
language in which to associated concepts in ways that
communicate mathematical can make sense for students.
exploration with both student
partners and foundation studies
students

295

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

to identify specific aspects of the didactic/pedagogic process that contributed to each


domain and the issues that arose, particularly, in fulfilling objectives of sensitivity
and challenge.
For example, in the Engineering Students Understanding Mathematics project,
activity in small groups with GeoGebra was designed (sensitively) to provide
opportunity for students to engage conceptually with mathematics through discussion
in the group and visualisation using GeoGebra. In the event, some students were
not able to sustain group mathematical dialogue or use visualisation possibilities
effectively. The Teaching Triad analysis drew teacher-researchers’ attention to the
issues and tensions in what had appeared to them as sensitive pedagogy (addressing
the mathematical challenges in engaging conceptually) in the light of some students’
apparent inability to engage. This drew attention to institutional and cultural factors
such as students’ school acculturation to procedural tasks, and the demands of
engaging more conceptually in university mathematics. Clearly sensitivity needed
to address more overtly these institutional and cultural factors.

A COMMUNITY OF INQUIRY ANALYSIS OF THE THREE CASES

In terms of the three-level model of inquiry-based practice:

Inquiry between Teachers and Students in Teaching-Learning Activity

We see this in different ways in all three cases. The Engineering Students
Understanding Mathematics project was founded on inquiry-based questioning in
lectures and tutorials. It was found, for some but not all students, that such questioning
resulted in a more overt level of engagement with mathematics, in lectures and in
tutorials, a deeper quality of thinking about mathematical concepts as students
addressed specially prepared questions in tutorials. In SYMBoL, the interns gained
insight from their design of resources into the teaching of mathematical concepts
that aided their own understanding of these concepts. The new resources were made
available and used by the peer leaders, so that second-year students gained from the
peer teaching and use of resources, to succeed well in their final assessment. In the
Catalyst project, the collaboration in preparation of tasks by the student-partners and
teacher-researchers was an inquiry process leading to new learning of mathematics
for the student partners, with the foundation studies students gaining ultimately from
use of new more visual tasks to support their mathematical learning.

Teachers Engaging in Professional Inquiry to Learn More about Creating


Mathematical Opportunities for Students

We see this also in all three projects. In the Engineering Students Understanding
Mathematics project, the teaching team engaged overtly in creating inquiry-based
tasks for students and reflecting on the teaching and learning that resulted. This

296

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

reflective learning both fed back into ongoing teaching and was the basis of study
within the project. In SYMBoL, the mathematician lecturers learned from their
inquiry with the student interns about the inherent difficulties in their modules and
about the design of resources which could help student understanding. The peer
leaders gained mathematical insights from their teaching of second year peers,
an implicit inquiry process. In Catalyst, teachers and student-partners inquired
into the production of mathematical tasks for the foundation studies students and
learned from the process respectively in terms of tasks design and its outcomes,
and of the mathematics in task design.

Inquiry in the Research Process to Learn More about Practical Implications and
Issues for Mathematical Development

Activity in all three projects involved developmental inquiry in a research process


involving new ways of engaging students in mathematics and new kinds of tasks, their
nature and use. We see here examples of developmental research projects in which
the overt development of materials or resources, designed to contribute to students’
learning, were significant in the learning of the designers and those conducting the
research. There is no doubt that these people learned significantly from the processes
involved. An important issue here is the extent to which this learning is generalisable
to other settings or transferable to other educators. Issues of scale appear here as we
think about ways in which, for example, student partners learning in task design can
be extended to larger groups or cohorts of students learning mathematics. These
issues are addressed further below.
Ideas of Community of Inquiry and Critical Alignment pervaded the three projects.
In Engineering Students Understanding Mathematics project the community of
inquiry was the teaching team, inquiring into new ways of engaging students to learn
mathematics more conceptually. A problem to some extent lay in the fact that the
students were not in any sense partners in this endeavour, and belonged to a culture
in which the value of their experience was not appreciated beyond the marks they
achieved. Thus, for the students, experiencing the innovation, critical alignment could
be seen in the activity of those who engaged well with inquiry-based tasks in tutorials,
and in the very good writing in the group projects. However, critical alignment was
resisted by students who were not able or willing to engage at the various levels.
Their previous experience and cultural embeddedness resulted in little obvious gain
from the innovation, as also made clear in post-teaching interviews (Jaworski et al.,
2012). Critical alignment for the teachers came through the recognition of the very
different ways in which students perceived the innovation. This was very significant
for the teaching team and led to some changes for future student cohorts; however,
we do not have data from these. In SYMBoL, the community of inquiry consisted
of mathematics staff including the lecturers and the student interns. Together they
learned to collaborate in mathematical activity and discussion which ultimately
benefitted the second-year students (use of the new resources to aid understanding,

297

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

and the peer tutorials). To some extent the peer leaders were also drawn into the
community of inquiry. Critical alignment could be seen in new ways of perceiving
the mathematical topics as a basis for student learning, as well as the use of new
resources to support their peer-teaching (Duah, 2017). Mathematics staff gained new
insights into students’ ways of thinking about mathematics and in finding ways to talk
mathematics with the interns. Sadly, we do not have data from subsequent lecturing
in the modules to report on modifications resulting from the project. In Catalyst,
the community of inquiry included the teacher-researchers, student-partners and
Analytical Assistants in creating activity resulting in new tasks for the foundation
studies students. The design process and its associated practical implementation led
to learning for all of these involved. Critical alignment could be seen in observations
of the use of the tasks by foundation studies students and in modifications required
for creation of tasks at both the student partner and teacher levels of preparation. The
lecturer has modified her practice both in lectures and in use of tasks in her tutorials
as a result of learning in the project and data is still being collected.

RESEARCH OUTCOMES AND DEVELOPMENTAL


LEARNING BEYOND THE PROJECTS

In any research project, publication of results and outcomes allows practitioners


beyond the projects to learn from what is reported. Our published papers in all the
above contribute to this. However, all that is reported here is small scale, and the very
nature of the projects restricts what is possible within the available resource/funding.
This was true also of the projects reported in the literature addressed earlier. The
funders of the Catalyst project were explicitly interested in dissemination beyond
the project itself and ways in which such dissemination could be instrumental in
spreading the knowledge and learning from the project. Those of us involved in
these projects are concerned to address how the significant small-scale learning
can be extended, particularly to larger cohorts of students. In SYMBoL, this started
to happen through the teaching of the peer leaders in tutorials for the second-year
students, with some success as reported above. In all three projects, the design
process and its subsequent modification resulted in resources which could be used
with future cohorts of students to promote a deeper quality of mathematics learning.
The most exciting findings are those in which we see student partners (or interns)
learning mathematics through deeper engagement with mathematics promoted
through their relationships with mathematics staff and teachers. In both SYMBoL
and Catalyst, this was very small scale, four students in each case. We ask the question
as to how we might engage more students as student partners with mathematics
teachers, so that their engagement in the teaching process will contribute to their
deeper learnings of mathematics. In essence, we would like all students to act as
teachers to learn mathematics with deeper quality. We are starting to engage with
this challenge in creating future projects.

298

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

A WAY AHEAD?

As a final codicil, I refer again to the recent paper by Wenger-Trayner and Wenger-
Trayner (2018). Their focus here is to develop the earlier work on Communities of
Practice to consider Landscapes of Practice and associated Knowledgeability. In an
interview with Etienner Wenger, Omidvar, & Kislov, 2014, write: “Wenger-Trayner
elaborates on the notion of knowledgeability as a relationship individuals establish
with respect to a landscape of practice that makes them recognizable as legitimate
actors in complex social systems” (p. 14).
Very briefly, a landscape of practice, according to Wenger-Trayner (cited in
Omidvar & Kislov, 2014, p. 15) elaborates on the notion of knowledgeability as
a relationship that individuals establish with respect to a landscape of practice that
makes them recognizable as legitimate actors in complex social systems. Further,
from this interview, Wenger-Trayner is cited as saying (p. 15):
Professional occupations, and indeed most non-professional endeavors, are
constituted by a complex landscape of different communities of practice –
involved not only in practicing the occupation, but also in research, teaching,
management, regulation, associations, and many other relevant dimensions. All
these practices have their own histories, domains, and regimes of competence.
Wenger-Trayner, further, expresses the idea of ‘knowledgeability’ within a
landscape: if a body of knowledge is a landscape of practice then, for each person
acting within the landscape, their personal experience of learning can be thought of
as a journey through this landscape. Wenger-Trayner is quoted further:
As a trajectory through a social landscape, learning is not merely the acquisition
of knowledge. It is the becoming of a person who inhabits the landscape with
an identity whose dynamic construction reflects our trajectory through that
landscape. (p. 19)
Briefly, although not doing justice to the complex arguments here, individual
knowledgeability is the totality of the experience of identity within the landscape,
where the landscape itself maps out the full range of possible experiences, including
tensions and contradictions. A final quotation from Wenger-Trayner:
We will use the term knowledgeability to refer to the complex relationships
people establish with respect to a landscape of practice, which make them
recognizable as reliable sources of information or legitimate providers of
services. (p. 23)
It perhaps does not require a great stretch of imagination to see how these thoughts
can be applied to the landscape of mathematics learning and teaching through
engagement in inquiry communities across local and international boundaries. The
activities described in the cases present some local activity from the United Kingdom,
while examples from the literature review take us to research activity in the rest of

299

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

the world. If we are to think of ‘scale’ rather in terms of social landscapes of practice
across international boundaries, than in terms of large-scale research in individual
situations or institutions, then the task becomes not only different but do-able over
time. As the research literature in mathematics learning and teaching at university
level grows, with researchers taking on the theoretical mantle of communities of
inquiry within landscapes of practice, we may start to discern a new paradigm in
our field. That is, the gaining of insight into practices in mathematics teaching
that engage students in mathematical inquiry, resulting in greater conceptual depth
with less reliance on procedure and memory. We aim to see students’ mathematical
knowledgeability grow with respect to inquiry-based practices and have much to
learn about the practices we employ, the relationships we engender, the tensions
and contradictions that inquiry reveals, and the inclusion of students in our teaching
enterprises. This learning is not local to the United Kingdom, or any other nation,
but a project for us all.

NOTES
1
The colleagues to whom I refer are many and, where I use pronouns “we” or “our,” I include this
wider group. I need to acknowledge especially the contribution of Simon Goodchild and Despina
Potari in developing concepts discussed in this chapter.
2
Space does not permit a detailed exposition of Wenger’s constructs. Readers are directed to chapters
2 and 8 of Wenger (1998).
3
These studies are already in the public domain as referenced. However, the author here was closely
associated with the three studies [1. As project leader; 2. As supervisor of the PhD researcher, 3. As
project leader] and can therefore draw on examples not necessarily reported in published material.

REFERENCES
Abdulwahed, M., Jaworski, B., & Crawford, A. R. (2012). Innovative approaches to teaching mathematics
in higher education: A review and critique. Nordic Studies in Mathematics Education, 17(2), 49–68.
Alsina, C. (2001). Why the professor must be a stimulating teacher: Towards a new paradigm of teaching
mathematics at university level. In D. Holton, M. Artigue, U. Kirchgräber, J. Hillel, M. Niss, & A.
Schoenfeld (Eds.), The teaching and learning of mathematics at university level (pp. 3–12). Dordrecht,
The Netherlands: Kluwer Academic.
Barton, B., Oates, G., Paterson, J., & Thomas, M. (2014). A marriage of continuance: Professional
development for mathematics lecturers. Mathematics Education Research Journal, 27(2), 147–164.
Biza, I., Jaworski, B., & Hemmi, K. (2014). Communities in university mathematics. Research in
Mathematics Education, 16(2), 161–176.
Duah, F. (2017). Students as partners and students as change agents in the context of university
mathematics (Unpublished PhD thesis). Loughborough University, Leicestershire, United Kingdom.
Duah, F., Croft, T., & Inglis, M. (2014). Can peer assisted learning be effective in undergraduate
mathematics? International Journal of Mathematical Education in Science and Technology, 45(4),
552–565.
Goodchild, S. (2014). Theorising community of practice and community of inquiry in the context of
teaching-learning mathematics at university. Research in Mathematics Education, 16(2), 177–181.
Goodchild, S., Fuglestad, A. B., & Jaworski, B. (2013). Critical alignment in inquiry-based practice in
developing mathematics teaching. Educational Studies in Mathematics, 84(3), 393–412.
Hawkes, T., & Savage, M. (Eds.). (2000). Measuring the mathematics problem. London: Engineering
Council.

300

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
UNIVERSITY MATHEMATICS TEACHING DEVELOPMENT

Jaworski, B. (1994). Investigating mathematics teaching: A constructivist enquiry. London: Falmer.


Jaworski, B. (2006). Theory and practice in mathematics teaching development: Critical inquiry as a
mode of learning in teaching. Journal of Mathematics Teacher Education, 9(2), 187–211.
Jaworski, B. (2014). Unifying complexity in mathematics teaching-learning development: A theory-
practice dialectic. In Y. Li, E. A. Silver, & S. Li (Eds.), Transforming mathematics instruction:
Multiple approaches and practices (pp. 439–458). New York, NY: Springer.
Jaworski, B. (2015). Intersubjectivity in mathematics teaching: Meaning-making from constructivist and/
or sociocultural perspectives? In P. Gates & R Jorgensen (Eds.), Shifts in the field of mathematics
education: Stephen Lerman and the turn to the social (pp.171–184). New York, NY: Springer.
Jaworski, B., & Matthews, J. (2011). How we teach mathematics: Discourses on/in university teaching. In
M. Pytlak, T. Rowland, & E. Swoboda (Eds.), Proceedings of the Seventh Congress of the European
Society for Research in Mathematics Education. CERME7 (pp. 2022–2032). Rzeszów, Poland:
University of Rzeszów and ERME.
Jaworski, B., & Potari, D. (2009). Bridging the macro-micro divide: Using an activity theory model to
capture complexity in mathematics teaching and its development. Educational Studies in Mathematics,
72(2), 219–236.
Jaworski, B., Robinson, C., Matthews, J., Croft, A. C. (2012). An activity theory analysis of teaching
goals versus student epistemological positions. International Journal of Technology in Mathematics
Education, 19(4), 147–152.
Jaworski, B., Treffert-Thomas, S., Hewitt, D., Feeney, M., Shrish-Thapa, D., Conniffe, D., … Anastasakis,
M. (2018). Student partners in task design in a computer medium to promote foundation student’
learning of mathematics. In. V. Durand-Guerrier, R. Hochmuth, S. Goodchild, & N. M. Hogstad
(Eds.), Proceedings of INDRUM 2018: Second Conference of the International Network for Didactic
Research in University Mathematics (pp. 316–325). Kristiansand, Norway: University of Agder &
INDRUM.
Johnson, E. M. S., & Larsen, S. P. (2012). Teacher listening: The role of knowledge of content and
students. Journal of Mathematical Behavior, 31(1), 117–129.
Kane, R., Sandretto, S., & Heath, C. (2002) Telling half the story: A critical review of research on the
teaching beliefs and practices of university academics. Review of Educational Research, 72, 177–228.
http://dx.doi.org/10.3102/00346543072002177
Lave, J., & Wenger, E. (1991). Situated learning: Legitimate peripheral participation. Cambridge:
Cambridge University Press.
Nardi, E. (1996). The novice mathematician’s encounter with mathematical abstraction: Tensions in
concept-image construction and formalisation (Unpublished doctoral thesis). University of Oxford,
Oxford.
Nardi, E. (2008). Amongst mathematicians: Teaching and learning mathematics at the university level.
New York, NY: Springer.
Omidvar, O., & Kislov, R. (2014). The evolution of the communities of practice approach: Toward
knowledgeability in a landscape of practice: An interview with Etienne Wenger-Trayner. Journal of
Management Inquiry, 23(3), 14–23, doi:10.1177/1056492613505908
Petropoulou, G., Potari, D., & Zachariades, T. (2011). Inquiring mathematics teaching at the university
level. In B. Ubuz (Ed.), Proceedings of the 35th Conference of the International Group for the
psychology of mathematics education (Vol. 3, pp. 386–392). Ankara, Turkey: PME.
Potari, D., & Jaworski, B. (2002). Tackling complexity in mathematics teacher development: Using the
teaching triad as a tool for reflection and enquiry. Journal of Mathematics Teacher Education, 5(4),
351–380.
Rasmussen, C., & Wawro, M., (2017). Post-calculus research in undergraduate mathematics education. In
J. Cai (Ed.), Compendium for research in mathematics education (pp. 551–581). Reston, VA: National
Council of Teachers of Mathematics.
Schoenfeld, A. H. (2010). How we think: A theory of goal-oriented decision making and its educational
applications. New York, NY: Routledge.

301

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
BARBARA JAWORSKI

Speer, N. M., & Wagner, J. F. (2009). Knowledge needed by a teacher to provide analytic scaffolding
during undergraduate mathematics classroom discussions. Journal for Research in Mathematics
Education, 40(5), 530–562.
Speer, N. M., Smith III, J. P., & Horvath, A. (2010). Collegiate mathematics teaching: An unexamined
practice. Journal of Mathematical Behavior, 29, 99–114.
Stenhouse, L. (1984). Evaluating curriculum evaluation. In C. Adelman (Ed.), The politics and ethics of
evaluation (pp. 77–86). London: Croom Helm.
Treffert-Thomas, S., & Jaworski, B. (2015). Developing mathematics teaching: What can we learn from
the literature? In M. Grove, T. Croft, J. Kyle, & D. Lawson (Eds.), Transitions in undergraduate
mathematics (pp. 259–276). Birmingham: University of Birmingham with Higher Education Academy.
Wagner, J. F., Speer, N. M., & Rossa, B. (2007). Beyond mathematical content knowledge: A
mathematician’s knowledge needed for teaching an inquiry-oriented differential equations course.
Journal of Mathematical Behavior, 26, 247–266.
Wells, G. (1999). Dialogic Inquiry: Towards a socio-cultural practice and theory of education.
Cambridge: Cambridge University Press.
Wenger, E. (1998). Communities of practice: Learning, meaning and identity. Cambridge: Cambridge
University Press.
Wenger-Trayner, E., & Wenger-Trayner, B. (2018). Learning in a landscape of practice: A framework. In
E. Wenger-Trayner, M. Fenton-O’Creevy, S. Hutchinson, C. Kubiak, & B. Wenger-Trayner (Eds.),
Learning in landscapes of practice: Boundaries, identity, and knowledgeability in practice-based
learning (pp. 13–29). London: Routledge.
Winsløw, C., Gueudet, G., Hochmut, R., & Nardi, E. (2018). Research on university mathematics
education. In T. Dreyfus, M. Artigue, D. Potari, S. Prediger, & K. Ruthven (Eds.), Developing research
in mathematics education. Twenty years of communication, cooperation and collaboration in Europe
(pp. 60–74). Oxon: Routledge.

Barbara Jaworski
Mathematics Education Centre
University of Loughborough

302

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI AND KONSTANTINOS STOURAITIS

11. TEACHER DECISION MAKING


Developments in Research and Theory

In this chapter we address research on decision making of mathematics teachers. We


attempt to trace shifts that have occurred in the research literature in relation to the
focus and theoretical perspectives of teacher decision making, discussing about the
links that this area of research in mathematics education has with others, such as
mathematics teacher noticing and mathematics teacher knowledge. In particular, we
discuss the nature of teacher decision making, its role in planning lessons and during
classroom teaching, factors that frame teacher decision making, and attempts that
have been made to support teachers to make decisions that could facilitate students’
learning. Finally, we present an example of our empirical study on teacher decision
making while reflecting on mathematics teaching enacted in the context of a new
curriculum to illustrate an Activity Theory perspective to study teacher decision
making.

INTRODUCTION

Teachers make decisions about teaching all the time both in planning teaching
and in interacting with the students in the classroom. Decision making has not
always been an explicit focus of research in mathematics education. However, it is
studied through teachers’ designs and the rationale behind them (Pepin, Gueudet,
& Trouche, 2013), through teachers’ in the moment actions (Stockero, Leatham,
Ochieng, Van Zoest, & Peterson, 2019) and through teachers’ reflections on their
own and others’ teaching and in particular while considering further teaching actions
(Sherin & Van Es, 2005). In the book edited by Clarkson and Presmeg (2008) on the
major contributions of Alan Bishop, decision making is a critical part of his work
since the beginning of his research career. He considers decision making at “the
heart of the teaching process” (Bishop, 1976, p. 42). Bishop argues that if we know
about teachers’ decisions, we can link teaching to a number of different aspects (e.g.,
objectives, intentions, children’s attitudes, children’s mathematical development)
and so search for ways of improving its quality. He also discusses the complexity of
the situation when the teacher needs to respond to a child’s question and talks about
practices that the teacher often uses in these circumstances, such as, for example,
“time-buying” (Bishop, 1976, p. 46).

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_012

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

In the same book Borko, Roberts, and Shavelson (2008) discuss how the work
of Bishop and Whitfield (1972), Shulman and Elstein (1975) and Shavelson
(1973), in relation to decision making, has influenced the field and how their
ideas have been extended. The authors argue that these researchers adopted a
cognitive perspective and “conceptualized teaching decisions as the fundamental
link between complex, real-time teaching situations and practical actions in
classrooms, postulating a cognitive framework or schema that underlies the link”
(p. 39). The authors refer to the frameworks that these researchers developed
about the process of decision making. They argue that Bishop and Whitfield
(1972) consider that prior experiences of the teachers filtered through their beliefs
and values and their lesson goals are the bases of teachers’ decision making.
They point out that the work and theorizing of Bishop and Whitfiled originated
from the actual teaching situations where teacher makes decisions while the
work of Shulman (with Elstein) and Shavelson started from theory and used it to
understand practice. Shavelson considers teachers’ decision making to integrate
information about students, subject matter, school and classroom environment
filtering with teachers’ beliefs and conceptions of the subject matter. All three
researchers recognize that usually decision making is made automatically while
in unexpected situations teachers’ consciousness of their decisions and past and
current judgments can influence subsequent decisions. Shavelson and Shulman
also point out that teachers’ beliefs about the nature of learning and teaching and
judgments of students’ knowledge, attitudes and behavior and expectations of
their performance also influence planning and consequently teaching.
Borko et al. (2008) also discuss studies that examine teachers’ decisions in
planning and interactive decisions and the methodologies adopted for studying
them. They also point out that some studies in the period of 1980 to 1990 begin
to recognize also external forces that seem to frame decision making and teaching
(they refer to an ethnographic research on reading of Borko & Eisenhart, 1989). In
that period, decision making extends to the area of teachers’ professional knowledge
and especially contrasting experiences with novice mathematics teachers. Borko
et al. comment that mathematics educators nowadays may not talk about decision
making, as they talk about teacher knowledge, building upon student thinking and
using artifacts and resources to improve teaching.
Mathematics teacher decision making was rather a non explicit focus of research
on mathematics teacher and mathematics teaching the decade before the publication
of the first edition of the Handbook of Mathematics Teacher Education in 2008.
However, it seems that currently there are some attempts to re-visit teacher decision
making extending the lenses that it is studied and understood. In this chapter, we
first address these attempts through a search in the literature for the last ten years.
We structure this review on the following issues: (a) the nature of decision making
process, (b) the interplay of decision making with perception and interpretation in
the context of noticing teaching phenomena, (c) the relation of decision making
to teacher knowledge, (d) decision making in planning teaching using curricular

304

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

resources, (e) mathematics teacher education approaches that support decision


making that can facilitate students’ learning and (d) on the factors that frame decision
making from individual and social perspectives. Second, we offer an example from
our own work to illustrate how the context and the sources upon which the teacher
bases his or her decisions frame the decision making process based on an Activity
Theory (AT) perspective. Finally, we discuss what we have learned from research
on teacher decision making and suggest future directions that this research needs
to take.

LITERATURE REVIEW

The Nature of Decision Making Process

A number of studies have focused on the nature of teacher decision making process.
The main issues that have been investigated are a) whether decisions are conscious
or unconscious b) what triggers conscious decision making and c) the process
itself. Watson (2019) refers to the work of the cognitive psychologists Evans and
Stanovich (2013) to characterize decision making as an unconscious and intuitive
Type 1 process or rational and conscious deliberative Type 2. Early research on
decision making shows that the first type of processes is mainly found on interactive
decision making and the second on planning and reflecting on the lesson. However,
more recent studies do not accept this dichotomic view and suggest that the nature of
decision making is somewhere between type 1 and type 2 reasoning. Watson (2019)
accepts the construct of algorithmic reasoning of Stanovich et al. (2011) to describe
a conscious process that makes use of heuristics and pre-established routines that the
teacher has developed through his or her long teaching experience. He claims that
teacher decision making needs to be seen through a multidisciplinary approach that
takes into account these culturally embedded routines and also affective aspects on
the basis of which the teacher makes decisions.
Schoenfeld (2011) from his research on problem solving and from analyzing
teacher-student interactions claims that decisions are consistent with the teachers’
goals, consciously or unconsciously. The decision making process is relatively
automatic when the situation is familiar to the teacher and routines are enacted while
in unfamiliar situations decision making can be modeled on the basis of orientations
of the individual. The implementation of decisions and their monitoring when the
situation triggers decision making takes place on an ongoing basis and the process
is iterative.
Engeström (2001b) adopts a more systemic approach to decision making and
identifies four dimensions of decision making: social-spatial, anticipatory-temporal,
moral-ideological and systemic-developmental. He argues that:
Decisions are not made alone, they are indirectly or directly influenced by
other participants of the activity. Decisions are typically steps in a temporally
distributed chain of interconnected events. Decisions are not purely technical,

305

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

they have moral and ideological underpinnings with regard to responsibility


and power. And the content of decisions is not restricted to the ostensible
problem or task at hand; they always also shape the future of the broader
activity system within which they are made. (p. 281)
Although this classification has been used by Engeström in health care, later in this
chapter we illustrate how we used his theoretical framework in studying mathematics
teachers’ decision making in the process or enacted a new curriculum.

Teacher Decision Making and Teacher Noticing

In-the-moment teacher decision making during classroom teaching has been a main
focus of research in mathematics education in the last decade. Stahnke, Schueler,
and Roesken-Winter (2016) conducted a literature review of teachers’ perception,
interpretation and decision-making. They considered decision making as one part
of this triplet and refer to Kaiser et al. (2015) who considered decision making
either as anticipating a response to students’ activities or as proposing alternative
instructional strategies. They also argued that decision making is a part of teachers’
noticing as it is related to how you act after interpreting what you notice. The
authors found about 32 papers published from 2007 to 2016 addressing decision
making. In most of the papers, decision making was studied in situations where
the teacher responded to specific classroom situations and in particular to students’
contributions. These situations were often made available to teachers through videos
in a video club setting. The results indicated that decision making was studied mainly
as a component of teachers’ noticing and it was particularly difficult for prospective
teachers. Most studies focused on in-the-moment decision making through case
studies and provided evidence for its complexity.
Research on teacher noticing has been extensive in mathematics education during
the last ten years. Mason (2002) introduced noticing to students’ learning process,
linking it to teacher awareness both in mathematics and in mathematics teaching.
In later developed frameworks, teacher noticing is considered as an activity that
involves description of a phenomenon, interpretation and proposition or enactment
of alternative teaching actions (van Es & Sherin, 2002). In most studies, teacher
noticing is related to students’ thinking, its interpretation and responding to it (see
for example Jacobs, Lamb, & Phillip, 2010). Teacher decision making is triggered
usually on what Stockero and Van Zoest (2013) call pivotal teaching moments
(PTM). The authors analyzed what types of decisions do novice mathematics teachers
make in these pivotal teaching moments when they occur in their teaching and on
what potential impact these decisions can have on student learning. The decisions
were categorized in relation to the teacher’s actions (extends mathematics, pursues
student thinking, emphasizes meaning, acknowledges pivotal teaching moments
but continues as planned or ignores/dismisses pivotal teaching moments), to the
degree of managing these actions (skilfully, moderately, poorly) and to the potential

306

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

impact on student learning (high positive, medium positive, low positive, neutral
and negative). They came to the conclusion that linking decision making to pivotal
teaching moments and to their impact on student learning is an important direction
that research on decision making can take.
The study of Van Zoest, Peterson, Leatham, and Stockero (2016) elaborates
further the relationship between decision making and productive teaching
actions by paying attention to teachers’ decisions on how to respond to students’
mathematical thinking. They recognize the complexity of this part of the noticing
especially in the whole class setting and they theorize this process as a collection
of teacher moves. These moves can be often unproductive especially in cases
where the teacher evaluates students’ contributions or uses funnelling questions.
The authors emphasize the need to respond to student thinking at the moment in
which it occurs rather than to monitor, select student work and then sequencing
the presentation of it later on in the lesson as Smith and Stein (2011) propose.
They consider the teaching practice of “building” as a productive decision in
acting on the moment. They conceptualize the teaching practice of building not
as a specific teaching action but as “several teacher moves woven together to
engage students in the intellectual work of making connections between ideas and
abstracting mathematical concepts from consideration of their peers’ mathematical
thinking” (pp. 1284–1285). So, to characterize teaching decisions as productive,
we need to study a series of moves that may result to the building goal of students
to understand important mathematical ideas. Van Zoest et al. also indicate that
the building of teaching practice consists of four sub-practices: make student
mathematical thinking clear; turn student mathematical thinking over to the
students in the class; orchestrating the classroom discussion in making sense of the
student mathematical thinking; and facilitate the extraction and articulation of the
important mathematical idea. This study of van Zoest et al. (2016) is an example
of how decision making as a part of noticing is linked to its relation and impact on
the quality of mathematics teaching.

Teacher Decision Making and Teacher Knowledge

Stahnke et al. (2016) link decision making and teacher professional knowledge.
Teacher knowledge has been considered both as a resource for decision making but
also decision making has been considered as a lens for studying teacher knowledge.
For example, the Knowledge Quartet (Rowland, Huckestep, & Thwaites, 2005),
discussed also in this volume of the Handbook by Tim Rowland and other chapter
authors, includes contingency as one unit related to the knowledge of the teacher to
respond to unexpected classroom situations. The two components of this category
are related to teacher’s “readiness to respond to children’s ideas and a consequent
preparedness, when appropriate, to deviate from an agenda set out when the
lesson was prepared” (p. 263). Rowland et al. relate the way that the teacher
responds to these contingency moments with the knowledge recourse available

307

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

to the teacher and to the use of opportunities for promoting students’ learning.
The codes for the contingency category (see the chapter by Tim Rowland, this
volume of the Handbook), developed through research using the Knowledge
Quartet, include also teacher insight during instruction and responding to the
availability of tools and resources as important elements of contingency. These
elements seem to be connected to teacher awareness and to the resources, which
as we have already discussed, in Schoenfeld’s (2011) framework are elements
of teacher decision making. In the Ball and colleagues’ (e.g., Ball, Thames, &
Phelps, 2008) framework of mathematical knowledge for teaching, the domains
of knowledge are also linked to teachers’ actions and decisions. For example, in
Ball et al., the subdomain of knowledge of content and teaching that combines
knowing about teaching and about mathematics is explicitly linked to teachers’
instructional decisions. In the following excerpt, the complex process of teacher
decision making is illustrated and is linked to the domain of knowledge of content
and teaching:
During a classroom discussion, a teacher must decide when to pause for more
clarification, when to use a student’s remark to make a mathematical point, and
when to ask a new question or pose a new task to further students’ learning.
Each of these decisions requires coordination between the mathematics at
stake and the instructional options and purposes at play. (p. 401)

Teacher Decision Making and Curricular Resources

Teachers’ use of curricular materials is also a part of the literature on teachers’


decision making. Considering teachers as active agents and designers, who shape the
enacted curriculum alongside their students and the materials, their decisions related
to the curricular materials have gained some attention. Brown (2009) refers to some
ways in which teachers interact with curriculum artifacts. Teachers select materials,
making “day-to-day decisions about which of the program’s available resources to
use” (p. 22). They interpret these materials and their interpretations are influenced
by their own capacities as well as features of the context. Often teachers modify
or omit parts of the materials and add their own makings as also suggested by the
following.
Lloyd (2008) studied the first two years of a teacher’s use of a new curriculum and
her findings include significant changes to classroom discourse towards the whole-
class instruction with which the teacher had the greatest experience. She concluded
that the teacher’s perception of students’ expectations and his own discomfort
associated with using the new curriculum were key factors in his decisions.
Choppin’s (2011) study focused on three teachers enacting instructional
sequences from curriculum materials, finding similarities and differences. One
of the teachers was developing adaptations in the use of the materials and the
researcher concluded that “learned adaptations suggest both an understanding of

308

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

the design rationale and empirically developed knowledge of how that rationale
plays out in practice” (p. 351).
Eisenmann and Even (2009) studied the differences of the algebraic activities
in two classes by the same teacher. They found that the students had different
opportunities to engage in global/meta-level algebraic activities during whole-class
work, and they connect the teacher’s decisions with her considerations about the
different classes and their capacities.
A more recent study by Earnest and Amador (2019) investigated how prospective
teachers use curriculum materials to design lesson plans and to visualize their
enactments. The study uses Gueudet and Trouche’s (2009) documentational genesis
to study the process of using the materials. The findings indicate that prospective
teachers decided to include, omit or adapt specific elements of the materials.
Prospective teachers’ rationale for such decisions was not always clear but it varied
from supporting students to learn mathematical ideas to student enjoyment of
mathematics.

Mathematics Teacher Education for Promoting “Productive” Decision Making

Different approaches have been developed that promote the development of


productive decision making, especially in the context of teacher noticing.
Encouraging teachers to use some of the developed frameworks on noticing
and decision making can be contexts for prospective and practicing teachers
in making decisions that could facilitate students’ learning. Researchers have
offered recommendations for using frameworks that have been developed. For
example, Stockero and Van Zoest (2013) suggest that asking teachers to identify
pivotal teaching moments and think about how the teacher might best respond to
promote student learning offer professional learning opportunities for novice and
experienced teachers. Thomas and Yoon (2014a) recommend that Schoenfeld’s
(2011) model of resources, orientations and goals could be an effective professional
development strategy for promoting teachers’ awareness of their in-the–moment
classroom decisions. However, few studies focus on the development of teachers’
awareness of their teaching decisions and their impact on students’ mathematical
learning.
Amador and Carter (2018) propose the structure of lesson study as a productive
context for developing prospective teachers’ capacity to observe, interpret and make
decisions on how to respond to students’ thinking. The Knowledge Quartet is another
example of a theoretical framework that has been used mostly with prospective
teachers as a main focus to support self reflection and develop teacher awareness of
their decisions and actions (Rowland & Turner, 2006).
However, in general, most attempts to consider decision making in mathematics
teacher education remain at the recommendation level; few examples were identified
from our search of the research literature.

309

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

Factors Framing Teacher Decision Making: Individual Perspectives

Concerning the factors that frame decision making, Stahnke et al. (2016) refer to
studies that provide evidence that teacher knowledge, beliefs and school climate
of trust have an impact on beginning teachers’ decision making. An example of a
study on the teachers’ decisions on selecting and enacting textbook problems (Son
& Kim, 2015) shows that teachers’ matching of their beliefs and goals with those
of the textbooks, their views on the textbook, their interpretation of curriculum and
assessment guidelines and their orientation toward student thinking are important
factors framing their decisions.
Schoenfeld (2011) offers a theoretical model of explaining how the teachers make
the moment-by-moment decisions focusing on the decision making process but also
on the factors that frame the process. He uses the constructs of goals, resources
and orientations to offer what he calls, coherent explanations of teachers’ decisions
and actions. He considers teaching as a goal-oriented activity where the teacher
has short- and long-term goals. A short-term goal can be responding to a student’s
response while a long-term response can be to prepare the students for an entrance
examination to the university. The way that teachers view their environment and
the way they react to this depends on their orientations that encompass “beliefs,
dispositions, values, tastes, and preferences” (Schoenfeld, 2011, p. 28). Schoenfeld
also points out that in most cases teachers are not aware of their orientations and
that they often make decisions on the basis of their routine activity. In this case
decision making is based on a default set of expectations that teachers have through
their teaching experience. Resources, according to Schoenfeld, include intellectual,
material and social resources where teacher knowledge is an important part of the
intellectual resources. Goals, orientations and resources that the teacher has are
activated when the teacher enters into a particular context. Schoenfeld describes
how the decisions are made and the interplay of the goals, resources and orientations
through a number of mathematics teaching situations. The model suggested by
Schoenfeld offers a way to scrutinize the process of decision making of a teacher
and address to some extent how this process is enacted.
Paterson, Thomas, and Taylor (2011) and Thomas and Yoon (2014a) use the
framework of Schoenfeld (2011) to study in-the-moment decision making. Paterson
et al. identify through two cases of university lecturers that their pedagogical goals
often conflicted with their goals as mathematicians in different ways of framing
different decisions. They concluded that these decisions are consistent with the
lecturers’ different orientations. One lecturer emphasized mathematical rigor,
aesthetics of the proof, accuracy and consistency of the mathematical notation,
the appropriate theoretical tools, and proof techniques for solving a mathematical
problem. The other lecturer paid more attention on the teaching and learning
processes, such as the appropriateness of the teaching materials, the exploration,
and the coverage of the syllabus. These orientations reflected on their decisions
and actions. In the study of Thomas and Yoon, the focus is on the decisions, goals,

310

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

resources and orientations of a secondary school mathematics teacher. The authors


identify four central goals that the teacher had that were conflicting each other at
different moments of the teaching (i.e., to prepare students to be successful in future
tasks; to use student–centered approaches; to fulfill the requirements of curriculum,
time and assessment; and to respect students’ cultural differences) and link them
to their orientations and resources. The study shows that these relationships are
complex and that there were cases where the constraint of time led the teacher to
take decisions that were different from his orientations.
Stockero et al. (2019) studied the role of teachers’ orientations for in-the-moment
decision making to unexpected students’ mathematical thinking. The authors used the
construct of thinking-as-a-recourse orientation that includes student mathematical
capability, student thinking informing instruction and the utility of student errors
in instruction and extended it to include a broader range of orientations. Some
orientations such as viewing students as capable of engaging with mathematical
ideas or believing in tapping into student thinking potentially support effective
decisions regarding teaching practices while orientations such as considering that
students do not need to respond to one another’s ideas or that student errors should
be avoided potentially hinder effective decision making.

Factors Framing Teacher Decision Making: Moving from Individual to Social


Perspectives

Considering decision making as a complex process requires lenses that could allow
us to study this complexity. Cooney (1988) argues that teacher decision making
both at the planning and at the classroom interaction stages is framed by a large
number of factors such as, for example, educational regulations, teacher’s objectives
for the lesson, expected use of the content in different realistic situations, students’
and teachers’ interest on the content, and predicted difficulty or even authoritative
judgments of professional groups and individuals. These factors go beyond cognitive
(e.g., students’ difficulties as reported in research), affective (e.g., emotions and
dispositions of students and teachers) or classroom management (ways of working,
established norms) situations. They also concern institutional, cultural, and social
factors that frame teachers’ decisions.
Most of the literature on teachers’ decision making is focused on the “what?”
question and is oriented towards individual and cognitive aspects of in-the-
moment decisions made in the classroom. Nevertheless, some of these researchers
acknowledge the need to broaden their vision to address the “why?” question, that is,
where teachers’ decisions come from. We have already discussed some studies that
attempt to address the sources on which teachers base their decisions focusing mainly
on teachers’ beliefs, and knowledge. Schoenfeld (2011) describes these sources in
more general terms but still with an individual and cognitive focus. Thomas and
Yoon (2014b), using Schoenfeld’s framework in their study, attempted to describe
the reasons of a teacher’s conflicting decisions regarding resources, orientations and

311

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

goals by providing interpretations of a social and systemic character. For example,


they included teacher’s inexperience of the specific curriculum approach and the
context of student preparation for the national assessment as factors influencing the
teacher’s decisions.
Lande and Mesa (2016) refer to planned and in-the-moment decisions of two
groups of mathematics teachers (full-time and part-time). They indicated that
different decisions are shaped by societal and institutional contexts. Considering
Ball et al.’s (2008) and Schoenfeld’s (2011) work, they stated that “Neither of these
programs or research investigates the conditions and constraints that impinge on
teaching or the decision-making process of individuals in the role of teacher” (p.
200). Seeking to understand why certain decisions are made, they investigated the
role of agency and the social context on teacher’s decision making.
Santagata and Yeh (2016) consider knowledge and beliefs as overlapping with
classroom practice, and that perception, interpretation, and decision-making skills
are at the centre of the overlapping space. These three skills are informed by
students, tools, and teacher communities. This reflects the authors’ “consideration
of what teachers know and believe, not as purely belonging to them and located
in their mind, but as socially and contextually negotiated” (p. 163). In their model
“communities become an intrinsic part of who teachers are and what they decide to
do in their classrooms” (p. 164).
Adopting a sociocultural perspective, Skott (2013) studying Anna, a teacher at the
beginning of her teaching career, finds four “significant practices or figured worlds”
that shape Anna’s meaning making and instructional decisions: ‘relationing’, ‘the
reform’, ‘mathematics’, and ‘teaming’ (p. 554). Skott uses the Patterns of Participation
approach “in an attempt to link the teacher’s contributions to the interaction to other
significant practices” (Skott, 2015, p. 16).
Using situated, social and sociocultural perspectives, researchers acknowledge the
restrictions and drawbacks related with cognitive perspectives that focus on teachers’
decisions as something happening in their minds in the flow of teaching. To a less
extent and in a different way, researchers using individual and cognitive perspectives
attempt to overcome such difficulties in looking for reasons for teachers’ decisions.
However, we view these approaches as having some limitations in addressing the
underlying reasons for teacher decision making. For example, first, there is the
view that teacher’s orientations and goals appear as stable and invariable structures;
knowing them one can even foresee teacher’s in-the-moment decisions. But why and
how these orientations and goals are formed and how can they explain differences
between different teachers in the same context or the same teacher in different
contexts? Second, resources are considered as a toolbox from which a teacher can
draw on any available tool. But the way the impact that tools have on the teacher and
teaching practice with their restrictions and affordances is underestimated. Finally,
the impact of teacher’s decisions is considered in the context of the classroom. But
teacher’s decisions can have a systemic impact on the teaching activity shaping the
way it is carried out on the students’ foreground.

312

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

DECISION MAKING USING ACTIVITY THEORY: AN EXAMPLE


FROM A STUDY

Cultural historical activity theory offers a lens that tries to capture the complexity
of teaching, by integrating dialectically the individual and the social/collective. The
activity is driven by a motive and directed towards an object (Leont’ev, 1978). Using
activity theory in teaching mathematics, we consider teachers (the subject) to be
involved in teaching activity with the motives of students’ learning of mathematics
and the fulfilment of other professional obligations. The unit of analysis is the activity
system (Engeström, 2001a), which incorporates social factors (rules, communities,
division of labour) that frame the relations between the subject and the object with
the mediation of tools (Figure 11.1). In our case, a tool with considerable influence
in the teaching activity is a new mathematics curriculum.

Figure 11.1. The activity system (adapted from Engeström, 2001a)

Every activity system is characterized by contradictions which are the driving


forces for the development of every dynamic system (Ilyenkov, 2009). Contradictions
may create learning opportunities for the subject and may broaden the activity, for
example, leading to reconsideration of the actions and goals (Engeström, 2001a;
Potari, 2013). In a recent study, the introduction and enactment of a new set of
mathematics curriculum materials produced or revealed contradictions in teaching
that emerged in teacher group discussions (Stouraitis, Potari, & Skott, 2017). In this
study, we introduced the construct of dialectical opposition to capture and interpret
the dialectical and epistemological nature of the contradictions related to teaching
and learning mathematics. Dealing with contradictions involves decisions about
the goals and the actions to be undertaken. Of particular importance are decisions
related to the “discrete individual violations and innovations” (Cole & Engeström,
1993), that is, the search for novel solutions as the first, individual response to the
emerged contradictions. Thus, although teachers’ decisions are part of the teaching
activity, they may have a transformational effect on this activity.
In our study, we use the dimensions of decision making of Engeström (2001b) that
we have described earlier in this chapter to capture the nature of the decisions that

313

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

teachers make and the underlying reasons. The systemic dimension is particularly
concerned with the way “this [decision] shapes the future of our activity” (Engeström,
2001b, p. 281). This dimension is connected to expansive learning, a theoretical
notion of the activity in which the learners are creating new ways to carry out the
activity, reconceptualising its object. This potentiality relates to the way subjects
think about the object of the activity and decide about the deliberate direction of
its transformation. Engeström, Engeström, and Kerosuo (2003) make a distinction
between action-based decisions about the actions to be undertaken, and future-
oriented envisioning, which is the imagination of the deliberate situation of the
object as outcome of the activity. Drawing from their interventional study in health
sector, they argue that intertwining these two aspects is necessary in any attempt
to transform the activity stating that “history is made in future oriented situated
actions” (p. 287).

The Research Context and Methodological Descriptions

Attempting to exemplify an activity theory perspective on decision making, we draw


from our research data. The second author collaborated with teachers in two lower
secondary schools that were piloting the materials of a newly prescribed curriculum
in their classrooms. The new materials emphasize students’ mathematical reasoning
and argumentation, connections within and outside mathematics, communication
through the use of tools, and students’ metacognitive awareness. It also attributes a
central role to the teacher in designing instruction. The collaboration took place in 8
and 10 group meetings at the respective schools, where the teachers discussed their
lesson planning and reflected on their experiences from teaching some modules of
the designed curriculum. The researcher participated in these meetings, providing
explanations about the rationale of the materials, as he was a member of the team
that developed the curriculum. He also discussed with the teachers their reflections
to provoke their explanations about the rationale of their choices. In this chapter, we
refer to two reflection groups, one of five teachers working in school A and one of
two teachers working in school B.
School A is a public experimental school with an innovative spirit. Our focus here
is on the teaching decisions of two of the teachers, Marina and Linda. They both
have more than 25 years of teaching experience and additional qualifications beyond
their teacher certification, as Marina has a masters’ degree in mathematics and Linda
has one in mathematics education. They both have experiences with innovative
teaching approaches and they have participated in teacher collaborative groups that
develop classroom materials. Further, Marina has written papers for conferences
and for journals for mathematics teachers, maintains links to communities dealing
with mathematics and is more informed than Linda about the recent activities of
the mathematics education community in Greece. Linda has also been involved in
producing materials and offering professional development courses for mathematics
teachers. Both have strong views about their instructional choices and a critical

314

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

stance on teaching innovations and materials introduced by various agents. Focusing


on Marina and Linda, we try to show how decisions could be made by experienced
teachers with innovative perspectives in a supportive school context.
School B is a normal public school with a culture open to innovations. Peter,
one of the teachers in focus here, has been teaching in public schools for about
15 years. Before this, he was teaching in private education preparing students for
examinations. Peter is an assistant principal of the school, attended a master course
in education in the period that the study was going on, and has been a teacher
educator preparing teachers to use digital tools in teaching mathematics. He is open
to the new curriculum, but he bases his teaching on the old textbooks. In the year of
the study, Peter was questioning the teaching practices he was involved in for many
years. Thus, Peter’s decision making is worth focusing on as a more usual case.
The research data consist of transcriptions of the audiotaped group discussions,
written documents (worksheets) and interviews conducted with each teacher in the
beginning of the study and six months after the end of the group meetings. The
transcriptions were analysed with methods inspired by grounded theory (Charmaz,
2006). The initial open coding resulted in the identification of the discussion
themes for each meeting and the corresponding contradictions that emerged in the
context of enacting the curriculum. Seeking an understanding of these emerging
contradictions, we used activity theory as a lens to study them and a language to
discuss their dialectical nature; integrating social, cultural and historical aspects
(Stouraitis, Potari, & Skott, 2017). Analysing the ways teachers decided to deal
with contradictions, we traced shifts in teachers’ discourse across different meetings
and interviews and we used activity theory and the relevant literature to interpret
these decisions and the factors framing them. Here we offer a glance to the teachers’
decisions and the potential shifts in teaching activity. We focus on two aspects: firstly,
on the dimensions of decision making as an interpretative framework and secondly,
on the future oriented envisioning as an indicator for the potential of transforming
the teaching activity. The first aspect has been initially discussed in Stouraitis (2016)
and the second in Stouraitis (2017).

Examples Teachers’ Decisions and the Potential Shifts in Teaching Activity

Below we analyse two examples as illustrative cases. With the first we exemplify
our use of Engeström’s (2001b) four dimensions to interpret decision making and
the concept of future oriented envisioning. The second one is used as a contrasting
case to discuss the potentiality for shifts in teaching activity based on future oriented
envisioning and its deficit.

First example: Teaching congruence involving geometrical transformations.


Geometrical transformations are introduced as a distinct topic in the new curriculum
with the rationale of supporting students’ development of spatial sense and of
using transformations when tackling issues of congruence and similarity. The use

315

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

of transformations as a proving tool is an alternative to the Euclidean perspective


in school geometry: the intuitive use of the moving figure is seen as incompatible
with the rigorous deductive rationale of Euclidean geometry. We consider the
contradiction between the two proving tools as a manifestation of the dialectical
opposition between intuition and logic.
In the fourth meeting (A4), Marina refered to her introductory lesson on triangle
congruence in grade 9 and to her students’ response that two triangles are congruent
if they “match after translation or reflection or rotation.” She considered using tasks
with geometrical transformations in parallel to or in combination with criteria of
triangle congruence. She described her goal saying, “I want them [the students]
to understand that when we compare angles or segments, we have two tools. One
is transformations and the other is the criteria of triangle congruence.” However,
she had not decided, and was wondering how she could do this, explaining:
“there is a need of investigation and inquiry before doing so.” Linda appreciated
Marina’s approach as a “nice idea,” nevertheless she prefered not to intertwine
the two topics. She pointed out that “the purpose [is for students] to learn how to
write [a justification], to observe the shape, to distinguish the given data from the
required claims, to make conclusions, and to prove,” implying that these goals can
be achieved through teaching congruence with a Euclidean perspective, without
involving transformations. Pursuing such goals in teaching Euclidean geometry,
Linda enriched and adapted the textbook tasks. Although Marina’s response was that
the same goals are relevant in every geometrical topic, Linda stated that in teaching
congruence she wanted to focus on Euclidean geometry and not on transformations.
In the next meeting (A5) Marina, having made the decision to combine the two
approaches, described how her students in grade 9 worked with the congruence of
triangles in combination with geometrical transformations to prove the congruence
of segments or angles. She noticed that this happened regularly in the class she
taught the previous year, but not very often in the one she was teaching now. In
this meeting, epistemological issues concerning the rigor and the intuition inherent
in different approaches were also discussed. Linda followed the discussion,
appreciating Marina’s approach as a “nice idea” and stating that she liked children
working in both ways (triangle congruence and geometrical transformations).
In the sixth meeting (A6) Marina had completed the topic of congruence.
Reflecting on her use of transformations in the classroom and on students’ work,
she explained her decision as creating an “opportunity to change the framework [of
proving] in grade 9” and to “get away from Euclidean geometry.”
In the 8th meeting (A8) Marina mentioned a seminar on transformations
she attended three years earlier at the university and her experimental teaching
of transformations in a school in which she was previously working. She also
mentioned that some students used transformations in other topics, such as
trigonometry, indicating that they used them as an operational tool to visualize and
prove congruence. In this discussion, Linda insisted on her decision not to intertwine

316

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

the different topics saying: “I like transformations per se. I don’t like overusing them
later in congruence … I don’t find the reason to [do so].”
Analysing Marina’s and Linda’s decisions across Engeström’s (2001b) four
dimensions, we make some interpretations. The social-spatial dimension is found
in the communities influencing the decisions. The group discussions in the meetings
appear to be supportive to Marina’s gradual formulation of goals and means, while
students’ predisposition to use geometrical transformations in congruence functioned
as trigger for her decision. Linda did not have such experiences with her students
and did not adopt Marina’s goals and decisions despite her involvement in the group
discussions. For both teachers, participating in communities before the year of the
study seems to influence their decisions. Especially for Marina, her comprehensive
experiences with mathematics and with teaching and her engagement in a learning
community specifically committed to discuss geometrical transformations may have
been important.
The anticipatory-temporal dimension can be found in the temporally distributed
steps of decisions. Marina’s decision to intertwine geometrical transformations
with Euclidean geometry in grade 9, came after her decision to teach systematically
transformations in grade 8. It is also a precursor for using transformations in other
topics such as trigonometry in grade 9.
The moral-ideological dimension is grounded on issues of power and teacher’s
responsibility about students’ well-being. Students’ positive reactions to Marina’s
attempts to consider transformations as a proving tool were crucial to her decisions.
Students’ involvement, understanding and positive dispositions are the grounds for
both teacher’s decisions, although in different ways.
The systemic-developmental dimension is found in the possibilities for action-
based decisions to shape the future of the broader activity. If adopted by the collective
subject (the community of mathematics teachers), Marina’s decisions can influence
the teaching activity. Using geometrical transformations as an alternative proving
tool alongside Euclidean geometry is a decision that can broaden the horizon of the
teaching activity, at least in Greek educational context.
Linda and Marina shared similar experiences and perspectives with the new set
of rules and tools in the form of the new curriculum materials. For them, significant
communities included the school they both worked at, and the same reflection group
that discussed approaches to teaching according to the new curriculum materials.
Both adopted a similar – but not identical – view for students’ learning as the object of
the activity: they prioritized understanding, mathematical reasoning and connections
with reality and within mathematics. Yet, there were significant differences between
the goals they were setting, the decisions they made and, consequently, the actions
they undertook. This is less striking if one considers these two teachers as having
“different positions and histories and thus different angles or perspectives on their
shared general object” (Engeström, 2001b, p. 286). Marina appears more familiar
with the mathematics of geometrical transformations to use them as a proving tool

317

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

alternative to Euclidean geometry, while Linda bases her goals on the affordances
of the Euclidean perspective. The apparent differences may possibly and in part
be explained by the different communities they have participated in and the tools
mediating the respective activities.
Focusing on the possibilities that their decisions have in shaping the future of
the teaching activity, we looked at the way the teachers envisioned the future of
their students’ learning. All of the aforementioned excerpts of Marina’s discourse
reveal a future-oriented envisioning of the object of the activity she was engaged in:
she imagined her students working fluently with both approaches and consciously
regarding the differences between them and she noted their development in this
direction. In the interview conducted in the following year, Marina said that she
uses the same approach with more elaborated tasks for her students. Like Marina’s
envisioning, Linda showed her motive in the relevant excerpts: she imagined her
students in the future to work on developing their understandings and proving
abilities based on the Euclidean perspective in school geometry.

Second example: The use of modelling in teaching algebra. The new curriculum
materials recommend mathematical modelling as an important aspect in students’
meaning making in algebra. Generating algebraic expressions and equations to
represent realistic situations and problems is introduced in 7th, 8th and 9th grades.
In group discussions about teaching polynomials in grade 9, a common contradiction
was about introducing polynomials and operations in a formal, abstract way
or involving realistic situations and modelling procedures. This contradiction
is a manifestation of the dialectical opposition between the abstract and the
concrete.
In the third meeting (B3), Peter described his introductory lesson of monomials
using only definitions, examples and counterexamples. He explained, “we begin
with the algebraic expression, they [the students] read the definition, and I give
them examples to discuss … then to monomials [with the same way].” After the
researcher and Manolis (a Peter’s colleague) questioned him about the “why” of
teaching polynomials, Peter refered to a similar student’s question. He reflected that
“he begins with the definitions” but considered that he “must pay more attention …
to the practical use of monomials.” Again, in the discussion with Manolis and the
researcher about modelling, Peter started thinking the potentials of it. After some
turns talking, he said that he liked the word “modelling” because “it shows exactly
what we are doing: we transform real situations to mathematics, verbal expressions
to mathematical ones.” With modelling “you give [students] a motive, a goal. Ok,
you must first pose the problem to create questions.”
Although Peter found modelling to be a useful idea, he was engaging in the
discourse by emphasizing the role of mathematics and his own teaching but not
the deliberate students’ development. For example, he described what “he did” and
what he “usually does,” and that modelling is what “we do in mathematics.” In
this discourse, no explicit or implicit longitudinal objective appears to be related

318

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

to the way his students should deal with modelling. This can be interpreted as
absence of any clear articulation of his future-oriented envisioning that could lead
his decisions.
In another meeting (B7), Peter referred to classroom discussions about functions
where the students and the teacher modelled realistic situations and phenomena
(mostly from physics) leading to linear and quadratic functions. He explained
that his goal was for “[the students] to understand that a function shows a relation
between two interdependent things, and that everything is a potential function.”
These formulations reveal Peter’s future-oriented envisioning about students’
understanding of functions and to connecting them to realistic situations and also
to physics. But again, there is not any similar envisioning about students’ work
on modelling per se. The modelling processes Peter incorporated in his classroom
discussions were limited at the level of actions subordinated to his teaching of
functions.
Analysing Peter’s decisions across Engeström’s (2001b) four dimensions, we see
the social-spatial dimension in the communities influencing the decisions. Peter’s
adoption of the idea of modelling in teaching polynomials and functions can be
interpreted as adoption of elements introduced by the new curriculum, based on
Peter’s reflection on and questioning of his teaching. The group discussions with
Manolis and the researcher seem to trigger Peter’s reflection and shift. Nevertheless,
Peter’s previous involvement in practices, like preparing students for examinations
in the private education, seem to have a strong influence on his decisions. This is
obvious in Peter’s choice to use more formal ways to introduce algebraic expressions
(monomials and polynomials).
The anticipatory-temporal dimension can be found in the connections with
modelling in previous grades (7th and 8th) where Manolis was teaching. Peter’s
decisions on teaching polynomials is also connected to his teaching of functions and
equations.
The moral-ideological dimension is grounded on the role the students’ reactions
have on Peter’s decisions. Giving students “a motive, a goal” and pursuing their
understanding “that everything is a potential function” are Peter’s moral commitment
on his student’s interests. Nevertheless, his own teaching practice and mathematical
content appear to prevail in his rationale.
The systemic-developmental dimension is found in the potentiality of Peter’s
action-based decisions to shift his teachings. Actually, these decisions did not give
rise to actions involving students in modelling procedures, especially in teaching
polynomials, although in teaching functions more clear connections were made
to realistic situations. Moreover, his decision about modelling did not have any
systemic influence on the teaching activity, since it was not connected with future
oriented envisioning.
In the interview conducted in the following year, the researcher asked Peter if he
used modelling in teaching polynomials this year. Peter responded that although he
thinks it is useful and keeps it in mind, he “hasn’t the time to do all this,” showing

319

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

that there is not any movement in the way he carries out the teaching activity about
modelling.

Discussion and Conclusion

The introduction of the new curriculum created or revealed contradictions that


provide opportunities for teachers to engage differently in mathematics teaching
and learning. With the examples provided above, we presented some of these
opportunities and teachers’ decisions to make or not make shifts in their teaching and
we attempted to offer brief interpretations based on activity theory. In both examples,
all three teachers seem to be aware of a contradiction related with the introduction
of the new curriculum. Marina and Linda appear to be more aware of it in relation
to its epistemological and dialectical nature. Peter also shows an understanding of
some aspects of the relevant contradiction. These examples suggest that teachers’
awareness of the contradiction is a necessary but insufficient driving force for the
development of the teaching activity. From this point, teachers’ decisions can lead to
one or the other direction.
In the two provided examples, three possibilities appear for teachers’ decisions and
the way these decisions may or may not influence the future of the activity. Marina’s
decision to combine geometrical transformations with Euclidean geometry is an
attempt to overcome the contradiction synthesizing dialectically the opposing poles.
On the other hand, Linda decides to keep the two opposing poles separated, pursuing
the affordances of Euclidean geometry. Somehow in the middle, Peter decides to
deal with the contradictions using aspects of modelling in teaching functions, but not
to use modelling as a meaning-making introductory activity in polynomials.
Although the activity is collective and the object is socially formulated, different
teachers can have “different positions and histories and thus different angles or
perspectives on their shared general object” (Engeström, 2001b, p. 286). In the
first example provided, Marina and Linda make different decisions about the same
contradiction. The difference may in part be explained by their different histories,
including Marina’s attending the seminar and her experimental teachings. In the
second example, Peter’s decision seems to be influenced by his previous activities
in the private education sector.
Marina’s decision has the potential to transform the teaching activity, broadening
the horizon of the possible modes in which this activity is carried out. The dialectical
overcoming of the contradiction is a discrete individual innovation, although its
evolvement is not already known. Linda’s approach does not transform the activity
but clarifies and strengthens some objectives of teaching Euclidean geometry.
Linda’s decision reinforces aspects of the existing way the teaching activity is carried
out, showing that learning is not necessarily expansive (Engeström, Engeström, &
Kerosuo, 2003). Peter’s decisions do not have any shifting effect to the way the
activity is carried out, nor reinforce any existing practice. Somehow this decision
seems to not have the power to affect the activity.

320

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

What is the difference between Marina’s and Linda’s decisions on the one hand,
and Peter’s decision on the other, that provide the different power on them? The
difference could not be the attempt to overcome or not the contradictions, since
Marina’s and Linda’s decisions differ at this point, although both are strong enough to
have an effect on the teaching activity. The difference is grounded in the connections
made between action-based decisions and future-oriented envisioning of the object.
Marina and Linda underpin their decisions about the actions they undertake with a
strong future-oriented projection of their students’ understanding. This adds fluency
in deciding among the possible actions realizing the relevant goals. At the same
time, it generates decisions with the potential to be stabilized, even if initially the
stabilization refers only to individual modes of carrying out the activity. On the other
hand, Peter’s decisions seem to be restricted to action level, without a grounding
on future envisioning of the object, namely the deliberate modelling processes
his students should be able to be involved in as outcome of the sequential actions
undertaken. The absence of future-orientation restricts the horizon of possible actions
and reduces the potentiality of stabilizing them. Our conclusions appear in line with
Engeström et al. (2003) who, researching developmental work in the health sector,
write that “professionals make history in future-oriented discursive actions” (p. 286).
Summing up, one can argue that for decisions to affect the activity some aspects
seem to be necessary: the emergence of a contradiction and some degree of
awareness about it, a willingness to deal with it and a future-oriented envisioning
about the outcomes of the activity. If there is to be a transformation of the activity,
the decision must aim at a dialectical overcoming of the contradiction by searching
for new solutions. Although schematic and perhaps simplistic, this sequence may
represent some crucial aspects of decision making, especially the relations between
action-based decisions and the future of the activity.
“Traditional views locate decision making in the heads of individuals at a given
point of time in a particular place” (Engeström, 2001b, p. 282) and thus, the social,
historical and systemic character of decision making are out of search. Searching for
what makes teachers form goals and what creates the horizon for possible actions
under an activity theoretical view contributes to our understanding of teachers’
decisions and mathematics teaching activity.

EMERGING ISSUES AND FUTURE DIRECTIONS

Research on mathematics teacher decision making has a long history and has been
always linked to the practice of mathematics teaching in different contexts. Teacher
decision making is studied in hypothetical teaching situations, actual teaching,
planning teaching using curricular resources, and in teacher collaborative work
with other teachers and researchers. It has been mostly studied as an element of
noticing mathematics teaching and in some studies is related to mathematics teacher
knowledge. It is one element of the triplet of perceiving, interpreting and decision
making in the context of noticing (Stanhke et al., 2016).

321

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

During the last decade, research on teacher decision making has started to move
from an emphasis on what decisions the teacher makes in the moment or in the
planning of teaching to why these decisions are made or why the teacher does not
usually make the same decisions even in similar situations. This shift of attention
has made more apparent that decision making is a very complex process that cannot
be described and interpreted if we focus only on the individual teacher and his or
her orientations. Teacher decision making depends on the classroom interaction and
on what the students bring into the discussion, on teacher’s participation in different
practices such as collaborative contexts of working with other teachers in her school
or in professional development context, figured worlds like curriculum innovations
(as Skott, 2015 argues in his pattern of participation theory).
The two examples that we share from our own research and the activity theory
analysis add to the discussion about this dynamic character of teacher decision
making. They trace aspects of the path leading from the contradictions that the
teachers faced and the processes of handling them to the transformation of the
teaching activity. The examples offer two particularly important aspects. Firstly, the
four dimensions of decision making (social-spatial, anticipatory-temporal, moral-
ideological and systemic-developmental) capture social and historical aspects
of teachers’ decisions. Secondly, the distinction between action-based decisions
and future oriented envisioning, provides a lens to interpret the possible power of
teachers’ decisions. At a theoretical and methodological level, our work considers
decision making in relation to teaching actions, goals and how these are interrelated
to the activity of mathematics teaching that is framed from the tools and resources
that a teacher has available and from the rules that the communities in which she or
he participates are brought into the way that her or she attempts to achieve her or
his goals.
Another direction of current research on teacher decision making is the study of
the impact that teachers’ decisions can have on the effectiveness of mathematics
teaching practices to promote students’ mathematical learning. Although there
are some studies that focus on the relationship between teaching and its impact
on learning, still the methodological tools used are emphasizing a cause-effect
relationship (Jentsch & Schlesinger, 2017). Links between teacher decision making
and students’ potential and actual learning pose new methodological challenges
in building these relations beyond a “cause-effect” perspective. Tracing teachers’
decisions on specific classroom events, their rationale and their impact on the
evolution of mathematics teaching and learning will help us to understand better
mathematics teaching and learning. Stockero and Van Zoest (2013) have initiated
such a direction.
Teacher decision making has been linked to the development of teacher
consciousness. This emphasis is both in studies with a focus on the individual and
the social aspects. Providing opportunities for reflecting on teacher decisions in
professional development and in initial mathematics teacher education is important
for the development of mathematics teaching and learning. Most studies on decision

322

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

making recommend the use of frameworks on teacher decision making that they
have developed in mathematics teacher education. However, very few examples
have been reported in research in relation to how these frameworks have been used
and what was their impact of teacher professional learning. Research in this area is
needed if we would like the teachers to become aware of their teaching decisions,
their dynamic character and their impact on students’ learning.

REFERENCES
Amador, J. M., & Carter, I. S. (2018). Audible conversational affordances and constraints of verbalizing
professional noticing during prospective teacher lesson study. Journal of Mathematics Teacher
Education, 21(1), 5–34.
Ball, D., Thames-Hoover, M., & Phelps, G. (2008). Content knowledge for teaching: What makes it
special? Journal of Teacher Education, 59(5), 389–407.
Bishop, A. J. (1976). Decision-making, the intervening variable. Educational Studies in Mathematics,
7(1-2), 41–47.
Bishop, A. J., & Whitfield, R. C. (1972). Situations in teaching. Berkshire: McGraw-Hill.
Borko, H., Roberts, S. A., & Shavelson, R. (2008). Teachers’ decision making: From Alan J. Bishop to
today. In P. Clarkson & N. Presmeg (Eds.), Critical issues in mathematics education (pp. 37–67).
New York, NY: Springer.
Brown, M. W. (2009). The teacher–tool relationship: Theorizing the design and use of curriculum
materials. In J. T. Remillard, B. A. Herbel-Heisenmann, & G. M. Lloyd (Ed.), Mathematics teachers
at work: Connecting curriculum materials and classroom instruction (pp. 19–36). New York, NY &
London, UK: Routledge.
Charmaz, K. (2006). Constructing grounded theory: A practical guide through qualitative analysis.
London: Sage publications.
Choppin, J. M. (2011). Learned adaptations: Teachers’ understanding and use of curriculum resources.
Journal of Mathematics Teacher Education, 14(5), 331–353.
Clarkson, P., & Presmeg, N. (2008). Critical issues in mathematics education. New York, NY: Springer.
Cole, M., & Engeström, Y. (1993). A cultural historical approach to distributed cognition. In G. Salomon
(Ed.), Distributed cognitions: Psychological and educational considerations (pp. 1–46). Cambridge,
UK: Cambridge University Press.
Cooney, T. (1988). Teachers’ decision making. In D. Pimm (Ed.), Mathematics teachers and children.
London, UK: Hodder & Stoughton.
Earbest, D., & Amador, J. M. (2019). Lesson planimation: Prospective elementary teachers’ interaction
with mathematics curricula. Journal of Mathematics Teacher Education, 22, 37–68.
Eisenmann, T., & Even, R. (2009). Similarities and differences in the types of algebraic activities in two
classes taught by the same teacher. In J. T. Remillard, B. A. Herbel-Heisenmann, & G. M. Lloyd
(Eds.), Mathematics teachers at work: Connecting curriculum materials and classroom instruction
(pp. 152–170). New York, NY & London: Routledge.
Engeström, Y. (2001a). Expansive learning at work: Toward an activity theoretical reconceptualization.
Journal of Education and Work, 14(1), 133–156.
Engeström, Y. (2001b). Making expansive decisions: An activity-theoretical study of practitioners
building collaborative medical care for children. In C. M. Allwood & M. Selart (Eds.), Decision
making: Social and creative dimensions (pp. 281–301). Dordrecht: Springer Science+Business Media.
Engeström, Y., Engeström, R., & Kerosuo, H. (2003). The discursive construction of collaborative care.
Applied Linguistics, 24(3), 286–315.
Evans, J. S. B., & Stanovich, K. E. (2013). Dual-process theories of higher cognition: Advancing the
debate. Perspectives on Psychological Science, 8(3), 223–241.
Gueudet, G., & Trouche, L. (2009). Towards new documentation systems for mathematics teachers?
Educational Studies in Mathematics, 71, 199–218.

323

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
DESPINA POTARI & KONSTANTINOS STOURAITIS

Ilyenkov, E. V. (2009). The ideal in human activity. Pacifica: Marxists Internet Archive.
Jacobs, V. R., Lamb, L. L., & Philipp, R. A. (2010). Professional noticing of children’s mathematical
thinking. Journal for Research in Mathematics Education, 41(2), 169–202.
Jentsch, A., & Schlesinger, L. (2017). Measuring instructional quality in mathematics education. In
T. Dooley & G. Gueudet (Eds.), Proceedings of the 10th Conference of the European Society for
Research in Mathematics Education (CERME10) (pp. 3073–3080). Dublin, Ireland: DCU Institute of
Education and ERME.
Kaiser, G., Busse, A., Hoth, J., König, J., & Blömeke, S. (2015). About the complexities of video-based
assessments: Theoretical and methodological approaches to overcoming shortcomings of research on
teachers’ competence. International Journal of Science and Mathematics Education, 13(2), 369–387.
Lande, E., & Mesa, V. (2016). Instructional decision making and agency of community college
mathematics faculty. ZDM Mathematics Education, 48(1–2), 199–212.
Leont’ev, A. N. (1978). Activity, consciousness and personality. Englewood Cliffs, NJ: Prentice-Hall.
Lloyd, G. M. (2008). Teaching mathematics with a new curriculum: Changes to classroom organization
and interactions. Mathematical Thinking and Learning, 10, 163–195.
Mason, J. (2002). Researching your own practice: The discipline of noticing. London: Routledge.
Paterson, J., Thomas, M., & Taylor, S. (2011). Decisions, decisions, decisions: What determines the path
taken in lectures? International Journal of Mathematical Education in Science and Technology, 42(7),
985–995.
Pepin, B., Gueudet, G., & Trouche, L. (2013). Re-sourcing teachers’ work and interactions: a collective
perspective on resources, their use and transformation. ZDM Mathematics Education, 45, 929–943.
Potari, D. (2013). The relationship of theory and practice in mathematics teacher professional development:
An activity theory perspective. ZDM Mathematics Education, 45(4), 507–519.
Rowland, T., Huckstep, P., & Thwaites, A. (2005). Elementary teachers’ mathematics subject knowledge:
The Knowledge Quartet and the case of Naomi. Journal of Mathematics Teacher Education, 8,
255–281.
Rowland, T., & Turner, F. (2006). A framework for the observation and review of mathematics teaching.
Mathematics Education Review, 18, 3–17.
Santagata, R., & Yeh, C. (2016). The role of perception, interpretation, and decision making in the
development of beginning teachers’ competence. ZDM Mathematics Education, 48(1–2), 153–165.
Schoenfeld, A. H. (2011). How we think: A theory of goal-oriented decision making and its educational
applications. New York, NY: Routledge.
Shavelson, R. J. (1973). What is the basic teaching skill? Journal of Teacher Education, 24(2), 144–151.
Sherin, M. G., & Van Es, E. A. (2005). Using video to support teachers’ ability to interpret classroom
interactions. Journal of Technology and Teacher Education, 13(3), 475–491.
Shulman, L. S., & Elstein, A. S. (1975). Studies of problem solving, judgment, and decision making:
Implications for educational research. In F. N. Kerlinger (Ed.), Review of research in education
(Vol. 3). Itasca, IL: F.E. Peacock.
Skott, J. (2013). Understanding the role of the teacher in emerging classroom practices: Searching for
patterns of participation. ZDM Mathematics Education, 45(4), 547–559.
Skott, J. (2015). Towards a participatory approach to ‘beliefs’ in mathematics education. In B. Pepin &
B. Roesken-Winter (Eds.), From beliefs to dynamic affect systems in mathematics education
(pp. 3–23). Cham: Springer International Publishing. doi:10.1007/978-3-319-06808-4_1
Smith, M. S., & Stein, M. K. (2011). Five practices for orchestrating productive mathematics discussions,
Reston, VA: National Council of Teachers of Mathematics.
Son, J.-W., & Kim, O.-K. (2015). Teachers’ selection and enactment of mathematical problems from
textbooks. Mathematics Education Research Journal, 27(4), 491–518.
Stahnke, R., Schueler, S., & Roesken-Winter, B. (2016). Teachers’ perception, interpretation, and decision-
making: A systematic review of empirical mathematics education research. ZDM Mathematics
Education, 48(1), 1–27.
Stockero, S. L., Leatham, K. R., Ochieng, M. A., Van Zoest, L. R., & Peterson, E. B. (2019). Teachers’
orientations toward using student mathematical thinking as a resource during whole-class discussion.
Journal of Mathematics Teacher Education. (Online first) https://doi.org/10.1007/s10857-018-09421-0

324

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
TEACHER DECISION MAKING

Stockero, S. L., & Van Zoest, L. R. (2013). Characterizing pivotal teaching moments in beginning
mathematics teachers’ practice. Journal of Mathematics Teacher Education, 16(2), 125–147.
Stouraitis, K. (2016). Decision making in the context of enacting a new curriculum: An activity-theoretical
perspective. In C. Csíkos, A. Rausch, & J. Szitányi (Eds.), Proceedings of the 40th Conference of the
International Group for the psychology of mathematics education (Vol. IV, pp. 235–242). Szeged,
Hungary: PME.
Stouraitis, K. (2017). Teachers’ decisions and the transformation of teaching activity. In T. Dooley &
G. Gueudet (Eds.), Proceedings of the 10th Conference of the European Society for Research in
Mathematics Education (CERME10) (pp. 3177–3184). Dublin, Ireland: DCU Institute of Education
and ERME.
Stouraitis, K., Potari, D., & Skott, J. (2017). Contradictions, dialectical oppositions and shifts in teaching
mathematics. Educational Studies in Mathematics, 95(2), 203–217. doi:10.1007/s10649-017-9749-4
Thomas, M., & Yoon, C. (2014a). The impact of conflicting goals on mathematical teaching decisions.
Journal of Mathematics Teacher Education, 17, 227–243.
Thomas, M., & Yoon, C. (2014b). The role of teaching decisions in curriculum alignment. In C. Nicol,
S. Oesterle, P. Liljedahl, & D. Allan (Ed.), Proceedings of the 38th Conference of the International
Group for the Psychology of Mathematics Education and the 36th Conference of the North American
Chapter of the Psychology of Mathematics Education (Vol. 5, pp. 241–248). Vancouver, Canada:
PME.
van Es, E. A., & Sherin, M. G. (2002). Learning to notice: Scaffolding new teachers’ interpretations of
classroom interactions. Journal of Technology and Teacher Education, 10(4), 571–596.
Van Zoest, L. R., Peterson, B. E., Leatham, K. R., & Stockero, S. L. (2016). Conceptualizing the teaching
practice of building on student mathematical thinking. In M. B. Wood, E. E. Turner, M. Civil, & J. A.
Eli (Eds.), Proceedings of the 38th annual meeting of the North American Chapter of the International
Group for the psychology of mathematics education (pp. 1281–1288). Tucson, AZ: University of
Arizona.
Watson, S. (2019). Revisiting teacher decision making in the mathematics classroom: A multidisciplinary
approach. Paper presented in CERME 11, Utrecht, The Netherlands.

Despina Potari
Department of Mathematics
National and Kapodistrian University of Athens

Konstantinos Stouraitis
Institute of Educational Policy

325

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
PART 5
FROM MATHEMATICS TEACHING PRACTICES TO
TEACHER EDUCATION

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER AND CRAIG POURNARA

12. EXEMPLIFYING WITH VARIATION AND ITS


DEVELOPMENT IN MATHEMATICS TEACHER
EDUCATION

This chapter focuses on mathematics teachers’ knowledgeable use of examples, what


we refer to as exemplification, with explicit attention to principles of variation within
and across example sets. We present a rationale for the co-ordination of principles
of variation and exemplification, and then illustrate what and how we apply this
in our professional development work with teachers. We distinguish between
modelling exemplification with variation in our mathematics teaching, and then
mediating this practice as we focus on the teaching of mathematics. Through this we
present opportunities in mathematics teacher education for teachers to learn about
exemplification and how the notion of variation provides a means for constructing,
critiquing and revising examples sets for use with learners.

INTRODUCTION

The role of examples in mathematics teaching and learning has gained considerable
attention in the last decade (e.g., Bills & Watson, 2008; Zaslavsky, 2017). Claims
have been made about the centrality of examples in mathematical learning, how
such learning is made possible by ‘careful and knowledgeable’ use of examples in
mathematics teaching (Watson & Chick, 2011), and in mathematics teacher education
(Peled & Balacheff, 2011). By pointing to the knowledgeable use of examples
in mathematics teaching as part of a teacher’s mathematical work (Ball, 2017),
this line of research contributes to the growing field of research on mathematical
knowledge for teaching. In Ball’s terms, the knowledge entailments of such work
are learnable, and consequently of significance for mathematics teacher education.
In this chapter, we pursue these ideas with a focus on the why, what and how of
working deliberately with secondary mathematics teachers on their use of examples
in their teaching, or what we refer to more simply as exemplification. The context
of our work is the Wits Maths Connect Secondary project where exemplification,
and more specifically example sets informed by principles of variation, is a core
component of our professional development work with teachers.
We begin the chapter with a brief introduction to the Wits Maths Connect Secondary
project, its theoretical orientation to teaching, including exemplification, and some
empirical research results. This provides the contextual basis for why we have come

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_013

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

to focus on exemplification. We then present a review of the literature on examples


and exemplification in mathematics education, the connection between this and
ideas about variation, and their combined significance for teachers’ knowledge and
practice. The review thus offers a more theoretically-informed rationale for the co-
ordination of exemplification and variation in our work with teachers. The remainder
of the chapter zooms into our teacher education practice, where we illustrate the
what and how of this work. We distinguish modelling of exemplification from
mediating exemplification with teachers. The former takes place in the mathematics-
focused aspects of our professional development work, while the latter is central
to the teaching-focused aspects. We describe opportunities in mathematics teacher
education for teachers to learn about exemplification and how the notion of variation
provides a means for constructing, critiquing and revising examples sets for use
with learners.1 We conclude with some reflection on the challenges of this focus in
mathematics teacher education for both teachers’ learning and mathematics teacher
educators’ teaching, pointing to the need for further research and development work.

WHY EXEMPLIFICATION? THE WITS MATHS CONNECT SECONDARY


PROJECT THEORY OF TEACHING AND INITIAL RESEARCH

The Wits Maths Connect Secondary project is a longitudinal research-linked


professional development project aimed at improving mathematics teaching in
socio-economically disadvantaged schools in one province in South Africa. Our
work is theoretically and empirically informed. It is shaped, on the one hand, by on-
the-ground realities of mathematics teaching and learning in such schools, and on
the other, by an orientation to the activity of teaching as ‘social’.
Theoretically, we draw from key tenets of sociocultural theory, where
mathematics is viewed as an inter-connected network of scientific concepts, and
mathematics teaching therefore as geared towards the mediation and appropriation
of the increasingly sophisticated and increasingly general ways of thinking that
constitute progression in the discipline (Vygotsky, 1978). Teaching as an activity is
not only goal-directed but also always about something (Alexander, 2000). Bringing
that something into learners’ focus – its mediation – is the teacher’s work. We call
this ‘something’ the object of learning2 – that which students are to know and be
able to do. In practical terms, it is akin to a lesson goal, but worded so that the
mathematics of the goal is made clear. Whatever the goal, it needs to be mediated. In
line with previous research, mediational means are understood as cultural tools and/
or resources in the practice of teaching (Adler, 2001).
In our initial observations of numerous mathematics lessons in our project
schools, and as anticipated, traditional forms of teaching predominated. This is not
surprising as traditional forms of teaching are common across the world (Nachlieli &
Tabach, 2018). Typically, a topic was announced by the teacher who proceeded with
whole-class teaching to exemplify and explain some content, be it a mathematical
idea or a procedure. This was followed by classwork and homework, typically on an

330

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

example set from a textbook or from other requisite materials. Learner participation
in lessons was typically limited to individual or chorused responses to the teachers’
questions and copying of worked examples from the board. In the South African
context, there are many factors that shape such learner participation in traditional
teaching forms, including overcrowded classrooms and limited resources, as well as
a relatively prescriptive curriculum with emphasis on coverage.
More careful analysis of the video-records of these mathematics lessons
showed that they were characteristically incoherent. While students might have
been ‘working’ (listening to the teacher, writing in their notebooks), and teachers
were following high levels of curriculum prescription, we struggled to interpret
the intended mathematical message in a lesson. We puzzled over how specific
mathematical goals prefigured lesson development activity for teachers. We described
this phenomenon as lessons where the mathematical object of learning was out of
focus, despite an announced topic or content. There was no apparent mathematical
‘story’ linking what learners were to know and be able to do. For example, in a four-
part lesson ostensibly on multiplying algebraic expressions, different rules reliant
on visible forms of the expressions were emphasized in each part of the lesson and
across a range of examples of such expressions, making available a fragmented and
incoherent notion of the products (Adler & Venkat, 2014).
Given the context described above, and a principle that good professional
development begins with what teachers bring and so who and where they are, it
made sense that our professional development work should attend to strengthening
teachers’ exemplification and the quality of their explanations. We were emboldened
in our decision by concurrent literature on teacher education advocating deliberate
attention to teaching practices that could lever up quality teaching – practices they
referred to as ‘high leverage’ (see for example Ball, Sleep, Boerst, & Bass, 2009;
Grossman, Hammerness, & McDonald, 2009). Criteria for such practices include
that they are (1) central to mathematics and effective for student learning; (2)
used frequently in mathematics teaching and apply across different approaches to
mathematics teaching; and most critically (3) possible to describe and illustrate thus
making them teachable and learnable. Exemplification meets these criteria.
As a high leverage practice, it is also important, particularly from a Vygostkian
perspective, to recognise that examples are symbolic mediators of mathematics.
Symbolic mediators include different signs, symbols, writing, formulae and graphic
organisers – all possible elements of mathematical examples. As Kozulin (2003)
explains, one cannot take for granted that learners will detect symbolic relations, no
matter how obvious they might seem to the teacher.
Symbols may remain useless unless their meaning as cognitive tools is properly
mediated… the mere availability of signs or texts does not imply that they will
be used by students as psychological tools…. (Kozulin, 2003, p. 24)
The implications for teaching and learning follow. Appropriation of psychological
tools and more connected scientific mathematical concepts requires deliberate

331

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

teaching of symbolic tools. This includes their systematic organisation and an


emphasis on their generality and application.
Symbolic tools (e.g., letters, codes, mathematical signs) have no meaning
whatsoever outside the cultural convention that infuses them with meaning
and purpose. (Kozulin, op cit., p. 26)
Indeed, instructional3 examples, embedded as they are in a pedagogic practice
are just such cultural conventions. Examples and exemplification thus form part
of a framework we have developed over time that informs both our research and
development work. This framing, named Mathematical Discourse in Instruction,
enables us to describe and analyse what it is teachers do, and then to work with them
developmentally on the mathematical quality of their teaching.
Elsewhere we have described how Mathematical Discourse in Instruction
operates as a boundary object in the Wits Maths Connect Secondary project (Adler,
2017), moving flexibly across practices in the project. It has framed the planning
of our professional development sessions in our 16-day, year-long mathematics
for teaching course (Pournara, Hodgen, Adler, & Pillay, 2015); it was used as a
tool for planning and reflection in our complementary lesson study work (Adler &
Alshwaikh, 2019); and we have operationalised it in our research on mathematics
teaching (Adler & Ronda, 2015). Mathematical Discourse in Instruction is a living
framework, with power lying in its iterative nature, moving between and supporting
our research and development work and teachers’ teaching practices.
As represented in Figure 12.1, Mathematical Discourse in Instruction focuses on
four key elements of mathematics teaching, with a lesson as our unit of analysis:
the object of learning, or lesson goal, and three mediational means, or cultural tools,
exemplification, explanatory talk and learner participation. The object of learning
in any lesson could be a concept, a procedure or mathematical practice, together
with the relevant capability. This leads to exemplification and more specifically to
examples and associated tasks that can be used to bring the object of learning into
focus with learners. With respect to examples, we are interested in their selection
and sequencing and how these accumulate within and across lessons. We draw
on the work of Watson and Mason (2006), who in turn draw on Marton and Tsui
(2004) to describe key features of a mathematical object and/or movement towards
generality across a sequence of examples. An example set that brings attention to
similarity across examples, and so to that which is invariant, offers opportunity to
identify key features and/or to generalise. If a set of examples brings attention to
contrast, and so to what something is in relation to what it is not, or to a different
class, opportunities are made available to recognise boundaries between classes of
examples. This provides further opportunity to generalise but not overgeneralise.
When two examples that are carefully varied are juxtaposed, they can draw quite
specific attention to a key feature of an object.
We note the third and fourth elements of the Mathematical Discourse in
Instruction framework, though they are not in focus in this chapter. Explanatory

332

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

Figure 12.1. Constitutive elements of Mathematical Discourse in Instruction (adapted from Adler & Ronda, 2015)

333

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

communication includes attention to naming/word use (what is said and what is


written) and substantiations of mathematics as specialised knowledge (what counts
as mathematical knowledge). Learner participation focuses on what learners do and
say with regards to the mathematics they are learning. We consider whether and how
learner talk moves beyond the very prevalent but limited offering of single words in
chorus, to responding to and asking questions, and to more open dialogue with the
teacher and/or other learners.
The overarching importance of the framework is the emphasis on the coherence
of a lesson and thus how all elements interact, how they link back to the object
of learning and open opportunities to learn mathematics. The constitutive elements
of Mathematical Discourse in Instruction apply to any mathematics classroom,
whatever the pedagogy. As already noted, our elaboration of exemplification in the
project and more specifically in our work with teachers was simultaneously informed
by the growing research on examples in mathematics education.

WHY EXEMPLIFICATION? MATHEMATICS EDUCATION RESEARCH

The setting up of exemplification as a research field in mathematics education can be


traced back to the Research Forum on Exemplifying at the Psychology of Mathematics
Education 30th conference in Prague 2006 (Bills et al., 2006). The literature on
examples dates back further. For example, Mason and Pimm (1984) explored the
role of generic examples in mathematics education. Dahlberg and Housman (1997)
investigated the generation of examples in learning advanced mathematics. It is
fair to say that interest in examples in mathematics and mathematics education
follows from a basic maxim that initial experiences of mathematical concepts and
procedures, given their abstract nature, will be through some exemplification:
through examples and the tasks in which they are embedded. Goldenberg and Mason
(2008) described examples as cultural mediating tools, and linked examples with the
notion of variation:
Examples can … usefully be seen as cultural mediating tools between learners
and mathematical concepts, theorems, and techniques. They are a major means
for ‘making contact’ with abstract ideas and a major means of mathematical
communication, whether ‘with oneself’, or with others. Examples can also
provide context, while the variation in examples can help learners distinguish
essential from incidental features and, if well selected, the range over which
that variation is permitted. (Goldenberg & Mason, 2008, p. 184, emphasis
added)
The resonance is apparent between this description, our theoretical orientation to
examples as symbolic tools, and to our work on and with exemplification in the Wits
Maths Connect Secondary project as briefly described above.
The 2006 Psychology of Mathematics Education Research Forum culminated in
a special issue of Educational Studies in Mathematics in 2008 (Bills & Watson,

334

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

2008) and was a catalyst for a follow-on conference focused on the role of examples
in argumentation and proof, and in conceptualizing and defining. The culmination
of this latter work was published in a special issue of the Journal of Mathematical
Behavior in 2011 (Antonini, Presmeg, Mariotti, & Zaslavsky, 2011).
More recently, a special issue in the Journal of Mathematical Behavior
(Zaslavsky, 2017) provides an overview of research on the role of examples in
concept learning and proving, together with current work with this specific focus.
We note the introductory papers in each of these special issues as they point firstly to
the substantial growth of research on exemplification in mathematics education, and
secondly, they alert readers to existing research reviews. We do not rehearse these
reviews, but rather zoom in on the papers focused on the role of examples in teachers’
learning of mathematics in teacher education, or teachers’ use and awareness of
examples in their teaching. Through this the crucial role in teaching of choosing
and using (instructional) examples becomes evident. The value of a deliberate focus
on exemplification in mathematics teacher education follows. While some of these
papers move into exemplification in the context of teacher education, the focus has
largely been on teachers’ learning of mathematics and has not yet extended to the
mediation of exemplification itself as a mathematics teaching practice.
Bills and Watson (2008) aptly argued that acting with and on examples is
fundamental to any theory of mathematical learning, and thus to describing and
theorizing mathematics teaching.
Examples have always played a central role in both the development and
the teaching of mathematics, … whether they are seen as raw material for
generalisation, illustrations of techniques or concepts, elements of classes, or
generic models for structure. The notion of an example, and its relation to
generality, is an integral part of the discipline of mathematics. We claim that
any theory of learning which does not deal with how learners and teachers
act with, and on, examples is likely to be incomplete as far as mathematics is
concerned. (p. 77)
Bills and Watson (2008) continued to argue that many papers about mathematics
pedagogy have not problematised the examples in use, and so their structure and
sequencing in a mathematics lesson. Yet, these are “at the heart of the learners’
mathematical experiences and also at the heart of much classroom data” (p.77).
We agree, and this is reflected in our attention to exemplification through the
Mathematical Discourse in Instruction framework.
Studies of the forms and functions of teachers’ example-use have extended
to both elementary and secondary mathematics teaching, and to prospective and
experienced teachers. Rowland (2008) explored example-use across 24 lessons
taught by prospective elementary teachers. He identified four categories of
example-use: variability, sequencing, representations, and lesson objectives. These
analytic distinctions in turn provided insight into aspects of teachers’ mathematical
knowledge-in-use in teaching: that this entails variation across a set of examples,

335

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

their sequencing and link with lesson goals. Each of these aspects features in
Mathematical Discourse in Instruction.
Using their own experience of example-use when working on a task about
polynomial functions, and example-use in the lessons of an experienced secondary
teacher teaching decimals, fractions and percentages, Watson and Chick (2011)
reinforce “how careful and knowledgeable teachers need to be” to bring about
“alignment between the learners’ engagement and teacher’s intentions” (p. 294). Put
differently, the affordances of an example or an example set are not self-evident to
learners. It takes a knowledgeable teacher, with some fluency in example-use to
draw learners’ attention to, and enable their engagement with, what is significant for
their mathematics learning.
Arzarello, Ascari, and Sabena (2011) were also interested in the distance between
teachers’ goals and students’ actions. They argued that mediating this distance is
complex and required the teacher to promote appropriate students’ actions with
examples. The issue here is that the genesis of examples in students’ activity would
not necessarily produce a structured space. Working semiotically with classroom
data they illustrate that in the process of producing examples:
[the] teacher’s intervention can be crucial in helping the students to modify a
wrong example, to generate the right one for the task and to start the long-term
process of building up the structure of their own space of examples. (p. 295)
These results and arguments for the importance of teaching with examples and
teachers’ knowledge of these in their teaching are particularly interesting in the light
of earlier work. Zodik and Zaslavsky’s (2008) study of example-use by experienced
secondary teachers illuminated that they were not necessarily aware of their example-
use and related rationales. From their observations and analysis of teaching, and
thus teachers’ tacit knowledge-in-use, they distinguished between teachers’ pre-
planned use of examples, and their spontaneous use as these arose in the course of
teaching. They also revealed that example choices can either facilitate or impede
students’ learning, and consequently the choice of examples in teachers’ work is not
trivial. They went on to lament the lack of deliberate attention to exemplification in
mathematics teacher education:
… numerous mathematics teacher education programmes do not explicitly
address this issue and do not systematically prepare prospective teachers to
deal with the choice and use of instructional examples in an educated way.
(p. 166)
Building on this earlier work and the significance of example-use in mathematics
teaching, Zodik and Zaslavsky have argued further for its place as part of specialised
knowledge for teaching (Ball, Thames, & Phelps, 2008):
The knowledge teachers need for meeting the challenge of judiciously
constructing and selecting mathematical examples is a special kind of

336

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

knowledge. It can be seen as core knowledge needed for teaching mathematics.


… engaging teachers in generating or choosing instructional examples can be
a driving force for enhancing these elements of their knowledge. (Zodik &
Zaslavsky, 2009)4
The role of examples in mathematics teacher education itself has received some
attention. For example, Peled and Balacheff (2011) used carefully and appropriately
designed task examples to elicit, identify and effect a change in teachers’ conceptions
about processes of modelling and problem solving on the one hand, and about
the role of mathematics in solving these problems on the other. While this study
contributes to the knowledge base on teacher education practice, its focus is the role
of exemplification in teachers’ learning of mathematics for teaching, and not on
teachers’ knowledge of exemplification itself.
Zazkis and Leikin’s (2008) study of the role of examples in defining and definitions
comes closer to the call for preparing teachers to deal with instructional examples and
so the development of this specialized content knowledge for teaching mathematics.
Zazkis and Leikin presented a group of 40 prospective secondary teachers with the
task of giving “as many examples as possible for a definition of a square” (p. 134).
Their study built on the important role of learner-generated examples (Watson &
Mason, 2005) in students’ mathematics learning, and how these could provide a
window into students’ mathematical conceptualization. They were interested in the
prospective teachers’ concepts of a square in the first instance, and then their meta-
mathematical concept of a definition. An additional research question related to the
usefulness of a three-dimensional framework for analysing examples: accessibility
and correctness, richness, and generality. The student teachers’ definitions and related
example-use generated a large number of examples of definitions, including more
and less rigorous definitions, as well as some incorrect ones. The study confirmed
their initial assumption that the exploration of learner-generated examples would
provide insight into the teachers’ understanding of the definition of a square and of
defining itself; and further that the examples generated served as “a lens” on this
understanding. The study identified the different understandings illuminated.
The interesting aspect of Zazkis and Leikin’s (2008) study for this chapter was the
follow-on task given to the same group of prospective teachers later in the year. Zazkis
and Leikin presented the prospective teachers with 24 examples of definitions, most
of which were selected from their teacher-generated examples of definitions and
with additions of some produced by ‘experts’. The task for the prospective teachers
was to evaluate the validity of each of the 24 definitions provided. Of interest to
us was the discussion they described amongst the teachers about the validity of the
various definitions, and how these moved between mathematical validations and
pedagogical ones. Some teachers evaluated validity in terms of what they thought
would be appropriate for school teachers, without attention to related mathematical
rigour. Zazkis and Leikin concluded by suggesting that the tasks offered “a valuable
activity, both mathematical and pedagogical, to promote a deeper conceptual

337

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

understanding of mathematics in general and of the nature and role of definitions in


particular” (p. 147).
The critical point here is that these tasks in a teacher education setting were
mathematical and meta-mathematical, intended to strengthen teachers’ knowledge
of examples of definitions and defining. The tasks were not designed for explicit
attention to choosing and using instructional examples. The value of this study is
that it points to the value of such activities in mathematics teacher education, and the
specialized knowledge these can lever up for teachers.
Through our review above we have pointed to the considerable literature
foregrounding the significance of exemplification in mathematics education, and
to a focus on exemplification as specialised knowledge for teaching. This leads to
the question of what and how this is attended to in mathematics teacher education,
and hence the focus of this chapter. Before turning to this work, we briefly bring
back into focus the attention to variation in much of the work on exemplification.
The importance of variation in mathematics teaching has a long history, dating
back to Dienes’ (1960) work on perceptual variation. In mathematics education
variation has come back into focus in more recent years through the work of Watson
and Mason (Mason, 2017; Watson, 2017; Watson & Mason, 2006), particularly
through their vivid illustration of variance amidst invariance as a tool for engaging
with generality and with mathematical structure, and through carefully structured
example sets.
More generally, application of Variation Theory to mathematics education has
ranged from studies of variation in textbook example sets (e.g., Sun, 2011) to
teachers’ learning through lesson study informed by variation theory (e.g., Runesson,
2008) and to learners’ learning (e.g., Kullberg, Runesson Kempe, & Marton, 2017).
Kullberg et al. (op. cit.) emphasise how attention to variation can enable critical
features of the object of learning to come into focus. They make the important
observation that multiple examples are not simply cumulative. The ordering of
examples, their simultaneous presentation, and the teacher drawing attention to
similarities and differences are critical. We agree, and we have argued previously
that research related to an entire lesson needs to attend to the accumulating example
space.
The research on examples, [however,] while illuminating of what teachers do
and why, does not enable a view of whether and how examples accumulate
to bring the object of learning into focus for learners, and whether there is
movement towards generality. (Adler & Ronda, 2017, p. 68)
This adds to the earlier research discussed above which highlighted challenges
for teaching with examples. For Goldenberg and Mason (2008), because an example
is always “an example of something” learners’ attention needs to be drawn to the
general case of which the example is an instance. On the one hand this implies
knowledgeable use of generic examples–those particulars in which it is possible
to discern the general (Mason & Pimm, 1984). In Variation Theory (Watson &

338

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

Mason, 2006) terms, learners’ attention also needs to be drawn to key features of
a mathematical object such as aspects of mathematical structure the teacher wishes
to make visible. This is a function of a set of examples, how they are organized
to illuminate variance amidst invariance and thus possibilities made available for
generalizing, and/or recognizing structure or key features. Hence our work and the
focus and contribution of this chapter.
Of course, examples are always embedded in a task. Thus, while examples are
selected as particular instances of the general case in focus, and for drawing attention
to relevant features, generality and/or structure, tasks are designed to bring particular
capabilities to the fore (Marton & Pang, 2006; Marton & Tsui, 2004). For example,
expanding a(b+c) and factoring ab+ac are different tasks. A further capability
involves working on the relation between these tasks or doing and undoing in
Mason’s (1988) terms. Different tasks require different actions, at different levels of
complexity, and so make available different opportunities for mathematics learning.
In our work, we link examples and tasks in our consideration of exemplification,
since an example or an example set is always embedded in a task. Indeed, it is this
that makes an example ‘instructional’.
Having discussed the why of our focus on exemplification in the Wits Maths
Connect Secondary project and in the literature, we now turn to engage with
what it might mean to have a deliberate focus on exemplification in mathematics
teacher education. We describe and illustrate the what and how of our professional
development work, and particularly how we model and mediate exemplification
with teachers. It is relatively easy to describe what we do in broad terms, but it is not
trivial to articulate how this is actually organised and accomplished.

EXEMPLIFICATION: THE WHAT AND HOW

As we noted earlier, the Mathematical Discourse in Instruction framework informs


our teaching of our mathematics-for-teaching course called Transition Maths 1. We
focus on the course since it is the major context in which we deliberately “teach”
exemplifying/exemplification as a key mathematics teaching practice. Traditional
Maths 1 is structured so that approximately two-thirds of the time teachers focus
on their own learning of mathematics. In these mathematics sessions, we model
exemplification as a mathematics teaching practice. Teachers work on mathematical
tasks where the objects of learning are key mathematical concepts, procedures and/
or practices, the selection of which has been influenced by the South African school
curriculum. In general, these tasks and activities provide opportunities to revisit and
deepen their knowledge of the mathematics they teach (Pournara, 2013; Zazkis,
2011), as well as activities that extend their knowledge of school mathematics. From
this activity they build generality, focus on mathematical structure and engage with
mathematical procedures and their rationales. The tasks and example sets offered
to teachers are carefully selected to model and illustrate the forms of variation
described earlier. For example, the task in Figure 12.2a deals with informal methods

339

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

Figure 12.2. Mathematics tasks for quadratic equations and trigonometric notation

of finding solutions to quadratic equations in factorised form. We ask teachers to


find numerical values without using formal procedures. We also ask them to reflect
on “what changes” and “what stays the same” and to consider the impact of this
variation on how they approached each focus.
The task in Figure 12.2b deals with trigonometric notation – a source of
considerably difficulty for teachers who have not done trigonometry for many years.
The examples provide opportunity to address typical errors and misinterpretations
of the notation, leading to questions such as “can I distribute the sin into the bracket
in (b)?” and “what difference does it make if the sin is not in the numerator in
(e)?” Once again these are mathematical questions posed by teachers to clarify the
mathematics for themselves.
While the course presenter frequently points to the variance amidst invariance in
the example sets, this is to mediate the mathematics in play and only an implicit form
of drawing attention to exemplification with variation. It is in the teaching sessions
that we deliberately mediate exemplification as a teaching practice.
In the remaining one-third of the course, which is our particular interest in this
chapter, teachers focus on their mathematics teaching. These teaching-focused
sessions are structured to mediate all the components of the Mathematical Discourse
in Instruction framework. We work from the assumption that better teaching
is characterised by more thoughtful selections of examples and tasks, and by
mathematical explanations that focus explicitly on the mathematics the teacher
intends the learners to learn. With respect to exemplification, we work with teachers
on articulating the mathematical goals for a lesson (objects of learning), and then
on choosing and using examples. Using principles of variation, we examine sets
of examples that either we have constructed or are available in textbooks or in
a prescribed lesson plan, to ascertain what is possible to come into focus. A key
strength here is that our focus is on issues that are sufficiently close to teachers’
current practice, and to curriculum demands, as to be possible to implement.
In a teaching session we might use a task similar to one used in an earlier
mathematics session but the focus of attention shifts. For example, in an example set
of equations, attention would be specifically directed to similarities and differences
between the different equations, their sequencing, and the extent to which the

340

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

example set might achieve the lesson goal. In the remainder of this chapter, we
elaborate our attention to exemplification as a key focus in our work and specifically
how we mediate this with teachers.

MEDIATING EXEMPLIFICATION IN MATHEMATICS TEACHER EDUCATION

In this section we provide three illustrative cases of how we work with exemplification
using variation in professional development. The three cases deal with number,
algebra and function all of which are given substantial attention in the course. We
distinguish between the learner task and the teacher education task, and hence make
explicit what it is about exemplification that we intend teachers to learn. In each
case, the choice of the example set, and its sequencing are in focus, so as to make
visible what it is teachers are to attend to. This means that while there is a learner
task, learner thinking per se is in the background.
Case 1 illustrates how we introduce teachers to ideas of variation in an example
set. Case 2 extends ideas of variation in an example set to focus on connections
between representations, leading to generalisation. Case 3 deals with applying
ideas of variation to produce a new example set. The three cases also illustrate
a progression in our mediation of exemplification. It is only in Case 3, and after
teachers have engaged with tasks like those in Cases 1 and 2, that we expect them
to be able to apply principles of variation to design a new task with coherent links
between the object of learning and the example set, and then to critique and revise
their example set.

Case 1 – Introducing Teachers to Ideas of Variation in an Example Set to Address


Learner Error

Our first case has its roots in lesson study work with teachers who had already
completed the course and so had been introduced to variation.5 It connects directly
into teachers’ practices in two ways: (1) it deals with a prevalent and persistent error
in the application of the distributive law and the use of brackets; and (2) it deals with
meaning of algebraic forms. We have drawn on this very specific problem of practice
to construct a learning opportunity in teacher education where we introduce teachers
to ideas of variation in an example set which deals with the use of the distributive law.
The learner task is framed by the following object of learning: “learners must
be able to simplify expressions with brackets that appear in different positions”
and contains the example set in Figure 12.3. Teachers would typically ask learners
to attempt the task individually and may then invite learners to work in pairs to
compare their answers. Thereafter the answers might be discussed in a whole-class
setting. The teacher would then draw attention to what is the same and different
about each of the law.

341

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

Figure 12.3. Learner example set involving application of the distributive law

Figure 12.4. Teacher example set for application of the distributive law

In this example set invariance lies in the selection and order of symbols (numeric
and algebraic). Variance is introduced in how the symbols are combined through
operations and the position of brackets.
In the teacher education task, the five examples are set up as a collection of
pairs of expressions, numbered 1–8, and with answers provided for convenience
(see Figure 12.4). These pairs are carefully juxtaposed to focus on particular learner
errors. In this way teachers are invited to compare the following pairs from the
learner task: (a)-(b), (a)-(c), (a)-(d) and (d)-(e). In comparing (a) and (b), we address
the common error where learners do not consider a letter to the right of the bracket
to be an instance of the distributive law. By contrast, in comparing (d) and (e) we
address the overgeneralisation “brackets mean multiply.”

342

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

When the task is presented as pairs of expressions, the contrast is more explicit
because of the juxtaposition of items with minor visual differences. Clearly the
effects of the variance are only discernible when one acts on the expressions. We do
not assume that teachers can see the full purpose of the selection of examples and the
pairings without actually completing the examples themselves. So we expect them
to do the simplification before any discussions begin. We then ask teachers three key
questions which we ask of all example sets: “what varies?”, “what is invariant”? and
“what mathematics is possible to learn from this variation?”
This teacher education task provides several learning opportunities for teachers.
Firstly, teachers see that working with variance amidst invariance provides a
teaching strategy for choosing and/or designing example sets to focus on particular
learner errors. Secondly, by focusing on pairs of examples, it is easier for teachers
to identify variance amidst invariance. This in turn shows how juxtaposition with
minor variation has the potential to bring the object of learning more clearly into
focus than it might in the original learner task.
Having identified variance amidst invariance, the next step is for teachers to
produce a similar pairing and then a full example set related to the distributive law.
The extent to which teachers can do this successfully gives us some insight into the
sense they have made of this introduction to the principles of variation as well as
their grasp of the structural aspects of the mathematics in focus. In Figure 12.5 we
illustrate two typical pairings that teachers propose.
In Figure 12.5a, the pairing draws attention to the matter of “do the brackets
first.” This new addition will provide the only instance in the example set where
the bracket can first be simplified. Thereafter it is similar to example (2) in Figure
12.4, where the 5 is then added. There is thus further potential for juxtaposition
with another example in the set. In Figure 12.5b the pairing draws attention to the
distinction between sign and operation. The bracket in the new expression shifts
the meaning from “add 3” to “positive 3.” This inevitably leads to some discussion
about whether the new example maintains the focus on the distributive law or
whether the focus on sign versus operation diverts attention away from the intended
object of learning.
We have learnt that teachers are easily able to identify the surface features of
the variation and to produce their own examples of variation. However, in doing
so, they may lose focus on the object of learning. Consequently, their suggested

Figure 12.5. Teachers’ extensions of the given example set

343

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

changes may simply generate an expression that varies rather than maintaining
focus on the intended object of learning. So we recognise this as part of the
journey of learning to work with principles of variation – that in the early stages,
variation may become an end in itself rather than a means to achieve the object
of learning.

Case 2: Extending Ideas of Variation in an Example Set to Attend to Connections


between Representations and to Generalise

This case is similar to the first in that it is drawn from lesson study work,6 and
therefore is also close to teachers’ practice. Here the example set (Figure 12.6) is
constructed to lead learners to generalise the impact of parameters on the graph of the
quadratic function y = ax 2 + q and hence to make connections between the equation
and the graph. In the original example set, the teacher dealt with the effects of a and
a
q on the rational function y = + q . In our first adaptation of the task for the course
x
we used the rational function but too many teachers had difficulty with the underlying
mathematics and so they were unable to pay attention to the aspects of exemplification

Figure 12.6. Example set for learner task to match equations and graphs

344

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

and variation that we sought to foreground. We therefore replaced the rational function
with the quadratic function.
The learner task involved a card matching activity where learners were given an
example set containing six equations and six graphs on separate cards. They were
required to match each graph with an equation. As the lesson progressed, the teacher
worked with learners to generalise the impact of a change in the sign of a and the
value of q on the graph.
In the session, teachers were first required to complete the card matching task.
They spotted the “twist” designed into the original example set by the teacher
in the lesson study: that equations 5 and 6 are the same and therefore there is no
corresponding equation for graph C. Hence the learner task also involved producing
an equation for graph C.
The teacher education task required teachers to compare the pairs of equations
and graphs shown in Figure 12.7, and to identify what varies, what is invariant and
what mathematics is possible to learn through the variation.
As can be seen in the example set for the teacher education task, equation 1/graph
D is held constant and the other member of the pair changes. This is intended to
draw teachers’ attention to how one might structure an investigation of the impact of
two parameters across two different representations. Teachers were able to identify
that in (a) and (b) the focus was the impact of q on the vertical position of the
graph, and in (c) attention moves to the impact of the sign of a on the orientation of
the graph. We then invited teachers to choose their own pair of equations/graphs to
compare and to identify what mathematics was possible to learn from the pairing. As
expected, a common pairing was equation 2/graph B and equation 5/graph E which
drew attention to the effect of q. We were encouraged that many groups also selected
pairs that led to the following generalisations: “if a and q have the same sign, then
the graph has no roots,” and conversely, “if a and q have opposite signs, then the
graph has two roots.”

Figure 12.7. Teacher task for function matching task

345

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

The use of juxtaposition in the teacher education task provides a structure for
introducing variation in task design when more than one representation and more
than one feature are in focus. In other words, the prompts for teachers provide a
scaffold to think about which elements to attend to when designing future card
matching tasks, and how to vary these features. Simultaneously the design of the
teacher education task suggests a possible teaching strategy for making the learner
task more accessible to learners who may have difficulty in dealing with twelve
different cards all at once.
Cases 1 and 2 provide some evidence of teachers’ take up of the principles of
variation in extending example sets during the course. We do not yet have evidence
of the extent to which teachers are able to do this in their own practice. However,
case 3 provides some insights into the challenges teachers may face in designing
their own example sets.

Case 3: Applying Ideas of Variation to Produce a New Example Set towards an


Articulated Goal

The third case is drawn from teacher education practice where teachers were invited
to choose an object of learning and to design an example set to achieve their object
of learning. The initial prompts for this task were short example sets (such as the
one in Figure 12.8) which we had designed as lesson starters to revise topics. We
have selected the work of a group of teachers who focused on a lesson starter task
for integer operations.
The learner task consisted of six examples involving subtraction and multiplication
of single-digit numbers. While 2 and 5 are the only digits that appear in the example
set, four examples include brackets and learners would need to distinguish between
sign and operation in order to determine the impact of the brackets.
As with the examples for the distributive law, the six examples here could be
usefully juxtaposed to emphasise particular issues, for example, (a)-(b) deals with
the commutative law and the well-known error of working right-to-left to avoid
subtracting a larger number from a smaller number (Vlassis, 2004). The juxtaposition

Figure 12.8. Learner task for operations on integers

346

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

of (b)-(c) addresses the sign/operation duality. The juxtaposing of (c)-(d)-(e) brings


attention to the (implicit) operation of multiplication in contrast to the explicit sign
in (c) and (d).
The teacher education task involved teachers working in groups on the following
sequence of tasks:
‡ Choose a particular object of learning and adapt the example set accordingly
‡ Make the object of learning explicit and motivate the changes made to the
example set
‡ Produce a poster to communicate your work with peers
‡ All groups give feedback to each other with particular focus on the example set
‡ Revise example set based on feedback and motivate additional changes or justify
rejecting advice of peers
We describe briefly below the sequence of events in this group’s progress because
this illustrates their take-up of ideas of example sets and variation within these. The
teachers first decided to retain the example set in Figure 12.8 and added the four new
examples in Figure 12.9a. They described the object of learning for their adapted
example set as: “learners must add, subtract and multiply single-digit integers.”
Their rationale for these four examples included attention to the operation (addition
vs subtraction), the position of the bigger number (first vs second) and the reversal
of the numbers.
In providing feedback, the other groups suggested they should focus only on
addition and subtraction and so in the first revision they removed all the examples
with brackets and proposed the six examples shown in Figure 12.9b, two of which
were in the original prompt with the other four being those they had introduced. They

Figure 12.9. Successive adaptations of the object of learning and example sets by teachers

347

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

revised the object of learning as follows: “learners must be able to add two integers,
subtract two integers; and subtract a smaller number from a bigger number.”
The revised object of learning is more clearly-specified with closer attention to the
relative size of the two numbers. The somewhat surprising reference to subtracting
a smaller number from a bigger number may suggest some confusion between sign
and operation, since they were referring to (b), (e) and (f) when they discussed
their rationale. Alternatively, it may reflect a “slip” in seeing, for example, –5 – 3
as subtracting –5 from 3. The example set contains three obvious pairs where the
juxtaposition suggests teachers intended learners to pay attention to the commutative
law. However, this was not the case; it is not indicated in the object of learning and
when we asked what learners might generalise from the example set, the teachers did
not mention the commutative law.
After writing down the answers to each example, we suggested that learners
might generalise about the order of the numbers and might thus conclude that “order
matters” for the first pair but not for the other pairs. Through our mediation, we
drew attention to the commutative law and that it did not hold for subtraction in the
first pair but did hold for the other pairs, including the (apparent) subtraction in the
third pair. Clearly this would be a contradiction of known mathematical laws and
potentially misleading for learners. Following from this, the teachers’ discussion
inevitably shifted to focus on sign and operation which had not been in focus earlier.
This led them to revise their example set to distinguish clearly between sign and
operation and to focus only on addition, indicating that a follow-up example set
should focus on subtraction.
The second revision of the example set (Figure 12.9c) still includes three pairs.
Teachers changed all 3’s to 2’s in order to reduce the variation and because this had
no significant impact on the example set. Given the new object of learning relating to
the impact of the order of numbers for addition, the example set provides instances
of similarity for (c)-(d) and (e)-(f) showing that order does not matter, whereas in
for (a)-(b) order does matter although the numbers being added in each case are not
the same.
Through this teaching task, teachers had opportunity to work with variance amidst
invariance as a tool for designing example sets to focus on topics that learners have
difficulty with. There was opportunity to revise the example set which typically led
to a clearer focus in the example set and a more clearly articulated object of learning.
The task also required greater attention to the relationship between the example set
and the object of learning when changes were made to both. The opportunity to
comment on the work of peers provided additional opportunity to work with ideas of
variation in relation to exemplification.
This case gives us insight into the range of issues that need attention when
constructing a carefully designed and focused example set. While teachers had paid
attention to the pairs they were juxtaposing, they did not pay sufficient attention to
what might be generalised from the combination of the pairs. We are reminded here

348

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

of Kullberg et al.’s (2017) warning that collections of examples are not necessarily
cumulative. In this instance the potential generalisation in Figure 12.9b would
have been incorrect. The work required to explain how this contradiction arose is
substantially mathematical and draws on specialised knowledge for teaching, in this
case the distinction between sign and operation.

DISCUSSION AND CONCLUSION

Our starting assumption, discussed earlier in the chapter, is that teaching, whatever
its context and/or pedagogy is purposive work. At the heart of this chapter is how
we work with teachers to develop their purposive and deliberate choice and use
of examples in their teaching. We have argued both from the growing literature
base and from our own research with teachers in the project that focusing on
exemplification in conjunction with principles of variation, and in particular
attention to variance amidst invariance, is an important and necessary component of
secondary mathematics teachers’ professional development. Furthermore, we have
argued that exemplification can be considered a high leverage practice that is used
frequently in teachers’ mathematical work and promotes learners’ mathematical
learning. With this rationale in place, we proceeded to elaborate the what and how of
exemplification informed by variation.
We described and discussed what we focus on and how we mediate exemplification
in our mathematics professional development course with teachers, through three
cases selected to illustrate three important features of this work. Firstly, separating
the learner and teacher education task is critical for being able to focus teachers’
attention on what it is they are to be coming to know and be able to do, that is,
principles of variation and how to apply these to structure focused example sets.
Secondly, the mathematical task for the learners needs to be familiar to teachers
so that their attention is on the learning of exemplification. We note that there may
still be mathematical learning opportunities for teachers, even in what could be
considered basic knowledge for secondary school teachers such as operations on
integers. Thirdly, it is important to organise the teacher education tasks so that there
is progression from becoming familiar with principles of variation at work in an
example set, to being able to work with these when there are two (and possibly
more) representational forms, and finally to applying these principles to constructing
such sets.
In conclusion, then, we reflected first on the nature of the tasks we use and then
on some of the challenges we have faced in our work – specifically mediating
knowledge and practice in relation to exemplification. We have indicated
how our tasks for mediating aspects of mathematics teaching, and specifically
exemplification, remain close to teachers’ practices, which as we described earlier
are predominantly ‘traditional’. Much of the literature on exemplification tends to be
related to teaching with rich tasks and inquiry-based pedagogies. We hope we have

349

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

shown the importance of explicit attention to exemplification in relation to more


traditional tasks focusing on key concepts and procedures in school mathematics to
support teachers’ learning to improve their practice.
Finally, in the discussion of the three cases, we have pointed to some challenges
as teachers engage with exemplification informed by principles of variation. Most
critical is that teachers easily notice variation at a visual level such as changes in
numbers, letters, orders of symbols, etc. While this is an important first step, it is
insufficient to engage only with the visual features of an example in its particular
representation. For example, example sets of algebraic equations do not immediately
reveal the nature of their solutions. Drawing teachers’ attention to variance amidst
invariance is not trivial if it is to move them beyond superficial use of such in
their teaching. We are cognisant that there is much still to explore in working with
mathematics teachers on examples and example sets. We hope that by sharing our
enthusiasm, and implicitly that of the teachers with whom we have worked on
exemplification, we elicit others’ interest in pursuing this line of work in teacher
education and research on mathematics teachers’ learning.

ACKNOWLEDGEMENT

This work is based on the research supported by the South African Research
Chairs Initiative of the Department of Science and Technology and National
Research Foundation (Grant No. 71218). Any opinion, finding and conclusion or
recommendation expressed in this material is that of the author(s) and the National
Research Foundation does not accept any liability in this regard.

NOTES
1
We use the terms learner/s and student/s interchangeably.
2
We have recruited but recontextualised the notion ‘object of learning’ from variation theorists (Marton
& Tsui, 2004), cognisant that our theoretical antecedents are different.
3
We use instructional examples and pedagogical examples interchangeably. These refer to examples in
use in mathematics teaching.
4
In identifying exemplification with specialized content knowledge, and referring simultaneously
to instructional examples, the literature here points to the blurred boundary between specialized
content knowledge (SCK) and knowledge of content and instruction (an element of Pedagogical
Content Knowledge) in Ball et al.’s (2008) terms. It is beyond the scope of this article to take up
the issue of definition of these terms in the literature. Much has been written of varied use of terms
related to mathematical knowledge for teaching in our field (e.g., Hoover, Mosvold, Ball, & Lai,
2016).
5
See Adler and Alshwaikh (2019) for a detailed discussion of this lesson study cycle, where there is a
focus on how exemplification was key in the planning, reflection and revision of the lesson, and in
teachers’ learning.
6
The lesson study cycle from which this case was constructed is reported in Adler and Ronda (2017),
where the rational function was in focus.

350

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

REFERENCES
Adler, J. (2001). Teaching mathematics in multilingual classrooms. Dordrecht: Kluwer.
Adler, J. (2017). Mathematics Discourse in Instruction (MDI): A discursive resource as boundary object
across practices. In G. Kaiser (Ed.), Proceedings of the 13th International Congress on Mathematical
Education, ICME-13 (pp. 125–143). Cham: Springer.
Adler, J., & Alshwaikh, J. (2019). A case of lesson study in South Africa. In R. Huang, A. Takahashi, &
J. da Ponte (Eds.), Theory and practices of lesson study in mathematics: An international perspective.
Dordrecht: Springer.
Adler, J., & Ronda, E. (2015). A framework for describing mathematics discourse in instruction and
interpreting differences in teaching. African Journal of Research in Mathematics, Science and
Technology Education, 19(3), 237–254. doi:10.1080/10288457.2015.1089677
Adler, J., & Ronda, E. (2017). Mathematical discourse in instruction matters. In J. Adler & A. Sfard
(Eds.), Research for educational change: Transforming researchers’ insights into improvement in
mathematics teaching and learning (pp. 64–81). Abingdon: Routledge.
Adler, J., & Venkat, H. (2014). Teachers’ mathematical discourse in instruction: Focus on examples and
explanations. In H. Venkat, M. Rollnick, J. Loughran, & M. Askew (Eds.), Exploring mathematics and
science teachers’ knowledge: Windows into teacher thinking (pp. 132–146). Abingdon: Routledge.
Alexander, R. (2000). Culture and pedagogy: International comparisons in primary education. Oxford:
Blackwell.
Antonini, S., Presmeg, N., Mariotti, M. A., & Zaslavsky, O. (2011). On examples in mathematical
thinking and learning. ZDM-The International Journal on Mathematics Education, 43(2), 191–194.
doi:10.1007/s11858-011-0334-5
Arzarello, F., Ascari, M., & Sabena, C. (2011). A model for developing students’ example space: The
key role of the teacher. ZDM-The International Journal on Mathematics Education, 43(2), 295–306.
doi:10.1007/s11858-011-0312-y
Ball, D. (2017). Uncovering the special mathematical work of teaching. In G. Kaiser (Ed.), Proceedings of
the 13th International Congress on Mathematical Education, ICME-13 (pp. 11–34). Cham: Springer.
Ball, D., Sleep, L., Boerst, T., & Bass, H. (2009). Combining the development of practice and the practice
of development in teacher education. The Elementary School Journal, 109(5), 458–474.
Ball, D., Thames, M., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407. doi:10.1177/0022487108324554
Bills, L., Dreyfus, T., Mason, J., Tsamir, P., Watson, A., & Zaslavsky, O. (2006). Exemplification in
mathematics education. In J. Novotná, H. Moraová, M. Krátká, & N. Stehliková (Eds.), Proceedings
of the 30th Conference of the International Group for the Psychology of Mathematics Education (pp.
126–154). Prague, Czech Republic: PME.
Bills, L., & Watson, A. (2008). Editorial introduction. Educational Studies in Mathematics, 69(2), 77–79.
Dahlberg, R. P., & Housman, D. L. (1997). Facilitating learning events through example generation.
Educational Studies in Mathematics, 33(3), 283–299. doi:10.1023/A:1002999415887
Dienes, Z. (1960). Building up mathematics. London: Hutchinson Educational Ltd.
Goldenberg, P., & Mason, J. (2008). Shedding light on and with example spaces. Educational Studies in
Mathematics, 69(2), 183–194. doi:10.1007/s10649-008-9143-3
Grossman, P., Hammerness, K., & McDonald, M. (2009). Redefining teaching, re-imagining
teacher education. Teachers and Teaching: Theory and Practice, 15(2), 273–289.
doi:10.1080/13540600902875340
Hoover, M., Mosvold, R., Ball, D., & Lai, Y. (2016). Making progress on mathematical knowledge for
teaching. The Mathematics Enthusiast, 13(1–2), 3–34.
Kozulin, A. (2003). Psychological tools and mediated learning. In A. Kozulin, B. Gindis, V. Ageyev,
& S. Miller (Eds.), Vygotsky’s educational theory in cultural context (pp. 15–38). New York, NY:
Cambridge University Press.
Kullberg, A., Runesson Kempe, U., & Marton, F. (2017). What is made possible to learn when using
the variation theory of learning in teaching mathematics? ZDM Mathematics Education, 49(4),
559–569.

351

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
JILL ADLER & CRAIG POURNARA

Marton, F., & Pang, M. F. (2006). On some necessary conditions of learning. The Journal of the Learning
Sciences, 15(2), 193–220.
Marton, F., & Tsui, A. B. M. (2004). Classroom discourse and the space of learning. Mahwah, NJ:
Lawrence Erlbaum Associates.
Mason, J. (1988). Project mathematics update: Doing and undoing. Milton Keyes: Open University Press.
Mason, J. (2017). Issues in variation theory and how it could inform pedagogic choices. In R. Huang & Y.
Li (Eds.), Teaching and learning mathematics through variation: Confucian heritage meets Western
theories (pp. 407–438). Rotterdam, The Netherlands: Sense Publishers.
Mason, J., & Pimm, D. (1984). Generic examples: Seeing the general in the particular. Educational
Studies in Mathematics, 15(3), 277–289.
Nachlieli, T., & Tabach, M. (2018). Ritual-enabling opportunities to learn in mathematics classrooms.
Educational Studies in Mathematics. (Advance online publication) doi:10.1007/s10649-018-9848-x
Peled, I., & Balacheff, N. (2011). Beyond realistic considerations: Modeling conceptions and controls
in task examples with simple word problems. ZDM-The International Journal on Mathematics
Education, 43(2), 307–315. doi:10.1007/s11858-011-0310-0
Pournara, C. (2013). Mathematics-for-teaching in pre-service mathematics teacher education: The case
of financial mathematics (Doctoral dissertation). University of Witwatersrand, Johannesburg, South
Africa. http://hdl.handle.net/10539/18836
Pournara, C., Hodgen, J., Adler, J., & Pillay, V. (2015). Can improving teachers’ knowledge of mathematics
lead to gains in learners’ attainment in mathematics? South African Journal of Education, 35(3), 1–10.
doi:10.15700/saje.v35n3a1083
Rowland, T. (2008). The purpose, design and use of examples in the teaching of elementary mathematics.
Educational Studies in Mathematics, 69(2), 149–163.
Runesson, U. (2008). Learning to design for learning. In P. Sullivan & T. Wood (Eds.), Knowledge and
beliefs in mathematics teaching and teacher development (pp. 153–172). Rotterdam, The Netherlands:
Sense Publishers.
Sun, X. (2011). “Variation problems” and their roles in the topic of fraction division in Chinese
mathematics textbook examples. Educational Studies in Mathematics, 76(1), 65–85. doi:10.1007/
s10649-010-9263-4
Vlassis, J. (2004). The role of mathematical symbols in the development of number conceptualization: The
case of the minus sign. Philosophical Psychology, 21(4), 555–570. doi:10.1080/09515080802285552
Vygotsky, L. S. (1978). Mind in society: The development of higher psychological processes. Cambridge,
MA: Harvard University Press.
Watson, A. (2017). Pedagogy of variations: Synthesis of various notions of variation pedagogy. In R.
Huang & Y. Li (Eds.), Teaching and learning mathematics through variation: Confucian heritage
meets western theories (pp. 85–103). Rotterdam, The Netherlands: Sense Publishers.
Watson, A., & Chick, H. (2011). Qualities of examples in learning and teaching. ZDM-The International
Journal on Mathematics Education, 43(2), 283–294. doi:10.1007/s11858-010-0301-6
Watson, A., & Mason, J. (2005). Mathematics as a constructive activity: Learners generating examples.
Mahwah, NJ: Lawrence Erlbaum.
Watson, A., & Mason, J. (2006). Seeing an exercise as a single mathematical object: Using variation to
structure sense-making. Mathematical Thinking and Learning, 8(2), 91–111.
Zaslavsky, O. (2017). There is more to examples than meets the eye: Thinking with and through
mathematical examples in different settings. The Journal of Mathematical Behavior. (Advance online
publication) doi:10.1016/j.jmathb.2017.10.001
Zazkis, R. (2011). Relearning mathematics: A challenge for prospective elementary school teachers.
Charlotte, NC: Information Age Publishing.
Zazkis, R., & Leikin, R. (2008). Exemplifying definitions: A case of a square. Educational Studies in
Mathematics, 69(2), 131–148. doi:10.1007/s10649-008-9131-7

352

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
EXEMPLIFYING WITH VARIATION IN TEACHER EDUCATION

Zodik, I., & Zaslavsky, O. (2008). Characteristics of teachers’ choice of examples in and for the
mathematics classroom. Educational Studies in Mathematics, 69(2), 165–182. doi:10.1007/s10649-
008-9140-6
Zodik, I., & Zaslavsky, O. (2009). Teachers’ treatment of examples as learning opportunities. In
M. Tzekaki, M. Kaldrimidou, & C. Sakonidis (Eds.), Proceedings of the 33rd Conference of the
International Group of the Psychology of mathematics education (Vol. 5, pp. 425–432). Thessaloniki:
PME.

Jill Adler
School of Education
University of Witwatersrand

Craig Pournara
School of Education
University of Witwatersrand

353

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS AND SEÁN DELANEY

13. MATHEMATICS TEACHING PRACTICES AND


PRACTICE-BASED PEDAGOGIES
A Critical Review of the Literature Since 2000

Two exciting lines of research have emerged or have been expanded over the past two
decades in relation to the content and practice of mathematics teacher education. The
first concerns the attempt to theoretically identify and codify practices of teaching
in general, and core or high leverage practices, in particular. The second concerns
the identification of practice-based pedagogies of teacher education that may
encourage the teaching and learning of the latter practices. Reviewing key findings
emanating from research that has been undertaken since 2000 in these two realms of
mathematics education, we discuss the advances made in rendering practice a key
aspect of understanding and improving teaching and teacher education. Adopting
a critical stance, we also point to challenges in conducting research in these areas,
including the lack of shared language, the absence of an agreed-upon suite of
methodologies to empirically examine theoretical arguments advanced in these
areas, and the need for stronger and more systematic empirical validation of the
potential of teaching practices and practice-based approaches to teacher education.
In doing so, we outline open issues worthy of investigation in the next decade in each
line of research separately and jointly.

INTRODUCTION

In his classic book, Schoolteacher, American sociologist Dan Lortie (1975) described
teacher isolation as one of the main impediments to teachers’ learning in and from
practice. Teachers, Lortie lamented, spend much time isolated from other adults,
largely interacting only with students. As a result, schools become sites for student
learning and only infrequently for teacher learning. For decades scholars have
advocated incorporating teacher learning into the work of teaching. Ball and Cohen
(1999) called for a practice-based curriculum in teacher education, a curriculum that
sets critical examination of the practice of teaching at its core. How teachers learn
joined the concern about what teachers need to learn with an argument that teaching
“must be learned in and from practice rather than in preparing to practice” (p. 12).
For such learning to take place, “professional development needs to be grounded in
the actual tasks, questions, and problems of practice” (p. 20).

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_014

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

Viewing practice as a potential cornerstone of teacher learning, has led to


concerted efforts to better understand the work of teaching and the practice-based
environments that can support teacher learning for over two decades now. Key to
these efforts have been attempts to not simply document the complexities of teaching
but to codify and explicitly communicate what is entailed in the work of teaching
(e.g., Cohen, 2011; Lampert, 2001). Lampert (2001) provides a comprehensive
account of what teaching entails and how its complexity can be managed. Along
similar lines, Cohen (2011) identifies questions to consider if teacher learning and
consequently instructional quality is to be improved: “What sort of an endeavor is
teaching? What kinds of problems must teachers solve, and how do they solve them?
And what would it take to solve them in ways that promote ambitious teaching and
learning?” (p. 3).
Two decades after Ball and Cohen’s call for a practice-based curriculum in
initial teacher education and ongoing professional development, analyzing practice
has become a key aspect of understanding and improving teaching and teacher
education. Therefore, it seems opportune to review what has been accomplished so
far in this area of mathematics education and to consider challenges and open issues
for future research. To do this we reviewed research published since 2000 on two
cutting-edge lines of research in mathematics teacher education. The first pertains
to codifying and understanding teaching practices that are learnable by teachers and
teachable by teacher educators, in the service of student learning; the second relates
to identifying practice-based pedagogies that facilitate the teaching and learning of
such practices.
In what follows, we first outline the methods pursued in identifying relevant
literature. In the next two sections, we focus on each strand – teaching practices and
practice-based pedagogies–discussing what has been accomplished and considering
challenges and open issues. We conclude by identifying issues and challenges for
future research.

METHODS

Figure 13.1 summarizes the steps taken to identify, annotate and synthesize relevant
literature. First, relevant databases and a set of keywords were identified for initially
screening suitable sources, guided by clear inclusion and exclusion criteria. For the
first strand, we included the keywords “practices,” “teaching practices,” “practices
of teaching,” “core (teaching) practices,” “high leverage (teaching) practices,”
“instructional practices” along with “mathematics”; for the second strand, we
considered keywords such as, “approximation of practice,” “representation of
practice,” “decomposition of practice,” “signature pedagogy,” “practice-based
pedagogy,” “rehearsals,” along with “mathematics.” We did not include keywords
such as “lesson study” or “video-viewing”/”video-clubs” because, as we argue later,
we consider them as being included under the overarching umbrella of representations,
decomposition and approximations of practice. Additionally, including these terms

356

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

Figure 13.1. Flowchart of the literature search and annotation

357

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

would result in an even longer list of studies which would be hard to handle in this
chapter. We, however, point the reader to (review) studies focusing on these topics.
Peer-reviewed journal articles, books and chapters, including only texts written
in English were sought. Conference proceedings and dissertations were excluded.
Chapters from key handbooks in the field were examined to complement the search.
These included the fourth edition of the Handbook of Research on Teaching (2001),
the Handbook of Research on the Psychology of Mathematics Education: Past,
Present, and Future (2006), the Second Handbook of Research on Mathematics
Teaching and Learning (2007), the four volumes of The International Handbook
of Mathematics Teacher Education (2008), the third edition of the Handbook of
Research on Teacher Education: Enduring Questions in Changing Contexts (2008),
the fifth edition of the Handbook of Research on Teaching (2016), and the third
edition of the Handbook of International Research in Mathematics Education
(2016). We skimmed chapters of these handbooks, looking for themes related to
teaching practices and practice-based pedagogies. Attempting to be more inclusive,
we included chapters that referred to teaching practice(s), as well as chapters that
pertained to teachers’ learning in general. Finally, three books related to the themes
were included: Lampert’s (2001) Teaching Problems and the Problems of Teaching,
Cohen’s (2011) Teaching and its Predicaments, and Grossman’s (Ed., 2018) book on
Teaching Core Practices in Teacher Education.
Next, articles identified were screened. We worked independently, reviewing
each abstract and justifying decisions for inclusion or exclusion. Sources for which
there was agreement on their unsuitability were dropped; at least one vote for a
source merited its inclusion. Excluded were papers focused primarily on policy
issues, teachers’ beliefs about practice, teachers’ knowledge and motivation about
teaching, measurement approaches in studying instructional quality, and the role of
technological advancements (e.g., whiteboards) in improving instructional quality.
We also excluded documents without an explicit focus on mathematics and its
teaching or generally if its focus was tertiary education, thus limiting our review to
pre-primary, primary, and secondary school education. During the third stage of our
search, we retained a couple of articles on tertiary education, which were deemed
to contribute something innovative or illustrative to the field. This initial screening
resulted in dropping 100 articles.
Third, the identified resources were read and annotated. Each document was
entered on an Excel spreadsheet, recording the focal points that appear at the bottom
of Figure 13.1. Following this step, 22 additional documents not meeting the criteria
for inclusion were excluded. In addition, four articles cited in the initial documents
but not identified in our search were included. As a result, our literature synthesis
is based on 149 documents (marked with an asterisk in the reference/bibliography
list). In the reference list we also present six chapters in Grossman’s edited 2018
book separately.
Fourth, we read all articles and independently identified initial themes under two
categories: (i) Practices of teaching and (ii) Practice-based pedagogies in teacher

358

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

education (or combinations thereof). These initial themes were compared and
common themes were identified, clarified and refined. Where themes emerged that
were not common, agreement was sought on their inclusion or exclusion through
discussion. Common themes were chosen which were deemed to represent either
advances made in the fields under exploration or to suggest (explicitly or implicitly)
open issues and areas where additional work is needed. Having identified themes,
we next sought similarities and differences in the literature (e.g., what teaching
practices have been identified or what practice-based pedagogies have been studied
so far; what evidence exists about their contribution to either improving teaching
quality and/or student learning). Study features such as the teacher population, the
context and the research methods used further informed the chosen themes. Below,
we present the themes that emerged from this analysis organized into two main
sections, one corresponding to practices of teaching and the second pertaining to
practice-based pedagogies in teacher education.

ATTENDING TO PRACTICES ENTAILED IN TEACHING: MAKING PRACTICE A


CENTRAL ISSUE OF INQUIRY

This section is organized in five parts. First, we focus on issues around defining and
understanding practice(s), pointing both to the different ways in which this term has
been conceptualized, but also to the lack of shared language around defining this
concept. Next, we zoom in on scholarly work around defining and classifying core
or high leverage practices. Third, we discuss the work of decomposing teaching to
identify such practices. Fourth, we attend to empirical evidence generated about how
core or high leverage practices contribute to student learning. In the fifth part, we
consider how practices (in general) have been classified and measured/assessed. In
the final part, we examine how issues related to equity and culture are interwoven in
the discourse around teaching practices.

Defining “Practice”

The term “practice” has been widely used in recent mathematics education literature.
But the term has many meanings, little consistency exists in how it is used, and
explicit definitions rarely accompany the term. One might infer that the meaning is
implicitly clear or that it can be deduced from the context in which the term is used.
Or perhaps the definition is elusive because its meaning is specific to contexts and
“the meaning of a practice comes from its use [italics in original] in a community
and the value of that practice in that community” (Staples, Bartlo, & Thanheiser,
2012, p. 461). Given the amount of literature generated around the term and its
centrality to analyzing teaching and organizing teacher education, more consistency
in use of the term would likely help advance this body of research.
Lampert (2010) attempted to “provoke” clarity (p. 21) in the field by investigating
four conceptions of practice. One conception of practice is that it involves

359

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

implementing an idea in a context and is different to having the idea; this is the
commonly made distinction between theory and practice. A second conception of
practice is something that is done repeatedly in order to improve performance in
it, in the sense of how one might practice playing tennis or the piano. Third, the
practice of teaching includes people who have adopted “the identity of a teacher”
who have been “accepted as a teacher” and who have taken “on the common values,
language, and tools of teaching” (p. 29). This meaning is often used in the context of
a medicine or engineering practice. A fourth meaning of practice, which is typically
used in the plural form, practices, relates to routines that are done “constantly and
habitually” (p. 25) in the classroom.
Other scholars complement this fourth definition, by pointing out that practices
are performed by “taking into consideration teachers’ working context, and their
meanings and intentions” (Maryono, Sutawidjaja, Subanji, & Irawati, 2017,
p. 12) and that they require both professional judgment and the involvement of
“meaningful intellectual and social community for teachers, teacher educators, and
students” (McDonald, Kazemi, & Kavanagh, 2013, p. 378); we revisit these, themes
later in the chapter. An example of such a practice is orchestrating group discussions
(Hatch & Grossman, 2009). It is this fourth definition of practice that has become a
topic of research interest for researchers as part of a renewed interest in practice in
teacher education (Zeichner, 2012).
Notwithstanding attempts by Lampert and others to bring clarity to the field,
several examples exist of multiple meanings of the term, sometimes even within a
single sentence, like noting that “successful enactment of these practices are [sic]
typically found only in isolated pockets of practice” (Rosenquist, Henrick, & Smith,
2015, p. 43). Elsewhere, Lerman and Zehetmeier (2008) use the term in several
contexts: “researching practice,” “communities of practicing mathematics teachers,”
“reflective practice,” “how to organize a practice community,” “the time for teacher
talk and student practice went down,” “they often lack knowledge and practice
regarding these new issues,” and “the relationship of theory and practice” (see also
Lai, Auhl, & Hastings, 2015; Lloyd, 2013; Mayrowetz, 2009). Such examples show
how the term can be used as a noun, a verb or an adjective and the meanings can
include performance, to be actively engaged in a career, a way of learning (through
reflection or interaction), to perform repeatedly, experience, and apply.
In addition to multiple definitions of practice, a widespread reluctance to define
the term and lack of agreement among definitions of practice, the muddiness is
compounded by the use of other terms as quasi-synonyms for one or other meaning
of practice. Such terms include procedure (Maccini & Gagnon, 2006; Lloyd, 2013),
technique (Maher, 2008), strategy (Merritt, Palacios, Banse, Rimm-Kaufman, &
Leis, 2017), and instructional decision (Shechtman, Roschelle, Haertel, & Knudsen,
2010). The term “task” has been used both as a synonym for practice (Russ, Sherin,
& Sherin, 2015) and as a sub-component of a practice (Sleep, 2012).
The meaning of practice used will depend on an author’s theoretical perspective.
Our intention is not to impose a definition of practice but to advocate care in

360

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

articulating what meaning of the term applies and the theoretical perspective which
underlies it. In this chapter we define practice as regular and habitual classroom
routines engaged in by the teacher (Lampert, 2010, fourth definition) in a particular
community, requiring professional judgment (McDonald et al., 2013). Specifically,
we refer to practices known as “core” or “high leverage.”

Focusing on Core or High Leverage Practices

Defining practices as core or high leverage. One area on which researchers


have intensively focused is on identifying practices that are generative (Lampert,
2010), salient (Lampert et al., 2013), fundamental (Santagata, 2005), foundational
(Shaughnessy & Boerst, 2018), core (McDonald, Kazemi, & Kavanagh, 2013;
Shaughnessy & Boerst, 2018) or high leverage (Sleep & Boerst, 2012). The latter
two terms are widely used, often interchangeably even though nuanced differences
exist – “core” refers to teaching practices generally whereas “high leverage” focuses
on practices that are worthwhile for prospective teachers to learn.
Core practices have been defined as “identifiable components of teaching
that teachers enact to support learning. These components include instructional
strategies, and the subcomponents of routines and moves. Core practices include
general and content-specific practices” (Grossman, 2018, p. 184). General
examples include “implementing norms and routines for classroom discourse and
work” (p. 165) “eliciting and responding to student thinking” (p. 171), (p. 165)
and “providing feedback to students” (p. 179). Subject-specific practices are, for
example, “constructing and interpreting models” in science (p. 179) and “selecting
and adapting historical sources,” in history (p. 181). The definition, which was
developed by the Core Practice Consortium (a United States-based multi-institution,
multi-disciplinary group of teacher educators who aspire to achieve shared under-
standings around practices of teaching; https://www.corepracticeconsortium.com/)
and likely involved compromise and accommodation among group members,
provides little guidance for how a practice might be identified as core. For example,
would instructional strategies such as questioning, planning a lesson, or valuing
diverse voices and perspectives be included? Would they be classified as practices
or subcomponents of practices? Ambiguity exists about the relationship between
practices and subcomponents of practice, routines and moves.
More specificity is provided by McDonald et al. (2013) who cite the work of
Grossmann, Hammerness, and McDonald (2009) and identify criteria that core
practices might share:
they occur with high frequency, they permit beginning mastery by novices,
they can be enacted by novices across different curricula or instructional
approaches, they allow novices to learn more about students and about teaching,
they preserve the integrity and complexity of teaching, and they are research-
based and have the potential to improve student achievement. (p. 277)

361

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

High leverage practices are conceived with an explicit focus on supporting


prospective teachers in learning to teach:
[T]hose practices at the heart of the work of teaching that are most likely to
affect student learning … [they] comprise the essential activities of teaching;
if teachers are unable to discharge them competently, they are likely to face
significant problems. Competent enactment of such practices also lays the
foundation for beginning teachers to develop into highly effective professionals.
(Ball & Forzani, 2010/2011, p. 43)
Among sample practices identified are eliciting and responding to students’ ideas
(Lampert et al., 2013; McDonald et al., 2013; Shaughnessy & Boerst, 2018; Sleep
& Boerst, 2012), organizing a mathematical discussion (Kosko & Wilkins, 2015;
Tyminski, Zambak, Drake, & Land, 2014), and providing instructional explanations
(Kline & Ishii, 2008; Russ et al., 2016).

Classifying practices as core or high leverage. For practices to be classified as core


or high leverage, they must be evidence-based or research-based (e.g., McDonald et
al., 2013). Some studies attempted to link the implementation of specific practices to
student outcomes (see more on this below). For example, Webb and colleagues (2017)
found that teacher support of student participation influenced student participation
which, in turn, influenced student achievement. Lerkkanen and colleagues (2012)
compared how child-centered practices and teacher-directed practices affected
kindergarten children’s interest in reading. Greater interest was associated with
experiencing more child-centered practices and fewer teacher-directed practices.
Although practices observed in classrooms may be supported by research, that is
not universally true. Xenofontos (2016) found that some of the practices used by
teachers are not always optimal. One reported practice was lowering of expectations
for what immigrant students can achieve in mathematics class. The author criticizes
this, observing that the practice removes responsibility from the teacher.
Despite the definitions and criteria, and consensus on a handful of practices,
identifying definitive core or high leverage practices of teaching remains elusive.
The lack of consensus is evident in how practices are classified, how components of
practices are related, how lists of practices differ and in justifications for classifying
them as core or high leverage. Given that the criteria and definitions proposed above
are relatively recent, it is not unusual to witness variation in how the concept has been
interpreted by researchers working with different lenses, priorities, commitments,
and experiences. Nevertheless, identifying disagreements helps recognize where
potential for consensus exists and where differences are substantive. Resolving
such differences will help move the field towards Lortie’s aspiration of a “common
technical vocabulary” (1975, p. 73), a goal of the United States Core Practice
Consortium (Grossman, Kavanagh, & Dean, 2018).

362

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

Decomposing Practice(s)

Core and high leverage practices are typically identified by decomposing the work
of teaching. Decisions need to be made about the scope of a practice; if naming a
practice is to be helpful, it cannot be too big or too small. Subsequently, practices
themselves may be further decomposed either for teacher education purposes or to
map the terrain more precisely. We first review studies that attempted to decompose
practice(s) before making suggestions for advancing such work in future.
Sleep and Boerst (2012) envisage domains of teaching that contain practices,
and practices of teaching which contain techniques that can be “specified, taught
to, and worked on by beginners” (p. 1039). Although technique is not explicitly
defined, an example is “checking whether correct answers are supported by correct
reasoning” (p. 1039). “Assessing student thinking” is the example they give of one
domain and practices and techniques nested within that domain are shown in Figure
13.2. Even this relatively straightforward decomposition of a domain of teaching,
is decomposed in a way that includes “nested practices of varying grain sizes” (p.
1039), some of which are also known as subcomponents, and one technique.

Figure 13.2. How techniques are nested within practices of different grain sizes which are
nested within the domain of “Assessing Student Thinking” (based on Sleep & Boerst, 2012)

363

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

Elsewhere Shaughnessy and Boerst (2018) classify eliciting, responding, and


interpreting as three “interactive” (p. 42) practices whereas Teaching Works (see
http://www.teachingworks.org/) combines two of them – eliciting and interpreting
individual students’ thinking – into one high leverage practice (Grossman, 2018).
Shaughnessy and Boerst (2018) focus on the “foundational” practice of “eliciting
student thinking” and they identified four core components of the practice that they
expected prospective teachers to demonstrate: “(a) eliciting the student’s process, (b)
probing the student’s understanding of key mathematical ideas, (c) attending to the
student’s ideas, and (d) deploying other moves that support learning about student
thinking” (p. 45). Each component is associated with moves that would make visible
or not that component of performance. Moves are “specific steps of talk that teachers
take as they interact with students … or particular actions” (p. 45).
Elsewhere Sleep (2012) refers to tasks as subcomponents of “steering instruction
toward the mathematical point.” Although she classifies “steering instruction toward
the mathematical point” as part of the “work of teaching” rather than a practice of
teaching, her goal in the research is to decompose this aspect of the work of teaching
and to identify subcomponents of it.
Similarly when Tyminski and colleagues (2014) decompose the practice of
organizing a mathematical discussion, they draw on Smith and Stein’s work (2011)
as “one way to accomplish the decomposition of a complex practice” (p. 468) into
the five practices of anticipating, monitoring, selecting, sequencing, and connecting.
Tyminski et al. imply that practices are nested within other practices. Santagata and
Yeh (2014) are also interested in instructional conversations, but they focus on how
student thinking is made visible. Their decomposition of the practice is presented
at three levels of sophistication: low (student thinking is minimally visible or not
visible), medium (student thinking is visible) and high (student thinking is both
made visible and probed further).
Briefly surveying how practices are decomposed illustrates the lack of
consistency around language, definition, detail, and grain size. Comparing lists of
practices articulated by different institutions and collaborators (see appendices in
Grossman, 2018) confirms this finding. Such differences and lack of consistency
may be symptomatic of a field in exploratory mode regarding practices of teaching
and practice-based approaches to teacher education. Indeed, if consensus is to be
achieved, it will likely take time. Yet, given the green shoots of research currently
visible, researchers in the field could strive for more consistency when studying
similar practice-related ideas. Little or slow progress seems likely unless the fruits
of collaboration among practice-focused research groups in the United States
(Grossman et al., 2018), become more evident in that country initially, because
much of the impetus for this work is sourced there. Building a common technical
vocabulary among researchers and teacher educators across and beyond the United
States poses additional challenges.
Although decomposition is necessary for understanding the work of teaching and
making it teachable and learnable, focus must remain on the bigger picture and the

364

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

complexity of teaching. Therefore, we encourage more researchers to attend to both


aspects – decomposition and recomposition of practices (Sleep, 2012) –showing
how prospective and practicing teachers can be helped to study and learn individual
practices, and the practice of teaching in general.

Empirical Evidence on the Contribution of High Leverage or Core Practices to


Student Learning

Most definitions of core or high leverage practices aspire to improve student learning.
Ball and Forzani (2009) claim that “proficient enactment of [such practices] by a
teacher is likely to lead to comparatively large advances in student learning” (p.
460). The Core Practice Consortium identifies practices that constitute “strategies,
routines, and moves that can be unpacked and learned by teachers to support student
learning” (cited in Grossman et al., 2018, p. 4), while Grossman, Hammerness, and
McDonald (2009) argued that these practices are research-based and can improve
student achievement. We now consider empirical evidence that these practices
support student learning, purportedly through improving teaching quality.
Defining practices remains a problem in reviewing such evidence. First,
studies that explore how practices of teaching contribute to student learning do
not necessarily identify these practices as such – let alone call them core or high
leverage; instead they refer more generally to different instructional aspects or
teaching factors. Second, several other studies implicitly or more explicitly cluster
individual strategies or teacher techniques under the term “teaching practice(s)”
without, however, having the features that teaching practices, as defined above, are
expected to have. Acknowledging such complications, the brief literature review
that follows aims to highlight certain topics instead of providing a comprehensive
account of what has been accomplished in the field.
Interestingly, few studies provide empirical evidence supporting the contribution
of these practices to student learning. Most studies that provide such evidence
were published during the last decade and pertain to kindergarten and primary-
school grades (e.g., Bargagliotti, Gottfried, & Guarino, 2017; Blazar, 2015; Bottia,
Moller, Michelson, & Stearns, 2014; Cohen, 2018; Firmender, Gavin, & McCoach,
2014; Ing et al., 2015); much rarer are studies focusing on secondary grades (e.g.,
Charalambous & Kyriakides, 2017; Fyfe & Rittle-Jonhson, 2017). Scholars seem
to have followed different data-collection approaches when examining teaching
practices, ranging from live or videotaped classroom observations (e.g., Blazar, 2015;
Cohen, 2018 Firmender et al., 2014) to teacher self-reports (e.g., Bargagliotti et al.,
2017; Bottia et al., 2014), and student surveys (e.g., Charalambous & Kyriakides,
2017). Although pursuing a suite of methodological approaches in studying this
effect is both desirable and necessary, given the complexity of teaching, scholars
are rarely explicit about the affordances and limitations of their chosen evaluative
approach(es) and the decisions that guided their selections. Explicitness around
such matters is critical for making more compelling arguments about incorporating

365

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

the practices in initial and practising teacher education (see similar arguments in
Praetorius & Charalambous, 2018).
Although some negative or non-significant effects have been identified (e.g.,
Bargalliotti et al., 2017), most studies reviewed provide encouraging results
suggesting that different practices such as teacher modeling (Cohen, 2018), providing
feedback (Fyfe & Rittle-Johnson, 2017), questioning and structuring (Charalambous
& Kyriakides, 2017), practices encouraging productive dialogue (Webb et al., 2017)
or immersing students in rich and cognitively challenging learning environments
(Blazar, 2015) are positively associated with student achievement or learning. The
promise of combining teaching practices that cut across different subject matters –
often called generic practices – with mathematics-specific practices in explaining
student learning has also been empirically suggested (Charalambous & Kyriakides,
2017). These authors drew on data from the Trends in International Mathematics and
Science Study (TIMSS) 2011 and 2015 cycles and showed that a higher percentage
of the unexplained variance in student learning could be explained when combining
practices as opposed to when considering generic or content-specific practices in
isolation. Interestingly, this percentage seemed to vary across different countries,
thus calling for a deeper exploration of the mediating role that different contextual
factors might have on the association. Pointing to a missing link between teaching
practices and student learning, another recent study (Ing et al., 2015) examined the
role of student participation in mediating the relationship between teacher support
and student learning. The authors advocate attending to both teaching and student
participation to understand how teaching can affect student learning.
Despite such work on studying associations of core or high leverage practices
with student learning, significant outstanding matters require consideration. First, in
line with other scholars (Ball & Forzani, 2009; Hlas & Hlas, 2012), more concerted
efforts are needed to empirically validate the contribution of these practices to student
learning. The evidence generated to date seems thin and unsystematic. This could
be addressed by developing a common lexicon and framework for describing and
analyzing instruction (cf. Grossman & McDonald, 2008) and seeking consensus on
a set of core practices considered important for teaching in general and for teaching
mathematics in particular. Achieving such a consensus would be a substantial
achievement.
Should such a consensus be achieved, it would create the possibility of explicitly
investigating how these teaching practices influence student learning. This is
particularly important because without clearly articulating specific hypotheses
around the mechanisms through which teaching practices can affect student
learning, developing and implementing research designs that lend themselves to
capturing these effects seems difficult. Along with Hlas and Hlas (2012), we believe
that the scholarly community could agree on what constitutes empirical evidence in
investigating the aforementioned association and on what methodological designs
might better lend themselves to generating such evidence. Although classroom
observations are often viewed as the gold standard in exploring instructional

366

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

quality (cf. Douglas, 2009), multiple approaches may more effectively capture how
instruction contributes to student learning. These methodological considerations also
need to be informed by a broad conception of student learning that moves beyond
the cognitive outcomes to also include other forms of learning, such as affective and
meta-cognitive outcomes (cf. Reynolds et al., 2016).
Ing and colleagues (2015) point to another critical issue. By overemphasizing the
role of teaching practices and the opportunities teachers craft for student learning,
as a research community we run the risk of missing a crucial link in the chain
connecting teaching and student outcomes: how the students themselves respond
to opportunities crafted for their learning. In fact, a recent synthesis of different
classroom observation instruments (Praetorius & Charalambous, 2018) revealed that
what teachers do or how they interact with their students is valued over closely
attending to students’ participation and the extent to which students avail of
opportunities to learn.
Finally, different contextual factors (e.g., available curriculum resources,
instructional approaches followed, school-level or system-level factors) can have a
mediating role on the association between teaching practices and student learning.
One characteristic identified by Grossman et al. (2009) is that these practices can be
enacted in classrooms across different curricula or instructional approaches. However,
the impact that such curricula or instructional approaches can have on the effect of
these teaching practices remains an open issue. Coupled with Charalambous and
Kyriakides’ (2017) findings suggesting differences across countries and educational
systems on the effect of generic and content-specific practices on student learning,
such explorations seem imperative.
Our review has so far been focused on core and high leverage practices. Because
many studies focus on practices of teaching without identifying them as such, we
now consider these studies briefly, while discussing issues of measuring/exploring
them.

Classifying and Measuring “Practices” as Other Than Core or High leverage

How practices are classified. In addition to core and high leverage practices,
authors categorize practices of teaching in multiple and sometimes nuanced ways.
Several researchers refer to the idea of best practices, either on the basis of research
warrants (e.g., Eddy, Converse, & Wenderoth, 2015) or based on teachers’ selection
of artefacts to represent their teaching at its best (Silver, 2010). Lampert (2010) sees
a problem with borrowing this term from the business world because reference to
“best practice” raises the question of what goal the practice is best for achieving and
what evidence supports a practice’s designation as “best.”
Frequently, practices are designated as “instructional” (e.g., Lee, Walkowiak, &
Nietfeld, 2017; Silver, 2010; Swars, Smith, Smith, Carothers, & Myers, 2018).
Lloyd (2013) defines instructional practices as “those that necessitate critical
thinking, reasoning, high levels of abstraction, and problem solving” (p. 107) and

367

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

she distinguishes them from social/emotional and management practices. Coming


with an interest in teaching mathematics to students with learning disabilities,
Maccini and Gagnon (2006) define instructional practices as being “both empirically
validated and recommended practices for teaching math” to students with learning
difficulties or emotional and behavioral difficulties (p. 218) and the requirement for
empirical validation is echoed in many articles.
Many authors contrast types of practice, like reform-based and traditional (e.g.,
McCaffrey et al., 2001; McClintock, O’Brien, & Jiang, 2005; Suurtamm, Koch, &
Arden, 2010), teacher-directed and student-centered (Morgan, Farcas, & MacZuga,
2015), and high press and low press (a term taken from Kazemi & Stipek, 2001
by Webb et al., 2009). Culturally responsive practices have also been identified
(e.g., Ukpokodu, 2011). Culturally responsive practices are derived from culturally
relevant or culturally responsive pedagogy and propose that all students experience
academic success, develop or maintain cultural competence, and develop critical
consciousness through which the current social order can be challenged (Ladson-
Billings, 1995). Cultural practices will be discussed further below.
A more specific kind of practice, responsive classroom, is studied by Ottmar,
Rimm-Kaufman, Larsen, and Berry (2015). Responsive classroom refers to “a set
of principles and practices for integrating social and academic learning across the
school day, creating classroom management processes well aligned with children’s
social and emotional needs, and fostering a caring and responsive environment for
students” (p. 792).
Other researchers avoid using the term practice. Webb and colleagues (2017)
use the term (teacher) “move” or “intervention” where others might use “practice.”
Examples include “help students formulate their own ideas and consider others’
perspectives” (p. 3), “asking students probing and clarifying questions” (p. 3) and
“‘positioning’ students as capable participants (p. 7).

Measuring and assessing practices of teaching. Here we focus only on the


(classroom observation) instruments that refer explicitly to measuring teaching
practices. Over the past two decades several other instruments have been developed to
measure certain instructional aspects, without, however, identifying them as teaching
practices. For a review of these instruments see Praetorius and Charalambous (2008).
Several instruments have been used by researchers to study or identify practices
of teaching, each with their own conception of practices of teaching. The instruments
include the Standards-based Learning Environment Protocol developed by Tarr
et al. (2008) to observe students, teachers, lessons and their interactions (used by
Swars et al., 2018); the Classroom Observation and Analytic Protocol developed
by Horizon Research (2000) which is an observational tool in which the observer
judges the lesson according to design, implementation, content and culture before
assessing the lesson’s likely impact on student learning; the Classroom Practices
Observational Measure (CPOM) developed by Abry, Brewer, Nathanson, Sawyer,
and Rimm-Kaufman (2010); and the Early Childhood Classroom Observation

368

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

Measure –ECCOM (see Lerkkanen et al., 2012). In this instrument, observers rate
management, climate and instruction in classrooms under two broad headings of
child-centered teaching practices and teacher-directed teaching practices.
Articles that document measures of teaching practices are typically explicit in
identifying the practices, even if the practices sometimes consist of a single word
and lack the detail or nuance contained in lists of core or high leverage practices.
The Teaching Practices Inventory was designed for use on university mathematics
and science courses and it measures the extent of practices rather than the quality
of their implementation, something that is more difficult to measure (Wieman &
Gilbert, 2014). It identifies two categories of practice: Practices that support learning
(such as knowledge organization, long-term memory and reducing cognitive load,
motivation, practice, feedback, metacognition, and group learning) and practices that
support teacher effectiveness (such as connecting with students’ prior knowledge
and beliefs, feedback to instructors on their effectiveness, such as mid-course
evaluations or repeated feedback from students, and gaining relevant knowledge and
skills). Another instrument for studying teaching practices in large university classes
is the Practical Observation Rubric to Assess Active Learning –PORTAAL (Eddy et
al., 2015). It measures teaching along the dimensions of practice, logic development,
accountability and reducing apprehension.
Shaughnessy and Boerst (2018) documented an innovative approach to assessing
practices. Faced with the challenge of fairly and consistently assessing the practice
of prospective teachers when real children responded in various ways, the authors
adapted an idea from medicine, the “standardized patient” and developed the idea of
the “standardized student” When assessing the practice “eliciting student thinking”
prospective teachers interacted with an adult who was scripted to respond to tasks
in a way many students might respond, thus presenting consistent challenges for
the teachers who were being assessed. Such an approach, though time-consuming
was innovative; each prospective teacher received a similar challenge in a setting
where they were required to act as teachers rather than write about teaching. Such
approaches suggest that in the future, classroom observations may be complemented
by other innovative ways to better capture teachers’ capacity to implement such
teaching practices.
We close this section by discussing attempts to interweave the exploration of
teaching practices with issues of equity and culture. Albeit short, this discussion
is deemed necessary since it documents scholarly awareness about the fact that
teaching needs to serve students with diverse needs and capabilities and that teaching
constitutes a cultural activity (Stigler & Hiebert, 1999).

Interweaving Issues of Equity and Culture in Exploring Teaching Practices

Practices to promote equity. Students may be considered marginalized in schools


for many reasons (e.g., bilingual students, students of color, students with learning
disabilities). Hence Gutiérrez (2002) provides a definition of equity practice. She

369

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

proposes a reorientation for how practices should be viewed in relation to equity


concerns, and she specifies questions to be addressed by researchers in this area. In
particular, equity practice refers
to the practice enacted between teachers, students, and mathematics that
empowers students to (a) develop proficiency in dominant mathematics, (b)
develop critical stances and new perspectives on the relationship between
mathematics and society, and (c) contribute toward a positive relationship
between mathematics, people and society in ways that erase inequities on this
planet. (p. 174)
Although this conception of equity practice expects students to become competent
in dominant mathematics, it also envisages students adopting a critical stance
towards what has been called an understanding of mathematics as “an exclusively
European product” (Dowling, 1998, p. 3). In this expression Dowling is referring
to a eurocentric, elitist view of mathematics and mathematics education that is
promulgated by school mathematics and school mathematical texts in particular,
despite the contention in mathematics education literature that mathematics features
centrally in all human cultures.
In a study of two classrooms where English language learners made high
achievement gains, Merritt and her colleagues (2017) found that the teachers used
multiple representations of concepts, they emphasized the building of mathematical
vocabulary, both checked frequently for understanding and spent time analyzing
students’ errors. In both classrooms, students spent relatively little time in small
group discussions and the researchers were surprised that neither teacher connected
the mathematics to students’ everyday lives. Perhaps, they hypothesize, teachers
were focusing on preparing students to do decontextualized tasks similar to those
they will encounter on high-stakes tests organized by the district.
These findings largely echo those of an earlier, similarly small study of three
teachers where the more successful teachers with high numbers of Latino students
were those who moved through the lesson more slowly, alternating whole class
time with individual/small group time, who built on student contributions when
introducing new mathematical vocabulary, and who responded explicitly to errors
(Zahner, Velazquez, Moschkovich, Vahey, & Lara-Meloy, 2012).
In an essay about supporting the learning of mathematics by English language
learners (i.e., students whose first language is not English and includes those both
with early and considerable proficiency in the language; Lacelle-Peterson & Rivera,
1994), Moschkovich (2013) recommends four practices all based around language.
Instead of focusing on accurate language or single words, teachers should focus on
mathematical reasoning and mathematical practices; teachers should support students
in engaging in complex mathematical language; and students’ home languages
should be seen as resources and not obstacles to their progress in mathematics.
Jackson and Wilson (2012) reviewed mathematics education research from 1989
to 2011, to identify instructional practices that are particularly conducive to African

370

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

American students’ learning. A key finding of their work was that current research
supports only the identification of broad principles for teaching these students and
the authors advocate continued investigation of practices of teaching that, with other
factors, may enhance student learning.
Several authors make recommendations about how students with learning
disabilities can be supported in learning mathematics (e.g., Maccini & Gagnon,
2006; Mayrowetz, 2009; Spooner, Saunders, Root, & Brosh, 2017). Spooner et al.
look at an evidence-based practice for teaching mathematical problem solving to
students with learning disabilities called schema-based instruction. This practice
has four components: make the problem accessible (e.g., through interactive read-
alouds of story problems or meaningful and motivating contexts); make the problem
conceptually comprehensive by providing graphic organizers and by sequencing
problems from easier to more difficult types; solve the problem procedurally; and
generalize in several ways. Mayrowetz (2009) looked specifically at the treatment
of tasks in inclusive classrooms with high incidence of students with special needs.
Although teachers helped students on a one-to-one basis, it was rarer for teachers to
modify their instruction for students with disabilities. Documenting how practices
are inclusive of all learners will greatly enhance research and practice in this area.

Practices of teaching as cultural artefacts. Much has been written about the
cultural nature of teaching (e.g., Stigler & Hiebert, 1999; Stylianides & Delaney,
2011) and the impact of culture is evident in the literature about practice. This
phenomenon refers to both differences in teaching across cultures and to how
mathematics teaching is largely influenced by the culture in which it is applied. We
elaborate further on each of these in relation to the articles reviewed.
Santagata (2005) compares the mistake management sequence in Italy and the
United States and finds that whereas United States teachers tend to move on from
mistakes quickly, Italian teachers more typically asked students to correct their errors,
often with support from the teacher. Although Santagata does not call the ‘mistake
management sequence’ a practice, it might be deemed the product of a practice in
countries where it is used or it may require the naming and identification of related,
nuanced practices in line with priorities in a given country. Lan and colleagues
(2009) studied classroom practices in China and the United States. They found that
practices associated with higher performance in China (e.g., large-group instruction)
are not inherently effective but that other factors are related to effective practice such
as student behavior and how lessons are enacted. “Whole-class instruction” is not
necessarily a practice in itself, and is certainly not a reliable term across cultures,
because it is enacted differently across countries: promoting active engagement in
China versus primarily lecture time in the United States.
Wager (2012) identified four practices in which teachers engaged to help students
in schools with large numbers of ethnic minority students connect their everyday
lives to school mathematics. First was using students’ out of school experiences
as contexts for word problems. Second was relating cultural activities to school

371

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

mathematics (e.g., mathematics of nutrition; fractions and cooking). Third was


encouraging students to use informal strategies – often different to those used in
school – to solve problems such as purchasing food for the family. Fourth, the
classroom was used as a site of culture (e.g., modelling a daily store, studying
butterfly migration to Mexico and money from around the world).
When identifying practices of mathematics teaching, although overlap likely
exists in practices from one country to another, some practices will be country- or
culture-specific. Furthermore, practices that acknowledge and value the diverse
backgrounds of all students as a foundation for fostering their mathematical
achievement must be included when practices are identified.
Given the current widespread interest in practices of teaching, it may be timely
for the field to strive to agree on a conception of what a practice of teaching is. This
could lead to mapping teaching through its practices, the description of categories of
practice – such as ‘core’ and ‘high leverage’ – and their relationship with domains,
tasks, techniques, and subcomponents of practice. The extent to which practices are
general or subject specific could be explored, as could their grain size. Means of
evaluating the impact of using classroom practices on student learning could also
be identified. Practices need to be articulated in a way that include students with
disabilities and members of ethnic or linguistic minorities. Some practices are likely
to be specific to or different in particular countries or cultures. In order to avoid
practices being viewed as thoughtless routines, the exercise of teacher judgment
in practices needs to be studied and highlighted in response to how McDonald et
al. (2013) include this as a criterion. Similarly, the community-basis for adopting,
implementing, and evaluating practices merits further study.

PRACTICE-BASED PEDAGOGIES: MAKING PRACTICE A CENTRAL SOURCE


OF TEACHER LEARNING

In this section, we focus on pedagogies that have the potential to foster teacher
learning and which put teaching practice at the core. We first explain what is meant
by practice-based pedagogies and we justify our decision to focus on representations
decomposition, and approximations of practice, as an overarching umbrella covering
such practice-based pedagogies. We then provide specific examples of implementing
these pedagogies and synthesize literature that explores their effectiveness in
promoting teacher learning. We conclude by identifying open issues that warrant
consideration as research on practice-based pedagogies accumulates.

Practice-Based Pedagogies: Focusing on Representations, Decomposition, and


Approximations of Practice

The terms practice-based pedagogy or practice-based teacher education have


been used in different ways to capture different approaches in teacher education
(cf. Forzani, 2014). Despite the differences in use, the term mostly emphasizes

372

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

the importance of engaging prospective and practicing teachers not in theoretical


discussions around practice, but in enacting practice as a means of learning to teach.
Realizing the limitations of traditional approaches in helping prospective teachers
and novice practicing teachers learn to do (rather than to think about) the complex
work of teaching (Forzani, 2014; Grossman et al., 2018; Grossman & McDonald,
2008; McDonald et al., 2013), the field of teacher education has recently witnessed a
shift, turning away from an intense focus on the knowledge needed for teaching to the
use of this knowledge in practice. As Grossman and colleagues (2018) suggest, this
increased emphasis on practice was preceded by similar attempts in the first half of
the previous century (e.g., the Commonwealth Teacher Training Study, see Forzani,
2014) and later in the 1960s and 1970s (competency-based teacher education), which
led to long lists of discrete skills that prospective teachers and novice practicing
teachers were expected to learn and practice. This reductive conception of teaching
shifted the pendulum to the end of teacher knowledge and judgment, from which the
field seems to gradually be departing (cf. Ball & Forzani, 2009; Gitomer & Zisk,
2015). Examples of such shifts can be seen in the restructuring of the teacher education
programs, in the United States (see, for example, Ball, Sleep, Boerst, & Bass, 2009;
Grossman et al., 2018), but also outside the United States [see, for example, Jao,
Wiseman, Kobiela, Gonsalves, and Savard (2018) for analogous efforts in Canada].
Comparing professional pedagogies across three professions, Grossmann et
al. (2009) developed a framework for the teaching of practice that includes three
elements: representations, decomposition, and approximations of practice. We now
explain each component and then explicate why we regard this framework as useful
in capturing different practice-based pedagogies.
Representations of practice are how practices of teaching are made visible to
prospective, novice or experienced teachers. Possible representations include records
of practice like videos of teaching, observations of live teaching, public lessons,
modeling teaching practices, classroom transcripts, lesson plans, student artifacts,
case studies about teaching, teacher narratives about teaching, and multimedia and
animated portrayals of teaching (Ball, Ben-Peretz, & Cohen, 2014; Danielson,
Shaughnessy, & Jay, 2018; Grossman et al., 2009; Han & Paine, 2010; Herbst &
Kosko, 2014; McGrew, Aston, & Fogo, 2018). Different representations illuminate
certain aspects of teaching, while obscuring others. Therefore, teacher educators
must carefully select representations to use, based on the purpose they are expected
to serve (cf. Lampert & Ball, 1998).
Decomposition of practice refers to partitioning the complex work of teaching
and its practices into identifiable, constituent parts. Teaching can be decomposed
to identify core or high leverage practices as detailed above and decomposition
can be performed at varying grain sizes. Decomposing practice allows for targeted
instruction and scaffolding to help teachers learn and improve aspects of teaching.
As Grossmann et al. (2009) note, “By decomposing complex practices, professional
educators can help [prospective teachers] learn first to attend to, and then to enact,
the essential elements of practice” (p. 2069). In essence, decomposition is useful

373

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

for developing professional vision (cf. Goodwin, 1994), and for enacting practice
in environments that support experimentation with and reflection upon different
components of teaching.
Approximations of practice, the third pillar of the framework, refer to creating
safe environments for prospective or novice teachers to practice teaching with
high degrees of support. By reducing complexity and focusing teachers’ attention
on specific aspects of teaching, approximations create productive spaces for
experimentation and deliberate and substantive reflection on teaching. Equally
important, they “reduce the error in the field,” while still helping teachers to focus on
“high-stakes practices” (Grossman et al., 2009, p. 2091). Approximations are used
widely in other professions like medicine, dentistry, law, and pilot training (Schutz,
Grossman, & Shaughnessy, 2018). In teacher education they can take different
forms, including role plays (Schutz et al., 2018), replays (Horn, 2010), rehearsals
(Ghousseini, 2017; Horn, 2010; Kelley-Petersen, Davis, Ghousseini, Kloser, &
Monte-Sano, 2018), microteaching (Cheng, 2017; Hong & Chai, 2017; Lai et al.,
2015), simulations of practice (Charalambous, 2008), field placement teaching,
fishbowls, co-teaching, and processing pauses (Schulz et al., 2018). These forms
differ in their level of authenticity.
Admittedly this framework cannot encompass all different forms of practice-based
pedagogies. For example, although it does create a space for teacher collaboration
and learning, it does not necessarily capture the complexity of such collaborations, as
manifested in communities of practice (e.g., Jaworski, 2006; Sowder, 2007). However,
it can serve as an umbrella for different other professional development approaches,
some of which have become particularly popular in the last two decades, such as
lesson study (see Huang & Shimizu, 2016 for a systematic review) and video viewing
in general (see Gaudin & Chaliès, 2015 for a review) or video clubs, in particular
(e.g., Charalambous, Philippou, & Olympiou, 2018; Russ et al., 2015; Sun & van Es,
2015). For example, lesson study can be considered an authentic approximation of
teaching, since it occurs in actual classrooms and it preserves the complexity of the
work; at the same time, the videotaped or live classroom observations that take part in
the context of lesson study provide representations of the work of teaching, which can
then be decomposed according to the foci set in each lesson-study session.
Hence, although Grossman et al. (2009) framework cannot comprehensively
capture existing work on practice-based pedagogies, it offers a good heuristic for
capturing three components of such pedagogies that are present to a greater or lesser
extent in pedagogies that render practice and its enactment at the core of teacher
professional development.

Examples of Incorporating Practice-Based Pedagogies in Teacher Education


Programs

Over the last decade teacher education programs have introduced or reported on
practice-based pedagogies. Some efforts are programmatic, involving restructuring

374

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

an entire teacher education program; others involve revamping individual courses


(cf. Cantun, Schutz, Kelley-Petersen, & Franke, 2018). Common to all such efforts
is the identification of core or high leverage practices, which are represented in
various ways, decomposed systematically, and approximated in controlled settings.
A programmatic level example is the systematic redesigning of the undergraduate
program at the University of Michigan to center it more on practice (see Teaching
Works, http://www.teachingworks.org). Faculty from the University have
collaborated with practicing teachers and graduate students to identify a set of high
leverage practices that cut across different subjects. Following these year-long
efforts, 19 such practices were generated, including leading a group discussion;
eliciting and interpreting student thinking; specifying and reinforcing productive
student behavior; and setting long and short-term learning goals for students (see
Grossman, 2018, pp. 164–169 for a complete list). Following this decomposition of
teaching, scholars designed courses to immerse prospective teachers in analyzing
representations of teaching (as exemplified below) and in enacting certain elements
of these practices in a heavily scaffolded environment. A similar endeavor was
undertaken under the University of Washington Accelerated Certification for
Teachers (U-ACT) initiative, which resulted in restructuring two graduate-level
teacher education programs around seven core teaching practices, including orienting
students to each other ideas, orienting students to the content, and assessing student
understanding (see Grossman, 2018, pp. 170–173 for a list of these practices).
At course level, several initiatives have been undertaken by individual or small
groups of faculty members, often within the same disciplinary area (for such examples
see Cartun et al., 2018, pp. 109–133). Our review yielded several examples of how
teacher educators have capitalized on the idea of representations, decomposition,
and approximations of practice to restructure their teaching programs (e.g., Ball et
al., 2009; Boerst, Sleep, Ball, & Bass, 2011; Erickson & Herbst, 2018; Ghousseini,
2015; Ghousseini & Herbst, 2016; Ghousseini, 2017; Herbst, Chieu, & Rougée,
2014; Horn, 2010; Shaughnessy & Boerst, 2018; Tyminski et al., 2014). We now
review two such examples which illustrate both the efforts invested in considering
and structuring such environments, and the several decisions involved in capitalizing
on this framework to inform the work. We highlight these examples because they
explain clearly how Grossman et al.’s (2009) framework was used, something that
seems to be the exception rather than the rule.
The first example pertains to a mathematics methods course for prospective
elementary schoolteachers designed at the University of Michigan (Boerst et al.,
2011). This course was revamped to emphasize four teaching practices: leading
a classroom discussion; planning mathematics lessons; assessing students’
knowledge, skill, and dispositions; and representing mathematical ideas. In the
study the authors focus primarily on leading a classroom discussion and provide
a detailed account of how they decomposed this practice into five identifiable
components. This decomposition enabled the teacher educators to design various
opportunities to engage prospective teachers in gradually learning this practice

375

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

through approximations of practice. Specifically, prospective teachers were initially


supported to engage in asking purposeful questions; then, to teach a “mini-problem”
during field placement, and finally, to teach an entire mathematics lesson. This
nesting of earlier approximations into subsequent ones and the gradual increase in
complexity and authenticity allowed prospective teachers to progressively immerse
themselves in the focal practice. Prospective teachers’ work was scaffolded by
representations of practice (e.g., videos of practice). In viewing these representations,
prospective teachers were provided with viewing lenses that enabled them to focus
their attention on specific aspects of teaching, starting narrowly, then expanding the
lens gradually to include more aspects of the relevant practice.
The second example pertains to a mathematics methods course intended for
prospective secondary school teachers (Ghousseini & Herbst, 2016). Although
again focusing on leading a classroom discussion, the teacher educators in this study
followed a different approach to that described above, thus illustrating another way
in which teacher educators have applied Grossman et al.’s (2009) framework of
practice-based pedagogies. The teacher educators began by representing a classroom
mathematics discussion in which the prospective teachers worked as students on
a warm-up problem and then discussed it as a whole class. One teacher educator
deliberately modeled some of the work involved in leading a discussion by enacting
instructional moves he wanted the prospective teachers to learn. This created a
productive space to decompose work entailed in leading a classroom discussion,
something that was done in collaboration with the prospective teachers themselves,
who were asked to collectively label and elaborate aspects of this work. Like in
the previous study, this was followed by different types of approximations of
practice. First, the prospective teachers were given a constructed dialogue between
an imaginary teacher and her students around a mathematical problem and were
asked to complete the teacher lines in this dialogue which had been purposefully
erased. Next the prospective teachers rehearsed leading a classroom discussion in
a fishbowl setting. In the third approximation, prospective teachers planned with
their cooperating teachers to lead a classroom mathematics discussion in their field
placement.
These two examples illustrate the potential of such courses to support prospective
teachers in learning the work of teaching. They also typify the careful work needed
in developing such courses, given that different decisions need to be made at several
junctures, both when planning and when enacting and reflecting upon such efforts.
In what follows, we review the literature with respect to evidence accumulated so far
about the effectiveness of such practice-based pedagogies.

Empirical Evidence on the Contribution of Practice-Based Pedagogies

In this part, we focus on studies that provide empirical evidence about the contribution
of practice-based pedagogies. Literature that discusses these issues theoretically or
documents the design of programs/courses (e.g., Ball et al., 2009; Boerst et al., 2011;

376

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

Grossman et al., 2009) were not considered. In addition to considering the evidence
generated in these studies, we discuss three other matters that help contextualize this
evidence: the opportunities afforded to teachers for enacting and reflecting upon
their (own) practice; the teaching practices considered; the participants engaged
in these studies; and the evidence generated with respect to (prospective) teacher
learning.
Our review pointed to a notable variation in what is reported regarding the
contribution of practice-based pedagogies to teacher learning. Some studies (e.g.,
Han & Paine, 2010; Tyminski et al., 2014) go into detail either illustrating how
teachers’ work changed as a result of participating in the practice-based learning
environments or discussing the mechanisms that seemed to facilitate participants’
learning (e.g., Horn, 2010). Other studies (e.g., Averill et al., 2016; Ghousseini,
2017), however, either present prospective teachers’ learning in broad strokes or
largely remain silent on this issue. In general, no common methodological framework
seems to exist for guiding scholarly explorations about what counts as evidence
on the contribution of these environments to teacher learning, let alone how this
evidence needs to be generated and reported. These issues represent open areas that
ought to be considered in the next decade, especially since work around practice-
based pedagogies seems likely to grow. Scholarly attempts need to produce stronger
and more systematic validity evidence about the potential of these practice-based
approaches in supporting (prospective) teacher learning.
Given the emphasis on practice-based pedagogies, it is unsurprising that many
studies report extensively on environments that were crafted to support prospective
teachers’ opportunities to enact practices of teaching (e.g., Averill, Drake, Anderson,
& Anthony, 2016; Ghousseini, 2015, 2017; Schutz et al., 2018; Tyminski et al., 2014).
Such reporting is reasonable, given the new (or renewed) emphasis on practice-
based pedagogies in the past decade and scholars drawing on it have attempted to
explicitly describe how they capitalize on its affordances to support teacher learning.
A critical component in this reporting is the opportunities afforded to prospective
teachers to reflect upon their own teaching or that of their peers. In rehearsals, the
key approximation of practice for which empirical evidence has been generated,
such opportunities are made available through debriefs often accompanying the
approximation, either with the prospective teacher engaged in the rehearsal or with
observers (see, for example, Averill et al., 2016; Ghousseini, 2015; Schutz et al.,
2018). Studies that mostly focused on representations of practice (e.g., Herbst &
Kosko, 2014; Herbst et al., 2014) report on how these records of practice facilitated
eliciting different teacher ideas and beliefs about teaching. Despite their significance,
these latter studies provide a window into teachers’ thinking rather than their actual
experimentation with the work of teaching.
Another pattern emerging from this analysis concerns the teaching practices that
have been the focus of recent scholarly efforts. Interestingly, most of the works that
provide empirical evidence on practice-based pedagogies focus on just a single
practice – something not surprising given the publishing space restrictions. What

377

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

is interesting, however, is that most of these works focus on the practice of leading
a classroom discussion (e.g., Averill et al., 2016; Ghousseini, 2015; Ghousseini &
Herbst, 2016; Tyminski et al., 2014). The focus on this practice can be attributed partly
to the fact that it represents a key component in several lists of core or high leverage
practices and partly because it has been significantly decomposed and elaborated
on in the literature (see, for example, Stein, Engle, Smith, & Hughes, 2008; Smith
& Stein, 2011); it is also considered core in attempts to offer students high-quality
instruction that elicits and builds upon their thinking (cf. Jacobs & Spangler, 2017).
The focus on leading classroom discussions implies that different manifestations
of practice-based environments around this practice have been generated, which
enables their comparability. At the same time, it represents a significant open issue
for years to come, because examples of other teaching practices are needed, to
illustrate how practice-based pedagogies can be implemented and to support teacher
learning in other teaching practices.
Unsurprisingly, the participants engaged in these studies are exclusively
prospective teachers. The number of the participants reported in the studies,
however, varies. In some studies (e.g., Ghousseini, 2015) only a single teacher is
considered; in other cases (e.g., Averill et al., 2016; Ghousseini, 2017; Tyminski et
al., 2014) many more participants are presented. Each approach has its affordances
and limitations: single case studies provide the opportunity to document in more
detail how the prospective teacher interacted with the practice-based learning
environment; what is left unattended, however, is whether and how this environment
actually functioned for the rest of the participants. Studies that report on a large
number of teachers unavoidably report on the average, thus leaving the differential
effect of such environments on their participants open to further inquiry. Therefore,
future work needs to strike a balance between foregrounding individual cases and
foregrounding the entire teacher sample.

Additional Open Issues and Possible Future Developments

Our review of literature on practice-based pedagogies surfaced additional open


issues. First, as already mentioned, although the framework considers learning to
take place in groups of teachers, more emphasis needs to be given to how these
practice-based pedagogies can support communities of practice (see, for example,
Sowder, 2007, for a discussion of such communities). Second, most of the extant
studies apply the ideas of practice-based pedagogies within a United States context.
Although this is reasonable, given that this is the context where the framework has
been developed and given the current emphasis on practice-based pedagogies in the
United States, future studies ought to test the applicability and transferability of these
ideas in more countries and in different educational systems. Such explorations will
allow for examining the extent to which certain contextual factors might impact on
teacher learning; cultural adaptations of these environments and the reasons that lead
to them will be particularly interesting to explore. Third, the matter of scaling needs

378

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

to be examined. What has been reported so far mostly concerns attempts undertaken
at small scale pertaining to single courses or even parts of these courses; as already
mentioned, often times, the impact of these approaches is reported only for a few
teachers and for individual teaching practices. It remains open to investigate the
effect of larger-scale attempts (including more courses, more practices, and more, as
well, as more diverse teacher participants) on teacher learning. Scaling up this effort
will allow investigating of what is feasible resource-wise, in terms of time, personnel,
course/program structures, and budget. Finally, given the over-emphasis placed on
prospective teachers, a call is made for exploring the role of such environments in
supporting the learning of novice practicing teachers or even practicing teachers at
different stages in their career.
In the next section, we briefly consider both areas – teaching practices and
practice-based pedagogies – and identify additional challenges and open issues for
future consideration.

WORKING AT THE CONFLUENCE OF TEACHING PRACTICES AND


PRACTICE-BASED PEDAGOGY: OTHER CHALLENGES WARRANTING
CONSIDERATION IN THE FUTURE

Reflecting on the call to develop a practice-based pedagogy of teacher education,


Peercy and Troyan (2017) identify additional challenges, besides those listed
above that need to be borne in mind. Although coming from a different discipline
(English Language Arts), the challenges these scholars outline can inform
mathematics education as well. In particular, they recommend more transparency
in how core practices are identified and developed. Given that lists of such
practices in mathematics education are fast accumulating, as discussed above, this
challenge needs careful consideration. The second challenge is that rarely is the
complexity inherent in designing and implementing practice-based pedagogies
made explicit. Hence, these scholars contend, “Specificity and transparency
regarding the enactment of practice-based pedagogy across a number of disciplines
would aid [teacher educators] in designing teacher education with core practices
as an organizing framework” (p. 34). Given the increased emphasis on practice,
these scholars warn that these practices might be divorced from their theoretical
underpinnings. They thus suggest helping prospective and practicing teachers to
focus on the work of teaching without, however, losing sight of the theoretical
background supporting certain practices. Another challenge relates to combining
practices of teaching with other aspects with which beginning teachers need to
be concerned such as textbooks as well as curriculum and behavior management.
Lampert and her colleagues (2013) advise that “practices, principles, and
mathematical knowledge must be used in relation to one another [original italics]
not in isolation” (p. 228).
Finally, we are mindful that attempts to achieve more consistent use of language in
relation to practices of teaching and practice-based pedagogies are limited by current

379

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

economic and political circumstances. Teacher educators are constrained by what is


possible through budgets and university structures. Research teams are constrained
by research commitments which may be inconsistent with selfless collaboration (see
Zeichner, 2012, for a discussion of such constraints).

CONCLUSION

This study reviewed literature on two exciting lines of research in mathematics


education: teaching practices and practice-based pedagogies. The review suggests
that 20 years after Ball and Cohen’s (1999) call to focus on practice and start
developing practice-based curricula, substantial scholarly advances have been
made in this area, in both breadth and depth. In particular, with respect to the
former strand, concerted efforts have been undertaken to decompose the work
of teaching into identifiable practices of varying grain sizes; and some attempts
have been made to study the effect of these practices on student learning. With
respect to the latter strand, Grossman et al.’s (2009) framework of representations,
decomposition, and approximations of practice appears to have offered some
common ground; the last decade has seen attempts to restructure teacher education
programs by including elements of practice-based pedagogies. Despite the rapidly
accumulating work in both strands, several issues remain open: at a conceptual/
theoretical level, at least more consistency in the language used is needed; at
a methodological level, the absence of an agreed-upon suite of methodologies
to empirically examine theoretical arguments advanced in these areas results in
studies that do not build on each other to produce cumulative knowledge. We
argue that stronger and more systematic empirical validation of the potential of
teaching practices and practice-based approaches to teacher education is needed,
something that appears to be one of the central open issues and challenges for the
future.
The challenges and open issues identified in this chapter and the promise of
the work produced to date underline the usefulness of continued, systematic, and
more collaborative work around these issues in the years to come, both within the
United States and internationally. In addition to continue decomposing teaching
to identify practices of teaching and developing practice-based pedagogies – and
finding productive ways of doing so – along with other scholars (e.g., Jacobs
& Spangler, 2017) we argue that this work ought to be directed toward more
systematically exploring how these teaching practices and their incorporation in
practice-based pedagogies can improve instructional quality and through that the
learning of all students. Bearing in mind the open issues and challenges, it seems
that both strands of research examined here offer substantial opportunities for
ongoing work, when examined in isolation, but more critically, when considered
together.

380

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

REFERENCES
[Items marked with an asterisk are the articles used in our literature synthesis, but not referenced in the
chapter.]

Abry, T., Brewer, A. J., Nathanson, L., Sawyer, B., & Rimm-Kaufman, S. E. (2010). Classroom practices
observation measure (Unpublished measure). University of Virginia, Charlottesville, VA.
*Akyuz, D., Dixon, J. K., & Stephan, M. (2013). Improving the quality of mathematics teaching with
effective planning practices. Teacher Development, 17(1), 92–106.
*Anghileri, J. (2006). Scaffolding practices that enhance mathematics learning. Journal of Mathematics
Teacher Education, 9(1), 33–52. http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-006-9005-9
*Averill, R., Drake, M., Anderson, D., & Anthony, G. (2016). The use of questions within in-the-moment
coaching in initial mathematics teacher education: Enhancing participation, reflection, and co-
construction in rehearsals of practice. Asia-Pacific Journal of Teacher Education, 44(5), 486–503. do
i:10.1080/1359866X.2016.1169503
*Awang, H., & Ismail, N. A. (2006). Teaching and learning practices: Their effects on mathematics
achievement. Journal on School Educational Technology, 1(3), 69–73.
*Ball, D. L. (2000). Bridging practices: Intertwining content and pedagogy in teaching and learning to
teach. Journal of Teacher Education, 51(3), 241–247.
*Ball, D. L., Ben-Peretz, M., & Cohen, R. B. (2014). Records of practice and the development of
collective professional knowledge. British Journal of Educational Studies, 62(3), 317–335. doi:10.
1080/00071005.2014.959466
Ball, D. L., & Cohen, D. K. (1999). Developing practice, developing practitioners: Toward a practice-
based theory of professional education. In G. Sykes & L. Darling-Hammond (Eds.), Teaching as the
learning profession: Handbook of policy and practice (pp. 3–32). San Francisco, CA: Jossey-Bass.
*Ball, D. L., & Forzani, F. (2009). The work of teaching and the challenge for teacher education. Journal
of Teacher Education, 60(5), 497–511. doi:10.1177/0022487109348479
Ball, D. L., & Forzani, F. (2010). Teaching skillful teaching. Educational Leadership, 68(4), 40–45.
*Ball, D. L., Sleep, L., Boerst, T. A., & Bass, H. (2009). Combining the development of practice and
the practice of development in teacher education. The Elementary School Journal, 109(5), 458–474.
*Bargagliotti, A., Gottfried, M. A., & Guarino, C. M. (2017). Educating the whole child: Kindergarten
mathematics instructional practices and students’ academic and socioemotional development.
Teachers College Record, 119(8), 1–41.
*Bass, H. (2017). Designing opportunities to learn mathematics theory-building practices. Educational
Studies in Mathematics, 95(3), 229–244. doi:10.1007/s10649-016-9747-y
*Biccard, P., & Wessels, D. (2017). Developing mathematisation practices in primary mathematics
teaching through didactisation-based teacher development. African Journal of Research in
Mathematics, Science and Technology Education, 21(1), 61–73. doi:10.1080/18117295.2017.1283
184
*Blazar, D. (2015). Effective teaching in elementary mathematics: Identifying classroom practices
that support student achievement. Economics of Education Review, 48, 16–29. doi:10.1016/
j.econedurev.2015.05.005
*Boaler, J. (2008). When politics took the place of inquiry: A response to the national mathematics
advisory panel’s review of instructional practices. Educational Researcher, 37(9), 588–594.
doi:10.3102/0013189X08327998
*Boerst, T. A., Sleep, L., Ball, D. L., & Bass, H. (2011). Preparing teachers to lead mathematics
discussions. Teachers College Record, 113(12), 2844–2877.
*Bottia, M. C., Moller, S., Mickelson, R. A., & Stearns, E. (2014). Foundations of mathematics
achievement: Instructional practices and diverse kindergarten students. The Elementary School
Journal, 115(1), 124–150. http://dx.doi.org.ucd.idm.oclc.org/10.1086/676950
*Boyd, D., Grossman, P., Hammerness, K., Lankford, H., Loeb, S., Ronfeldt, M., & Wyckoff, J. (2012).
Recruiting effective math teachers: Evidence from New York City. American Educational Research
Journal, 49(6), 1008–1047. doi:10.3102/0002831211434579

381

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

*Cantun, A., Schutz, K. M., Kelley-Petersen, M., & Franke, M. (2018). Core practices and the teacher
education curriculum: Stories of practice. In P. Grossman (Ed.), Teaching core practices in teacher
education (pp. 107–133). Cambridge, MA: Harvard Education Press.
*Cavanagh, M., & Prescott, A. (2010). The growth of reflective practice among three beginning
secondary mathematics teachers. Asia-Pacific Journal of Teacher Education, 38(2), 147–159.
http://dx.doi.org.ucd.idm.oclc.org/10.1080/13598661003678968
Charalambous, C. Y. (2008). Preservice teachers’ mathematical knowledge for teaching and their
performance in selected teaching practices: Exploring a complex relationship (Unpublished doctoral
dissertation). Ann Arbor, MI: University of Michigan.
*Charalambous, C. Y. (2016). Investigating the knowledge needed for teaching mathematics: An
exploratory validation study focusing on teaching practices. Journal of Teacher Education, 67(3),
220–237. doi:10.1177/0022487116634168
*Charalambous, C. Y., & Kyriakides, E. (2017). Working at the nexus of generic and content-specific
teaching practices: An exploratory study based on TIMSS secondary analyses. The Elementary School
Journal, 117(3), 423–454.
Charalambous, C. Y., Philippou, S., & Olympiou, G. (2018). Reconsidering the use of video clubs for
student-teachers learning during field placement: Lessons drawn from a longitudinal multiple case
study. Teaching and Teacher Education, 74, 49–61. https://doi.org/10.1016/j.tate.2018.04.002
*Charalambous, C. Y., & Pitta-Pantazi, D. (2016). Perspectives on priority mathematics education:
Unpacking and understanding a complex relationship linking teacher knowledge, teaching, and
learning. In L. English & D. Kirshner (Eds.), Handbook of international research in mathematics
education (3rd ed., pp. 19–59). London: Routledge.
*Cheng, J. (2017). Learning to attend to precision: The impact of micro-teaching guided by expert
secondary mathematics teachers on pre-service teachers’ teaching practice. ZDM Mathematics
Education, 49(2), 279–289. doi:10.1007/s11858-017-0839-7
*Chieu, V. M., Kosko, K. W., & Herbst, P. G. (2015). An analysis of evaluative comments in teachers’
online discussions of representations of practice. Journal of Teacher Education, 66(1), 35–50.
doi:10.1177/0022487114550203
*Christman, J. B., Ebby, C. B., & Edmunds, K. A. (2016). Data use practices for improved mathematics
teaching and learning: The importance of productive dissonance and recurring feedback cycles.
Teachers College Record, 118(11), 1–32.
*Cohen, D. (2011). Teaching and its predicaments. Cambridge, MA: Harvard University Press.
*Cohen, J. (2018). Practices that cross disciplines?: Revisiting explicit instruction in elementary
mathematics and English language arts. Teaching and Teacher Education, 69, 324–335.
http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.tate.2017.10.021
*Copur-Gencturk, Y., & Papakonstantinou, A. (2016). Sustainable changes in teacher practices: A
longitudinal analysis of the classroom practices of high school mathematics teachers. Journal of
Mathematics Teacher Education, 19(6), 575–594. http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-
015-9310-2
*Crisan, C., Lerman, S., & Winbourne, P. (2007). Mathematics and ICT: A framework for conceptualising
secondary school mathematics teachers’ classroom practices. Technology, Pedagogy and Education,
16(1), 21–39. http://dx.doi.org.ucd.idm.oclc.org/10.1080/14759390601167991
*Danielson, K. A., Shaughnessy, M., & Jay, L. P. (2018). Use of representations in teacher education.
In P. Grossman (Ed.), Teaching core practices in teacher education (pp. 15–33). Cambridge, MA:
Harvard Education Press.
*da Ponte, J. P., & Chapman, O. (2006). Mathematics teachers’ knowledge and practices. In A. Gutiérrez
& P. Boero (Eds.), Handbook of research on the psychology of mathematics education: Past, present,
and future (pp. 461–494). Rotterdam, The Netherlands: Sense Publishers.
*da Ponte, J. P., & Chapman, O. (2016). Prospective mathematics teachers’ learning and knowledge for
teaching. In L. English & D. Kirshner (Eds.), Handbook of international research in mathematics
education (3rd ed., pp. 275–296). London: Routledge.

382

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

'HOLFH$$\GÕQ( dHYLN.6  0DWKHPDWLFVWHDFKHUV¶XVHRITXHVWLRQV,VWKHUHDFKDQJH


of practice after the curriculum change? Eurasia Journal of Mathematics, Science & Technology
Education, 9(4), 417–427. http://dx.doi.org.ucd.idm.oclc.org/10.12973/eurasia.2013.9410a
*Di Muro, P. (2006). Best practices in mathematics instruction: Teaching for understanding. NADE
Digest, 2(1), 1–8.
*Doabler, C. T., Nelson, N. J., Kennedy, P. C., Stoolmiller, M., Fien, H., Clarke, B., … Baker, S.
K. (2018). Investigating the longitudinal effects of a core mathematics program on evidence-
based teaching practices in mathematics. Learning Disability Quarterly, 41(3) 144–158.
doi:10.1177/0731948718756040
*Doabler, C. T., Nelson, N. J., Kosty, D. B., Fien, H., Baker, S. K., Smolkowski, K., & Clarke, B.
(2014). Examining teachers’ use of evidence-based practices during core mathematics instruction.
Assessment for Effective Intervention, 39(2), 99–111. http://dx.doi.org.ucd.idm.oclc.org/
10.1177/1534508413511848
Douglas, K. (2009). Sharpening our focus in measuring classroom instruction. Educational Researcher,
38(7), 518–521.
Dowling, P. (1998). The sociology of mathematics education: Mathematical myths/pedagogic texts.
Oxon: Routledge, Falmer Press.
*Eddy, S. L., Converse, M., & Wenderoth, M. P. (2015). PORTAAL: A classroom observation tool
assessing evidence-based teaching practices for active learning in large science, technology,
engineering, and mathematics classes. CBE – Life Sciences Education, 14(2), 1–16.
*Erickson, A., & Herbst, P. (2018). Will teachers create opportunities for discussion when teaching proof
in a geometry classroom? International Journal of Science and Mathematics Education, 16(1), 167–
181. doi:10.1007/s10763-016-9764-4
*Escudero, I., & Sánchez, V. (2007). How do domains of knowledge integrate into mathematics teachers’
practice? The Journal of Mathematical Behavior, 26(4), 312–327. http://dx.doi.org.ucd.idm.oclc.org/
10.1016/j.jmathb.2007.11.002
*Essien, A. A. (2010). What teacher educators consider as best practices in preparing pre-service teachers
for teaching mathematics in multilingual classrooms. Perspectives in Education, 28(4), 32–42.
*Feiman-Nemser, S. (2008). Teacher learning: how do teachers learn to teach? In M. Cochran-Smith, S.
Feiman-Nemser, D. J. McIntyre, & K. E. Demers (Eds.), Handbook of research on teacher education:
Enduring questions in changing contexts (3rd ed., pp. 697–705). New York, NY: Routledge.
*Firmender, J. M., Gavin, M. K., & McCoach, D. B. (2014). Examining the relationship between teachers’
instructional practices and students’ mathematics achievement. Journal of Advanced Academics,
25(3), 214–236.
*Forzani, F. (2014). Understanding “core practices” and “practice-based” teacher education: Learning
from the past. Journal of Teacher Education, 65(4), 357–368. doi:10.1177/0022487114533800
*Franke, M. L., & Kazemi, E. (2001). Learning to teach mathematics: Focus on student thinking. Theory
into Practice, 40(2), 102–109. doi:10.1207/s15430421tip4002_4
*Franke, M. L., Kazemi, E., & Battey, D. (2007). Mathematics teaching and classroom practice. In F.
K. Lester (Ed.), Second handbook of research on mathematics teaching and learning (pp. 225–256).
Reston, VA: National Council of Teachers of Mathematics.
*Fyfe, E. R., & Rittle-Johnson, B. (2017). Mathematics practice without feedback: A desirable difficulty
in a classroom setting. Instructional Science, 45(2), 177–194. http://dx.doi.org.ucd.idm.oclc.org/
10.1007/s11251-016-9401-1
*Gainsburg, J. (2012). Why new mathematics teachers do or don’t use practices emphasized in
their credential program. Journal of Mathematics Teacher Education, 15(5), 359–379.
http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-012-9208-1
Gaudin, C., & Chaliès, S. (2015). Video viewing in teacher education and professional development:
A literature review. Educational Research Review, 16, 41–67. http://dx.doi.org/10.1016/
j.edurev.2015.06.001
*Gelzheiser, L. M., Griesemer, B. A., Pruzek, R. M., & Meyers, J. (2000). How are developmentally
appropriate or traditional teaching practices related to the mathematics achievement of

383

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

general and special education students? Early Education and Development, 11(2), 217–238.
http://dx.doi.org.ucd.idm.oclc.org/10.1207/s15566935eed1102_6
*Ghousseini, H. (2015). Core practices and problems of practice in learning to lead classroom discussions.
The Elementary School Journal, 115(3), 334–357.
*Ghousseini, H. (2017). Rehearsals of teaching and opportunities to learn mathematical knowledge for
teaching. Cognition and Instruction, 35(3), 188–211. doi:10.1080/07370008.2017.1323903
*Ghousseini, H., & Herbst, P. (2016). Pedagogies of practice and opportunities to learn about classroom
mathematics discussions. Journal of Mathematics Teacher Education, 19(1), 79–103. doi:10.1007/
s10857-014-9296-1
Gitomer, D. H., & Zisk, R. C. (2015). Knowing what teachers know. Review of Research in Education,
39(1), 1–53. doi:10.3102/0091732X14557001
*Goodchild, S., Fuglestad, A. B., & Jaworski, B. (2013). Critical alignment in inquiry-based practice in
developing mathematics teaching. Educational Studies in Mathematics, 84(3), 393–412.
Goodwin, C. (1994). Professional vision. American Anthropologist, 96(3), 606–633. http://dx.doi.org/
10.1525/aa.1994.96.3.02a00100
*Griffin, C. C., League, M. B., Griffin, V. L., & Bae, J. (2013). Discourse practices in inclusive elementary
mathematics classrooms. Learning Disability Quarterly, 36(1), 9–20. http://dx.doi.org.ucd.idm.oclc.org/
10.1177/0731948712465188
*Grossman, P. (Ed.). (2018). Teaching core practices in teacher education. Cambridge, MA: Harvard
Education Press.
*Grossman, P., Compton, C., Igra, D., Ronfeldt, M., Shahan, E., & Williamson, P. (2009). Teaching
practice: A cross-professional perspective. Teachers College Record, 111(9), 2055–2100.
Grossman, P., Hammerness, K., & McDonald, M., (2009). Redefining teaching, re-imagining teacher
education. Teachers and Teaching: Theory and Practice, 15(2), 273–289.
*Grossman, P., Kavanagh, S. S., & Dean, C. G. P. (2018). The turn towards practice-based education:
Introduction to the work of the core practices consortium. In P. Grossman (Ed.), Teaching core
practices in teacher education (pp. 1–14). Cambridge, MA: Harvard Education Press.
Grossman, P., & McDonald, M. (2008). Back to the future: Directions for research in teaching and teacher
education. American Educational Research Journal, 45(1), 184–205.
*Gujarati, J. (2013). An “inverse” relationship between mathematics identities and classroom practices
among early career elementary teachers: The impact of accountability. The Journal of Mathematical
Behavior, 32(3), 633–648. http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.jmathb.2013.08.002
*Gutiérrez, R. (2002). Enabling the practice of mathematics teachers in context: Toward a new equity research
agenda. Mathematical Thinking and Learning, 4(2–3), 145–187. http://dx.doi.org.ucd.idm.oclc.org/
10.1207/S15327833MTL04023_4
*Han, X., & Paine, L. (2010). Teaching mathematics as deliberate practice through public lessons. The
Elementary School Journal, 110(4), 519–541. http://dx.doi.org.ucd.idm.oclc.org/10.1086/651194
Hatch, T., & Grossman, P. (2009). Learning to look beyond the boundaries of representation: Using
technology to examine teaching (Overview for a digital exhibition: Learning from the practice of
teaching). Journal of Teacher Education, 60(1), 70–85.
*Heller, V. (2015). Academic discourse practices in action: Invoking discursive norms in mathematics
and language lessons. Linguistics and Education, 31, 187–206. http://dx.doi.org.ucd.idm.oclc.org/
10.1016/j.linged.2014.12.003
*Herbel-Eisenmann, B., Lubienski, S. T., & Id-Deen, L. (2006). Reconsidering the study of mathematics
instructional practices: The importance of curricular context in understanding local and global teacher
change. Journal of Mathematics Teacher Education, 9(4), 313–345. http://dx.doi.org.ucd.idm.oclc.org/
10.1007/s10857-006-9012-x
*Herbst, P., Chazan, D., Kosko, K. W., Dimmel, J., & Erickson, A. (2016). Using multimedia questionnaires
to study influences on the decisions mathematics teachers make in instructional situations. ZDM
Mathematics Education, 48(1–2), 167–183. doi:10.1007/s11858-015-0727-y
*Herbst, P., Chieu, V., & Rougée, A. (2014). Approximating the practice of mathematics teaching:
What learning can web-based, multimedia storyboarding software enable? Contemporary Issues in
Technology and Teacher Education, 14(4), 356–383.

384

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

*Herbst, P., & Kosko, K. W. (2014). Using representations of practice to elicit mathematics teachers’ tacit
knowledge of practice: A comparison of responses to animations and videos. Journal of Mathematics
Teacher Education, 17(6), 515–537. doi:10.1007/s10857-013-9267-y
*Hertzog, H. S., & O’Rode, N. (2011). Improving the quality of elementary mathematics student teaching:
Using field support materials to develop reflective practice in student teachers. Teacher Education
Quarterly, 38(3), 89–111.
*Hill, H. C. (2004). Professional development standards and practices in elementary school mathematics.
The Elementary School Journal, 104(3), 215–231. http://dx.doi.org.ucd.idm.oclc.org/10.1086/499750
*Hill, H. C., Blunk, M. L., Charalambous, C. Y., Lewis, J. M., Phelps, G. C., Sleep, L., & Ball, D.
L. (2008). Mathematical knowledge for teaching and the mathematical quality of instruction: An
exploratory study. Cognition and Instruction, 26(4), 430–511. doi:10.1080/07370000802177235
+ODV$& +ODV&6  $UHYLHZRIKLJKဨOHYHUDJHWHDFKLQJSUDFWLFHV0DNLQJFRQQHFWLRQV
between mathematics and foreign languages. Foreign Language Annals, 45(S1), 76–97. doi:10.111/
j.1944-9720.2012.01180.x
*Hong, H., & Chai, C. S. (2017). Principle-based design: Development of adaptive mathematics teaching
practices and beliefs in a knowledge building environment. Computers & Education, 115, 38–55.
http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.compedu.2017.07.011
Horizon Research. (2000). Inside the classroom observation and analytic protocol. Retrieved from
http://www.horizon-research.com/instruments/clas/cop.pdf
*Horn, I. S. (2010). Teaching replays, teaching rehearsals, and re-visions of practice: Learning from
colleagues in a mathematics teacher community. Teachers College Record, 112(1), 225–259.
*House, J. D. (2002). Instructional practices and mathematics achievement of adolescent students in
Chinese Taipei: Results from the TIMSS 1999 assessment. Child Study Journal, 32(3), 157–178.
*Huang, R., Barlow, A. T., & Haupt, M. E. (2017). Improving core instructional practice in mathematics
teaching through lesson study. International Journal for Lesson and Learning Studies, 6(4), 365–379.
doi:10.1108/IJLLS-12-2016-0055
*Huang, R., & Shimizu, Y. (2016). Improving teaching, developing teachers and teacher educators, and
linking theory and practice through lesson study in mathematics: An international perspective. ZDM
Mathematics Education, 48(4), 393–409. doi:10.1007/s11858-016-0795-7
+XJKHV(03RZHOO65/HPENH(6 5LOH\ဨ7LOOPDQ7&  7DNLQJWKHJXHVVZRUNRXW
of locating evidence-based mathematics practices for diverse learners. Learning Disabilities Research
& Practice, 31(3), 130–141. http://dx.doi.org.ucd.idm.oclc.org/10.1111/ldrp.12103
*Ing, M., Webb, N. M., Franke, M. L., Turrou, A. C., Wong, J., Shin, N., & Fernandez, C. H. (2015).
Student participation in elementary mathematics classrooms: The missing link between teacher
practices and student achievement? Educational Studies in Mathematics, 90(3), 341–356. doi:10.1007/
s10649-015-9625-z
*Jackson, K., & Wilson, J. (2012). Supporting African American students’ learning of mathematics:
A problem of practice. Urban Education, 47(2), 354–398. http://dx.doi.org.ucd.idm.oclc.org/
10.1177/0042085911429083
*Jacobbe, T., Ross, D. D., Caron, A. D., Barko, T., & Busi, R. (2014). Connecting theory and practice:
Preservice teachers’ construction of practical tools for teaching mathematics. Teacher Education and
Practice, 27(2), 24.
*Jacobs, V. R., & Spangler, D. A. (2017). Research on core practices in K-12 mathematics teaching. In
J. Cai (Ed.), First compendium for research in mathematics education (pp. 766–792). Reston, VA:
National Council of Teachers of Mathematics.
*Jao, L., Wiseman, D., Kobiela, M., Gonsalves, A., & Savard, A. (2018). Practice-based pedagogy in
mathematics and science teaching methods: Challenges and adaptations in context. Canadian Journal
of Science, Mathematics and Technology Education, 18(2), 177–186.
*Jaworski, B. (2006). Theory and practice in mathematics teaching development: Critical inquiry as a
mode of learning in teaching. Journal of Mathematics Teacher Education, 9(2), 187–211.
*Jong, C., Pedulla, J. J., Reagan, E. M., Salomon-Fernandez, Y., & Cochran-Smith, M. (2010). Exploring
the link between reformed teaching practices and pupil learning in elementary school mathematics.
School Science and Mathematics, 110(6), 309–326.

385

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

*Kaminski, E. (2003). Promoting pre-service teacher education students’ reflective practice in mathematics.
Asia-Pacific Journal of Teacher Education, 31(1), 21–32. http://dx.doi.org.ucd.idm.oclc.org/
10.1080/13598660301619
*Karsenty, R., & Arcavi, A. (2017). Mathematics, lenses and videotapes: A framework and a language
for developing reflective practices of teaching. Journal of Mathematics Teacher Education, 20(5),
433–455. doi:10.1007/s10857-017-9379-x
*Kazemi, E., Ghousseini, H., Cunard, A., & Turrou, A. C. (2016). Getting inside rehearsals: Insights from
teacher educators to support work on complex practice. Journal of Teacher Education, 67(1), 18–31.
doi:10.1177/0022487115615191
Kazemi, E., & Stipek, D. (2001). Promoting conceptual thinking in four upper-elementary mathematics
classrooms. The Elementary School Journal, 102(1), 59–80. http://dx.doi.org/10.1086/499693
*Kelley-Petersen, M., Davis, E. A., Ghousseini, H., Kloser, M., & Monte-Sano, C. (2018). Rehearsals
as examples of approximation. In P. Grossman (Ed.), Teaching core practices in teacher education
(pp. 85–105). Cambridge, MA: Harvard Education Press.
*Kiemer, K., Gröschner, A., Pehmer, A., & Seidel, T. (2015). Effects of a classroom discourse intervention
on teachers’ practice and students’ motivation to learn mathematics and science. Learning and
Instruction, 35, 94–103. http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.learninstruc.2014.10.003
*Kline, S. L., & Ishii, D. K. (2008). Procedural explanations in mathematics writing: A framework for
understanding college students’ effective communication practices. Written Communication, 25(4),
441–461. http://dx.doi.org.ucd.idm.oclc.org/10.1177/0741088308322343
*Kosko, K. W., & Wilkins, J. L. M. (2015). Does time matter in improving mathematical discussions?
The influence of mathematical autonomy. The Journal of Experimental Education, 83(3), 368–385.
doi:10.1080/00220973.2014.907225
Lacelle-Peterson, M. W., & Rivera, C. (1994). Is it real for all kids? A framework for equitable assessment
policies for English language learners. Harvard Educational Review, 64(1), 55–75.
Ladson-Billings, G. (1995). But that’s just good teaching! The case for culturally relevant pedagogy.
Theory into Practice, 34(3), 159–165.
*Lai, M. Y., Auhl, G., & Hastings, W. (2015). Improving pre-service teachers’ understanding of
complexity of mathematics instructional practice through deliberate practice: A case study on “study
of teaching.” International Journal for Mathematics Teaching and Learning, 1–25.
*Lampert, M. (2001). Teaching problems and the problems of teaching. New Haven, CT: Yale University.
*Lampert, M. (2010). Learning teaching in, from, and for practice: What do we mean? Journal of Teacher
Education, 61(1–2), 21–34. doi:10.1177/0022487109347321
Lampert, M., & Ball, D. L. (1998). Teaching, multimedia, and mathematics: Investigations of real
practice. New York, NY: Teachers College Press.
*Lampert, M., Boerst, T. A., & Graziani, F. (2011). Organizational resources in the service of school-wide
ambitious teaching practice. Teachers College Record, 113(7), 1361–1400.
*Lampert, M., Franke, M. L., Kazemi, E., Ghousseini, H., Turrou, A. C., Beasley, H., … Crowe, K.
(2013). Keeping it complex: Using rehearsals to support novice teacher learning of ambitious teaching.
Journal of Teacher Education, 64(3), 226–243. doi:10.1177/0022487112 473837
*Lan, X., Ponitz, C. C., Miller, K. F., Li, S., Cortina, K., Perry, M., & Fang, G. (2009). Keeping their
attention: Classroom practices associated with behavioral engagement in first grade mathematics
classes in China and the United States. Early Childhood Research Quarterly, 24(2), 198–211.
http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.ecresq.2009.03.002
*Lee, C. W., Walkowiak, T. A., & Nietfeld, J. L. (2017). Characterization of mathematics instructional
practises for prospective elementary teachers with varying levels of self-efficacy in classroom
management and mathematics teaching. Mathematics Education Research Journal, 29(1), 45–72.
doi:10.1007/s13394-016-0185-z
*Lerkkanen, M. K., Kiuru, N., Pakarinen, E., Viljaranta, J., Poikkeus, A. M., Rasku-Puttonen, H., …
Nurmi, J. E. (2012). The role of teaching practices in the development of children’s interest in
reading and mathematics in kindergarten. Contemporary Educational Psychology, 37(4), 266–279.
doi:10.1016/j.cedpsych.2011.03.004

386

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

*Lerman, S., & Zehetmeier, S. (2008). Face-to-face communities and networks of practising mathematics
teachers: Studies on their professional growth. In K. Krainer & T. Wood (Eds.), The international
handbook of mathematics teacher education: Participants in mathematics teacher education (Vol. 3,
pp. 33–154). Rotterdam, The Netherlands: Sense Publishers.
*Liang, S., Glaz, S., DeFranco, T., Vinsonhaler, C., Grenier, R., & Cardetti, F. (2013). An examination
of the preparation and practice of grades 7–12 mathematics teachers from the Shandong province in
China. Journal of Mathematics Teacher Education, 16(2), 149–160. http://dx.doi.org.ucd.idm.oclc.org/
10.1007/s10857-012-9228-x
*Lloyd, G. M. (2014). Research into teachers’ knowledge and the development of mathematics classroom
practices. Journal of Mathematics Teacher Education, 17(5), 393–395. http://dx.doi.org.ucd.idm.oclc.org/
10.1007/s10857-014-9285-4
*Lloyd, M. E. R. (2013). Transfer of practices and conceptions of teaching and learning mathematics.
Action in Teacher Education, 35(2), 103–124.
Lortie, D. C. (1975). Schoolteacher. Chicago, MA: The University of Chicago Press.
*Maccini, P., & Gagnon, J. C. (2006). Mathematics instructional practices and assessment accommodations
by secondary special and general educators. Exceptional Children, 72(2), 217–234.
*Maher, C. A. (2008). Video recordings as pedagogical tools in mathematics teacher education. In D.
Tirosh & T. Wood (Eds.), The international handbook of mathematics teacher education: Tools and
processes in mathematics teacher education (Vol. 2, pp. 65–84). Rotterdam, The Netherlands: Sense
Publishers.
*Martínez, J. F., Stecher, B., & Borko, H. (2009). Classroom assessment practices, teacher judgments,
and student achievement in mathematics: Evidence from the ECLS. Educational Assessment, 14(2),
78–102. http://dx.doi.org.ucd.idm.oclc.org/10.1080/10627190903039429
*Maryono, M., Sutawidjaja, A., Subanji, S., & Irawati, S. (2017). Implementation of Pedagogical Content
Knowledge (PCK) of mathematics teachers in teaching practice: A case study. International Education
Studies, 10(3), 11–25.
*Mayrowetz, D. (2009). Instructional practice in the context of converging policies: Teaching mathematics
in inclusive elementary classrooms in the standards reform era. Educational Policy, 23(4), 554–588.
*McCaffrey, D. F., Hamilton, L. S., Stecher, B. M., Klein, S. P., Bugliari, D., & Robyn, A. (2001).
Interactions among instructional practices, curriculum, and student achievement: The case of
standards-based high school mathematics. Journal for Research in Mathematics Education, 32(5),
493–517. http://dx.doi.org.ucd.idm.oclc.org/10.2307/749803
*McClintock, E., O’Brien, G., & Jiang, Z. (2005). Assessing teaching practices of secondary mathematics
student teachers: An exploratory cross case analysis of voluntary field experiences. Teacher Education
Quarterly, 32(3), 139–151.
*McDonald, M., Kazemi, E., & Kavanagh, S. S. (2013). Core practices and pedagogies of teacher
education: A call for a common language and collective activity. Journal of Teacher Education, 64(5),
378–386. doi:10.1177/0022487113493807
*McDuffie, A. M. R., & Mather, M. (2006). Reification of instructional materials as part of the process of
developing problem-based practices in mathematics education. Teachers and Teaching: Theory and
Practice, 12(4), 435–459. http://dx.doi.org.ucd.idm.oclc.org/10.1080/13450600600644285
*McDuffie, A. R. (2004). Mathematics teaching as a deliberate practice: An investigation of elementary
pre-service teachers’ reflective thinking during student teaching. Journal of Mathematics Teacher
Education, 7(1), 33–61. http://dx.doi.org.ucd.idm.oclc.org/10.1023/B:JMTE.0000009970.12529.f4
*McGrew, S., Alston, C. L., & Fogo, B. (2018). Modeling as an example of representation. In P. Grossman
(Ed.), Teaching core practices in teacher education (pp. 35–55). Cambridge, MA: Harvard Education
Press.
*McKinney, S. E., Robinson, J., & Berube, C. T. (2013). “Real teaching” in the mathematics classroom:
A comparison of the instructional practices of elementary teachers in urban high-poverty schools.
Teacher Education and Practice, 26(4), 797–815.
*Merritt, E. G., Palacios, N., Banse, H., Rimm-Kaufman, S., & Leis, M. (2017). Teaching practices
in Grade 5 mathematics classrooms with high-achieving English learner students. The Journal of

387

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

Educational Research, 110(1), 17–31. http://dx.doi.org.ucd.idm.oclc.org/10.1080/00220671.2015.10


34352
*Miller, S. P., & Hudson, P. J. (2007). Using evidence-based practices to build mathematics competence
related to conceptual, procedural, and declarative knowledge. Learning Disabilities Research &
Practice, 22(1), 47–57. http://dx.doi.org.ucd.idm.oclc.org/10.1111/j.1540-5826.2007.00230.x
*Morgan, P. L., Farkas, G., & Maczuga, S. (2015). Which instructional practices most help first-grade
students with and without mathematics difficulties? Educational Evaluation and Policy Analysis,
37(2), 184–205. http://dx.doi.org.ucd.idm.oclc.org/10.3102/0162373714536608
*Moschkovich, J. (2013). Principles and guidelines for equitable mathematics teaching practices and
materials for English language learners. Journal of Urban Mathematics Education, 6(1), 45–57.
*Ottmar, E. R., Rimm-Kaufman, S. E., Berry, R. Q., & Larsen, R. A. (2013). Does the responsive classroom
approach affect the use of standards-based mathematics teaching practices? The Elementary School
Journal, 113(3), 434–457.
*Ottmar, E. R., Rimm-Kaufman, S., Larsen, R. A., & Berry, R. Q. (2015). Mathematical knowledge for
teaching, standards-based mathematics teaching practices, and student achievement in the context
of the responsive classroom approach. American Educational Research Journal, 52(4), 787–821.
http://dx.doi.org.ucd.idm.oclc.org/10.3102/0002831215579484
Peercy, M. M., & Troyan, F. J. (2107). Making transparent the challenges of developing a practice-
based pedagogy of teacher education. Teaching and Teacher Education, 61, 26–36. http://dx.doi.org/
10.1016/j.tate.2016.10.005
Praetorius, A. K., & Charalambous, C. Y. (2018). Classroom observation frameworks for studying
instructional quality: Looking back and looking forward. ZDM Mathematics Education, 50(3),
535–553. doi:10.1007/s11858-018-0946-0
Reynolds, D., Chapman, C., Clarke, P., Muijs, D., Sammons, P., & Teddlie, C. (2016). Conclusions: The
future of educational effectiveness and improvement research, and some suggestions and speculations.
In C. Chapman, D. Muijs, D. Reynolds, P. Sammons, & C. Teddlie (Eds.), The Routledge international
handbook of educational effectiveness and improvement: Research, policy, and practice (pp. 408–
439). New York, NY: Routledge.
*Rosenquist, B. A., Henrick, E. C., & Smith, T. M. (2015). Research–practice partnerships to support
the development of high quality mathematics instruction for all students. Journal of Education for
Students Placed at Risk, 20(1–2), 42–57. http://dx.doi.org.ucd.idm.oclc.org/10.1080/
10824669.2014.988335
*Russ, R. S., Sherin, B. L., & Sherin, M. G. (2016). What constitutes teacher learning? In D. H. Gitomer
& C. A. Bell (Eds.), Handbook of research on teaching (5th ed., pp. 391–438). Washington, DC:
American Educational Research Association.
*Santagata, R. (2005). Practices and beliefs in mistake-handling activities: A video study of Italian
and US mathematics lessons. Teaching and Teacher Education, 21(5), 491–508.
http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.tate.2005.03.004
*Santagata, R., & Yeh, C. (2014). Learning to teach mathematics and to analyze teaching effectiveness:
Evidence from a video- and practice-based approach. Journal of Mathematics Teacher Education,
17(6), 491–514. http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-013-9263-2
*Schutz, K. M., Grossman, P., & Shaughnessy, M. (2018). Approximations of practice in teacher education.
In P. Grossman (Ed.), Teaching core practices in teacher education (pp. 57–83). Cambridge, MA:
Harvard Education Press.
*Shaughnessy, M., & Boerst, T. A. (2018). Uncovering the skills that preservice teachers bring to teacher
education: The practice of eliciting a student’s thinking. Journal of Teacher Education, 69(1), 40–55.
doi:10.1177/0022487117702574
*Shechtman, N., Roschelle, J., Haertel, G., & Knudsen, J. (2010). Investigating links from teacher
knowledge, to classroom practice, to student learning in the instructional system of the middle-school
mathematics classroom. Cognition and Instruction, 28(3), 317–359. http://dx.doi.org.ucd.idm.oclc.org/
10.1080/07370008.2010.487961
*Sileo, J. M., & van Garderen, D. (2010). Creating optimal opportunities to learn mathematics: Blending
co-teaching structures with research-based practices. Teaching Exceptional Children, 42(3), 14–21.

388

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
MATHEMATICS TEACHING PRACTICES AND PRACTICE-BASED PEDAGOGIES

*Silver, E. (2010). Examining what teachers do when they display their best practice: Teaching
mathematics for understanding. Journal of Mathematics Education at Teachers College, 1(1), 1–6.
*Silver, E. A., Clark, L. M., Ghousseini, H. N., Charalambous, C. Y., & Sealy, J. T. (2007). Where
is the mathematics? Examining teachers’ mathematical learning opportunities in practice-based
professional learning tasks. Journal of Mathematics Teacher Education, 10(4–6), 261–277.
http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-007-9039-7
*Sleep, L. (2012). The work of steering instruction toward the mathematical point: A
decomposition of teaching practice. American Educational Research Journal, 49(5), 935–970.
doi:10.3102/0002831212448095
*Sleep, L., & Boerst, T. A. (2012). Preparing beginning teachers to elicit and interpret students’
mathematical thinking. Teaching and Teacher Education, 28(7), 1038–1048. doi:10.1016/j.
tate.2012.04.005
Smith, M. S., & Stein, M. K. (2011). 5 practices for orchestrating productive mathematics discussions.
Reston, VA: National Council of Teachers of Mathematics.
*Sowder, J. T. (2007). The mathematical education and development of teachers. In F. K. Lester (Ed.),
Second handbook of research on mathematics teaching and learning (pp. 157–223). Reston, VA:
National Council of Teachers of Mathematics.
*Spooner, F., Saunders, A., Root, J., & Brosh, C. (2017). Promoting access to common core mathematics
for students with severe disabilities through mathematical problem solving. Research and Practice for
Persons with Severe Disabilities, 42(3), 171–186. doi:10.1177/1540796917697119.
*Staples, M. E., Bartlo, J., & Thanheiser, E. (2012). Justification as a teaching and learning practice: Its
(potential) multifacted role in middle grades mathematics classrooms. The Journal of Mathematical
Behavior, 31(4), 447–462. http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.jmathb.2012.07.001
Stein, M. K., Engle, R. A., Smith, M. S., & Hughes, E. K. (2008). Orchestrating productive mathematical
discussions: Five practices for helping teachers move beyond show and tell. Mathematical Thinking
and Learning, 10, 310–340.
Stigler, J. W., & Hiebert, J. (1999). The teaching gap: Best ideas from the world’s teachers’ for improving
education in the classroom. New York, NY: The Free Press.
Stipek, D., & Byler, P. (2004). The early childhood classroom observation measure. Early Childhood
Research Quarterly, 19, 375–397.
*Stockero, S. L., & Van Zoest, L. R. (2013). Characterizing pivotal teaching moments in beginning
mathematics teachers’ practice. Journal of Mathematics Teacher Education, 16(2), 125–147.
http://dx.doi.org.ucd.idm.oclc.org/10.1007/s10857-012-9222-3
Stylianides, A. J., & Delaney, S. (2011). The cultural dimension of teachers’ mathematical knowledge.
In T. Rowland & K. Ruthven (Eds.), Mathematical knowledge in teaching (Vol. 50, pp. 179–191).
Dordrecht: Springer.
Sun, J., & van Es, E. A. (2015). An exploratory study of the influence that analyzing teaching has
on preservice teachers’ classroom practice. Journal of Teacher Education, 66(3), 201–214.
https://doi.org/10.1177/0022487115574103
*Suurtamm, C., Koch, M., & Arden, A. (2010). Teachers’ assessment practices in mathematics:
Classrooms in the context of reform. Assessment in Education: Principles, Policy & Practice, 17(4),
399–417. http://dx.doi.org.ucd.idm.oclc.org/10.1080/0969594X.2010.497469
*Swars, S. L., Smith, S. Z., Smith, M. E., Carothers, J., & Myers, K. (2018). The preparation experiences
of elementary mathematics specialists: Examining influences on beliefs, content knowledge, and
teaching practices. Journal of Mathematics Teacher Education, 21(2), 123–145. doi:10.1007/s10857-
016-9354-y
Tarr, J. E., Reys, R. E., Reys, B. J., Chávez, Ó., Shih, J., & Osterlind, S. J. (2008). The impact of middle
grade mathematics curricula and the classroom learning environment on student achievement. Journal
for Research in Mathematics Education, 39(3), 247–280.
*Tchoshanov, M. A. (2011). Relationship between teacher knowledge of concepts and connections,
teaching practice, and student achievement in middle grades mathematics. Educational Studies in
Mathematics, 76(2), 141–164.

389

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
CHARALAMBOS Y. CHARALAMBOUS & SEÁN DELANEY

*Thomas, K. (2013). Changing mathematics teaching practices and improving student outcomes through
collaborative evaluation. Teacher Education and Practice, 26(4), 779–796.
*Tyminski, A. M., Zambak, V. S., Drake, C., & Land, T. J. (2014). Using representations, decomposition,
and approximations of practices to support prospective elementary mathematics teachers’ practice of
organizing discussions. Journal of Mathematics Teacher Education, 17(5), 463–487. doi:10.1007/
s10857-013-9261-4
*Ukpokodu, O. N. (2011). How do I teach mathematics in a culturally responsive way? Identifying
empowering teaching practices. Multicultural Education, 19(3), 47–56.
*Wager, A. A. (2012). Incorporating out-of-school mathematics: From cultural context to embedded
practice. Journal of Mathematics Teacher Education, 15(1), 9–23. http://dx.doi.org.ucd.idm.oclc.org/
10.1007/s10857-011-9199-3
*Webb, N. M., Franke, M. L., De, T., Chan, A. G., Freund, D., Shein, P., & Melkonian, D. K. (2009).
‘Explain to your partner’: Teachers’ instructional practices and students’ dialogue in small groups.
Cambridge Journal of Education, 39(1), 49–70. doi:10.1080/03057640802701986
*Webb, N. M., Franke, M. L., Ing, M., Turrou, A. C., Johnson, N. C., & Zimmerman, J. (2017). Teacher
practices that promote productive dialogue and learning in mathematics classrooms. International
Journal of Educational Research. (Advance online publication) doi:10.1016/j.ijer.2017.07.009
*Webb, N. M., Franke, M. L., Ing, M., Wong, J., Fernandez, C. H., Shin, N., & Turrou, A. C. (2014).
Engaging with others’ mathematical ideas: Interrelationships among student participation, teachers’
instructional practices, and learning. International Journal of Educational Research, 63, 79–93.
doi:10.1016/j.ijer.2013.02.001
*Wieman, C., & Gilbert, S. (2014). The teaching practices inventory: A new tool for characterizing
college and university teaching in mathematics and science. CBE-Life Sciences Education, 13(3),
552–569. doi:10.1187/cbe.14-02-0023
*Xenofontos, C. (2016). Teaching mathematics in culturally and linguistically diverse classrooms:
Greek-Cypriot elementary teachers’ reported practices and professional needs. Journal of Urban
Mathematics Education, 9(1), 94–116.
*Yeh, C. (2017). Math is more than numbers: Beginning bilingual teachers’ mathematics teaching
practices and their opportunities to learn. Journal of Urban Mathematics Education, 10(2), 106–139.
*Yu, R., & Singh, K. (2018). Teacher support, instructional practices, student motivation, and mathematics
achievement in high school. The Journal of Educational Research, 111(1), 81–94. doi:10.1080/0022
0671.2016.1204260
*Zahner, W., Velazquez, G., Moschkovich, J., Vahey, P., & Lara-Meloy, T. (2012). Mathematics
teaching practices with technology that support conceptual understanding for Latino/a students. The
Journal of Mathematical Behavior, 31(4), 431–446. http://dx.doi.org.ucd.idm.oclc.org/10.1016/j.
jmathb.2012.06.002
*Zeichner, K. (2012). The turn once again toward practice-based teacher education. Journal of Teacher
Education, 63(5), 376–382.

Charalambos Y. Charalambous
Department of Education
University of Cyprus

Seán Delaney
Marino Institute of Education

390

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP, JOHN (ZIG) SIEGFRIED AND
EVA THANHEISER

14. SEEING MATHEMATICS THROUGH THE LENS


OF CHILDREN’S MATHEMATICAL THINKING
Perspective on the Development of Mathematical Knowledge for Teaching

Just as effective teaching of mathematics requires that practicing teachers focus


simultaneously on mathematics and on children’s ways of reasoning about
mathematics, we make the case that this dual focus is also important for the
preparation of prospective teachers. Viewing mathematics through the lens
of children’s mathematical thinking motivates prospective teachers’ learning
of mathematics for teaching, and, more important, changes and deepens the
mathematics they see. Finally, we highlight how a focus on mathematics through
the lens of children’s mathematical thinking, including a focus on culturally relevant
mathematical differences, can be leveraged to support teachers while their view of
children becomes richer and more nuanced.

INTRODUCTION

To teach mathematics in the 21st century, teachers must possess robust knowledge
of mathematics, of pedagogical content knowledge (Shulman, 1986), of children
as learners, and of the social contexts of mathematics teaching and learning
(Association of Mathematics Teacher Educators, 2017). Furthermore, because the
knowledge, skills, and dispositions required of teachers are so complex and are
constantly changing, teachers must also learn how to learn from teaching. We focus
on one dimension of learning to teach: the development of mathematical knowledge
for teaching mathematics to children.1 Many have discussed knowledge for teaching
mathematics, and later in the chapter we review that work, but our central thesis
in this chapter is different from that put forth by others: We propose that the order
in which prospective elementary school teachers develop their knowledge for
teaching mathematics is important. In particular, when prospective elementary
school teachers focus upon children’s mathematical thinking either concurrently
with or even before learning mathematics, then their mathematical development is
qualitatively richer. We refer to this approach as viewing mathematics through the
lens of children’s mathematical thinking, and in this chapter, we make the case that
taking this approach will not only support the development of richer mathematics

© KONINKLIJKE BRILL NV, LEIDEN, 2020 | DOI: 10.1163/9789004418875_015

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

but will also provide a foundation for prospective elementary school teachers to
learn how to learn from practice. We begin with an example.

EXAMPLE 1: MARIA

Consider the following situation: Maria has 4 marbles. How many marbles does
Maria need to get so that she has 9 marbles altogether? Many adults, including
prospective elementary school teachers, view this problem as subtraction because
the answer, 5, is the result of subtracting 4 from 9. When analyzing the mathematical
issues in this task, prospective elementary school teachers might consider memorizing
subtraction facts or counting (down) using fingers. However, younger children
tend to view this situation as additive, matching the action in the problem context
(Carpenter, Fennema, Franke, Levi, & Empson, 2015). Further-more, children apply
a range of solution strategies when solving this task. Children may directly model
the action, whereby they represent each quantity in the problem (Carpenter et al.,
2015). For example, they may count out four manipulatives, then count up from 4 to
9, and then count the manipulatives they added. Or children may start with 4 “in their
heads,” and then count up to 9 using manipulatives or fingers, a strategy that is more
sophisticated than the previous strategy because the child did not directly model the
initial value. Further, children may use derived facts, whereby they use a fact that
they already know, in this case likely 4 + 4 = 8 or 5 + 5 = 10, to arrive at the fact they
do not know, 4 + 5 = 9. Eventually, children will use their memorized number-fact
family, 4 + 5 = 9 and 9 – 4 = 5, to immediately know the answer.

A FRAMEWORK FOR LOOKING AT SCHOOL MATHEMATICS THROUGH THE


LENS OF CHILDREN’S MATHEMATICAL THINKING

Our purpose for this chapter is to highlight how viewing mathematics through the
lens of children’s mathematical thinking broadens what one sees and consequently
has implications for teacher education. Figure 14.1 represents what an adult, such as
a prospective elementary school teacher, sees when looking at school mathematics,

Figure 14.1. Prospective elementary school teachers looking at school mathematical issues

392

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

Figure 14.2. Prospective elementary school teachers looking at school mathematical issues
through the lens of children’s mathematical thinking see richer and different mathematics

and Figure 14.2 represents what an adult sees when viewing school mathematics
through the lens of children’s mathematical thinking. In the case of Maria, viewing
the mathematics from the adult perspective, represented by the yellow oval in Figure
14.1, might include thinking of memorizing facts or counting down. Both of these
strategies are mathematically valid. However, when viewing the mathematics through
the lens of children’s mathematical thinking, the adult may see the same mathematics
but might also see richer mathematics involving more connections among ideas,
represented by the green oval. Furthermore, the adult may see new or different
mathematics, represented by the expanded blue region. Maria’s solution as viewed
through the lens of children’s mathematical thinking (which might be represented
by the green oval) represents richer mathematics involving more connections than
viewing mathematics solely from the adult’s perspective and includes prospective
elementary school teachers’ seeing the compensation involved in the relationship
between 4 + 5 and 4 + 4 (or 4 + 5 and 5 + 5). Examples that might be represented
by the blue oval, representing new or different mathematics, include prospective
elementary school teachers’ explicating the relationship between a situation in life
and a mathematical model to represent that relationship so that, in this case, what
adults see as subtraction might more validly be modeled for children as addition, or
the role that mathematical properties, in this case, the associative property, play in
everyday mathematics.
Fundamentally, the goal for mathematics instruction is for teachers to support
children in developing their mathematical thinking, and all teachers of mathematics
are committed to this outcome (Stigler & Hiebert, 1999). However, although teachers
are committed to supporting children’s mathematical learning, they may do so by
focusing in different ways. A teacher may focus on (a) the correctness of a child’s
work; (b) whether the child’s reasoning matched the intended reasoning of the
teacher; (c) the ways in which the child is mathematically making sense, irrespective
of the correctness of the child’s solution; and (d) what mathematical issues are
raised by a child’s reasoning. Although some teachers focus on all four of these

393

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

areas simultaneously, teachers’ beliefs and orientations more commonly direct them
toward the first two: the correctness of children’s answers and whether children’s
reasoning matches the teachers’ ways of thinking (Philipp, 2007; Thompson, 1992;
Thompson, Philipp, Thompson, & Boyd, 1994).
When teachers focus primarily on the correctness of children’s work or on the
children’s learning to apply the teacher’s approach, the reasoning of children may
remain implicit and hence not be the focus of the class. However, when the teacher’s
focus is on children’s reasoning, including when the child’s reasoning is partially
or entirely incorrect, then the teacher may elevate mathematical thinking as the
central focus of the lesson. Furthermore, when focusing on children’s reasoning, the
teacher enhances opportunities for children to grapple with the mathematical issues,
sometimes issues that are deep and rich.

DEFINING CHILDREN’S MATHEMATICAL THINKING

Implicit in a framework for looking at school mathematics through the lens of children’s
mathematical thinking is a definition of children’s mathematical thinking. We define
children’s mathematical thinking as children making sense of a mathematical situation,
context, or task with the tools the children have available. This definition may seem no
different from a definition of adult’s mathematical thinking, but we offer an implication
for a difference in how children and adults learn school mathematics because of the
difference in tools the two groups have available. When prospective elementary school
teachers revisit school mathematics from a conceptual perspective, they approach the
mathematical topics having already learned as procedures for calculating (Browning
et al., 2014; Olanoff, Lo, & Tobias, 2014; Thanheiser & Browning, 2014; Thanheiser,
Browning, et al., 2014; Thanheiser, Whitacre, & Roy, 2014), but when children learn
a topic, they may be exposed to concepts prior to learning the associated procedures.
At first glance, one might think that prospective elementary school teachers, therefore,
have more entry points into the mathematics; we suspect, however, that the procedures
will dominate, thereby decreasing the likelihood that they will grapple with the
underlying concepts (Pesek & Kirshner, 2000). In contrast, children, who have yet to
learn procedures, can be encouraged to grapple with the concepts, and, in doing so,
begin to discover the procedures for themselves.

ASSUMPTIONS

We identify several assumptions associated with looking at school mathematics


through the lens of children’s mathematical thinking. First, these are assumptions
that we, the authors, hold, and we share them so that we might explicate for the
readers the foundational beliefs on which our theoretical framework rests. Second,
these are beliefs that we hope practicing teachers hold, although we recognize that
teachers fall along a spectrum vis-à-vis the extent to which they hold these beliefs
(Philipp, 2007). Finally, these are beliefs that we hope would be associated with

394

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

courses and experiences for prospective elementary school teachers so that they
might begin to develop these assumptions as shared beliefs.
The notion that viewing mathematics through the lens of children’s mathematical
thinking will help prospective elementary school teachers see different and richer
mathematics is based upon epistemological and pedagogical assumptions. When
these assumptions are not embraced, we see limited effect of the lens of children’s
mathematical thinking. Consider an example related to the mathematical topic
of fraction division, with two hypothetical teachers, Teacher A and Teacher B,
representing differing sets of epistemological and pedagogical assumptions. Teacher
A views the topic of fraction division as involving the relationships among rich
conceptual understanding, adaptive reasoning, problem solving, and procedural
fluency, as highlighted in the strands of mathematical proficiency (National
Research Council, 2001), whereas Teacher B views the topic in a solely procedural
manner whereby the quotient of two fractions involves decontextualized tasks
such as 1 1/2 ÷ 1/3 . Furthermore, whereas Teacher A takes a pedagogical approach
associated with children’s grappling with mathematical ideas and hence believes that
children must be given novel tasks about which to think, in Teacher B’s pedagogical
approach, the expectation is that teachers must explicitly demonstrate a procedure
that children learn through imitation and practice. Figures 14.3 and 14.4 highlight
these two approaches.
Both of these teachers might think that they are paying attention to children’s
thinking, but Teacher B is really attending only to anticipated difficulties students
might have in applying a procedure. This view for Teacher B, the teacher with the
procedural-only view of mathematics for whom teaching is telling, might involve
attending to the kinds of challenges children experience with the procedure, which
might include difficulty converting a mixed number into an improper fraction,

Figure 14.3. An approach to the topic of fraction division as an integration of concepts,


procedures, reasoning, and problem solving

1 1 3 1 3 3 9 1
1 ÷ = ÷ = × = =4
2 3 2 3 2 1 2 2

Figure 14.4. A solely procedural approach to the topic of fraction division

395

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

the tendency to forget the steps in the procedure, or difficulty with converting the
final improper fraction back into a mixed number. Although this perspective might
enhance the procedural instruction, fundamentally the children’s mathematical lens
is absent from Teacher B’s view and hence cannot affect the problem’s mathematical
richness for the teacher.
Now consider Teacher A’s view of this topic through the lens of children’s
mathematical thinking. For example, a child might understand a ÷ b as “How many
bs are in a?” and reason that there are 3 thirds in 1, another third in 1/2 with 1/6
left, and the child might incorrectly provide the answer of 4 1/6 as opposed to 4 1/2.
(This is precisely the reasoning of Elliot, a child in Video #6 at this webpage:
http://www.sci.sdsu.edu/CRMSE/sdsu-pdc/nickerson/imap.htm.) Several mathematical
issues are raised for one viewing this mathematics through the lens of children’s
mathematical thinking. First, the child’s answer is incorrect, so one might
conclude that he does not understand division. However, that conclusion would be
oversimplified, because the child correctly conceptualized the task as finding “How
many 1/3 s are in 11/2?” Furthermore, the child correctly determined that there are 4
whole 1/3 s in 11/2, with a remaining 1/6 (of a whole). However, when the child wrote
the answer as 4 1/6, he was unaware that he was treating the 1/6 as a remainder instead
of as part of the quotient. (Note that the issue of remainder versus quotient is also
relevant with whole number division, for example, consider the context of sharing 14
grapes evenly among 4 children, which might be represented as 14 ÷ 4. The answer
might be represented as 3 R 2, or 3 1/2; both are correct, but they involve different
mathematical understandings because whereas children at the age of 5 or 6 can share
14 grapes among 4 people and see that 2 are left over, they may not yet be able to
conceptualize the remaining 2 grapes as 1/2 of a grape for each of the 4 children.)
This common error raises important mathematical issues that come to the fore when
prospective elementary school teachers are required to grapple meaningfully with
fraction division, or even with division more generally.
We explicate four assumptions that we make regarding mathematics, regarding
children as sense makers, about the need to care about children’s reasoning, and
about teachers’ learning. When appropriate, we draw upon the previous example in
presenting these assumptions.

Assumption #1: Mathematics Entails Rich Interconnections

Seeing through the lens of children’s mathematical thinking must be associated with
a view of mathematics as connections among concepts, procedural fluency, and
problem solving, built upon an underlying foundation of mathematical reasoning,
including justifying one’s reasoning.
Assumption #1 was highlighted in the previous example. By focusing solely
on learning procedures, Teacher B demonstrated a view of mathematics as not
comprised of rich interconnections. Consequently, even when challenges in applying
the algorithm are viewed through the lens of children’s mathematical thinking, this

396

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

teacher’s view of mathematics constrained the extent to which employing the lens of
children’s mathematical reasoning enhanced the mathematical richness.

Assumption #2: Children Are Sense Makers

Seeing through the lens of children’s mathematical thinking is based upon a view
of children as sense makers, including the view that children hold pre-instructional
knowledge on which subsequent learning must be built, and that, when given
opportunities, children of all ages are able to engage in rich and creative mathematical
reasoning.
Teachers who do not hold Assumption #2 would be unlikely to inquire about
children’s reasoning prior to formal instruction. Furthermore, if prospective
elementary school teachers do not hold this assumption, then even when they view
children’s rich or creative reasoning, they will attribute that reasoning to the children’s
having been taught rather than to their engaging creatively with mathematical ideas.
An anecdote might help the reader understand Assumption #2: A mathematics
teacher educator arranged for prospective elementary school teachers enrolled in
her elementary mathematics methods class to interview kindergarten children in
the fall, during the second week of school. One goal of this teacher educator was
for prospective elementary school teachers to recognize how much mathematical
understanding children bring to school, because at the time that this interview took
place, kindergarten was the first year of schooling in the United States. After the
prospective elementary school teachers conducted the interviews, they met to discuss
the variety of ways of reasoning about addition and subtraction situations they had
observed during the interviews. One of the prospective elementary school teachers
then shared that she was shocked to see how much these children had learned in just
one week of school! Without holding Assumption #2, this prospective elementary
school teacher was unable to validly associate the source of these children’s
understanding to their outside-of-school experiences.
Associated with Assumption #2 is the notion that understanding is never an all-
or-nothing entity and instead students hold understandings that might be thought of
as existing along a continuum (Van de Walle, Karp, Lovin, & Bay-Williams, 2014)
or at varying levels of development (Carpenter et al., 2015). That is, almost no one
understands everything about anything, and almost everyone understands something
about anything, and effective teaching involves finding ways to connect to each
student’s understanding.

Assumption #3: Caring about the Reasoning of Children Is Essential for Effective
Teaching

Seeing through the lens of children’s mathematical thinking must be associated with
a commitment to understanding children’s mathematical thinking.

397

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Historically, mathematics teaching has not been associated with teachers’ focusing
on children’s mathematical reasoning. Instead, teachers and children approached
the classroom with the expectation that the children would watch and listen while
the teacher explained, and then the children would practice (Stanic & Kilpatrick,
2003), an approach consistent with the historically prevailing behavioristic view
of learning (National Research Council, 2001). Theories of learning have evolved
beyond behaviorism, and with them we identify three reasons that seeing through
the lens of children’s mathematical thinking requires a focus on children’s thinking:
first, so that teachers might build instruction on children’s understandings; second,
because children differ in their understandings and so teachers must learn what
different children understand; and third, when a teacher approaches teaching with
the intellectual humility associated with recognizing that the teacher’s approach is
but one way, not the way, to reason, then the teacher can approach children with
sincere intellectual curiosity.

Assumption #4: Teachers Must Continue to Learn throughout Their Careers

The knowledge, skills, and dispositions required for effective instruction cannot be
mastered during a teacher preparation program.
Teachers cannot learn all that they must learn from teacher preparation or teacher
credential programs because teaching is far too complex. Consequently, teachers
must continue to learn throughout their careers; they must learn how to learn from
practice. However, to accomplish this, teachers must approach their profession
through a stance of learning, and this must start with their preparation as teachers.

THE CHALLENGE OF INVOKING THESE ASSUMPTIONS AND BELIEFS

When juxtaposed with earlier views of education, these assumptions about


mathematics, about children as sense makers, about caring about the reasoning
of children, and about the need for teachers to be career-long learners create a
challenging set of beliefs. In the past, children were viewed as passive receivers
of mathematical knowledge who did not have their own creative ways of thinking
mathematically (Stanic & Kilpatrick, 2003). However, when educators assume that
children are sense makers and that the way they make sense, even if not entirely
correct, is important, then the role of teachers of mathematics becomes far more
complex than in the past. Teachers now not only must create a welcoming intellectual
climate wherein children are encouraged to share their reasoning but also must be
able to make sense of the children’s thinking and use it when making instructional
decisions (Smith & Stein, 2011). For example, in addition to attending to children’s
thinking, teachers must interpret the children’s thinking from the points of view
of both the underlying mathematical ideas and the research on children’s thinking.
Further, they must respond, in the moment, to children’s thinking to build on that
thinking and also help other children understand the reasoning of their classmates

398

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

(Jacobs, Lamb, & Philipp, 2010). When teachers shift from focusing primarily on
mathematical correctness to also considering the richness of children’s mathematical
reasoning, the classroom becomes pedagogically rich because the teacher may
elevate mathematical thinking as the central focus of the lesson. Furthermore, when
focusing on children’s reasoning, the teacher enhances the children’s opportunities
to grapple with the mathematical issues, sometimes issues that are deep and rich.
Focusing on children’s mathematical thinking is difficult. A teacher who is
open to listening to children’s thinking may discover that some children are more
mathematically creative than the teacher. Note, we do not claim that the child
knows more mathematics than the teacher but rather that the child’s approach to
mathematics may be more open, inquisitive, and creative than the teacher’s approach.
Preparing teachers to work in such environments must include supporting teachers in
developing mathematical flexibility (Rowland, 2013); seeing mathematics through
the lens of children’s mathematical thinking can support this approach.
We do not suggest that prospective elementary school teachers must become
experts on children’s mathematical thinking nor that children’s mathematical
thinking must become the focus of prospective teachers’ mathematical knowledge
for teaching. We do suggest that when prospective teachers become practicing
teachers, they need to have developed rich expertise about children’s mathematical
thinking as it relates to the children they are teaching, and this expertise involves
the integration of the mathematical domain and children’s thinking. We suggest also
that viewing mathematics through the lens of children’s mathematical thinking not
only begins to orient prospective teachers toward the integration of mathematics
and children’s thinking but also supports the development of an orientation toward
looking at children’s mathematics thinking through a new and rich lens, prompting
them to wonder, “What is valid and meaningful in children’s mathematical
reasoning, and what are we still missing?” We have found in our own work that
to answer such a question requires a kind of mathematical understanding that
prospective teachers, in general, are not developing. In the next section, we review
research on professional knowledge for teaching, but first we reiterate our central
thesis in this chapter and discuss how it differs from that put forth by others. Our
focus on looking through the lens of children’s mathematical thinking is different
from the focus taken by others in two ways. First, whereas other researchers have
investigated the effects of teachers’ focusing on children’s mathematical thinking as
a means by which to support teachers while they refocus their instruction, we are
also interested in the effects on the mathematical development of teachers. Second,
our focus has implications for the order in which teachers develop knowledge of
mathematics content and children’s mathematical thinking, because if viewing
mathematics through the lens of children’s mathematical thinking leads to deeper
and richer mathematical learning, then prospective elementary school teachers must
not be asked to delay learning about children’s mathematical thinking until after they
study mathematics for teaching.

399

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Having shared our framework for viewing mathematics through the lens of
children’s mathematical thinking, defined children’s mathematical thinking, and
shared our assumptions, we turn to the literature to consider professional knowledge
for teachers of mathematics. After reviewing the literature, we return with additional
examples of viewing mathematics through the children’s mathematical-thinking
lens.

PROFESSIONAL KNOWLEDGE FOR TEACHERS OF MATHEMATICS: A


REVIEW OF THE LITERATURE

Professional knowledge for teachers of mathematics has been conceptualized in


multiple ways for various purposes (Neubrand, 2018). Building on Shulman (1998)
and Kunter et al. (2013), Neubrand listed characteristic features of a profession as
follows:
(1) The core is that professions are obliged to serve others, which also embraces
some personal responsibility. (2) To that end, they need an understanding of
a scholarly or theoretical kind, i.e., a profession refers to a domain of skilled
performance or practice. (3) Professionals are used to exercising judgments
under conditions of unavoidable uncertainty. (4) They therefore need learning
from experience, as theory and practice in a profession always interact. (5)
Finally, professionals need systematically organized training and are part of a
professional community to monitor quality and aggregate knowledge. (p. 601)
In this chapter, we focus on the professional knowledge for teachers of
mathematics, with a particular focus on the kind of understanding that comes from
attending to children’s mathematical thinking. We build on the above definition
of features of a profession, especially Items (2) and (4), because teachers need to
understand the mathematics they teach (professional knowledge for teachers of
mathematics) and need to continue to learn. One way to continue to learn is to attend
to children’s mathematical thinking (Franke, Carpenter, Levi, & Fennema, 2001).
As such teachers can learn about both mathematics and children’s mathematical
thinking by noticing children’s mathematical thinking. (For more about the construct
of noticing, see Schack, Fisher, & Wilhelm, 2017; Sherin, Jacobs, & Philipp, 2011).
First, though, we must consider the knowledge needed to teach mathematics.
Researchers studying professional knowledge for teachers of mathematics have
considered the quality of mathematical knowledge held, such as instrumental/
relational (Skemp, 1978), procedural/conceptual (Hiebert & Lefevre, 1986),
profound understanding of mathematics (Ma, 1999), key developmental
understandings (Simon, 2006), and strands of mathematical proficiency (National
Research Council, 2001). Silverman and Thompson (2008) built on Simon’s (2006)
key developmental understandings and developed a Framework for Mathematical
Knowledge of Teaching with the following categories: (a) a developed key
developmental understanding within a topic, (b) models of the variety of ways

400

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

children may understand the content (decentering), (c) an image of how someone
else might come to think of the mathematical idea in a similar way, (d) an image of
the kinds of activities and conversations about those activities that might support
another person’s development of a similar understanding of the mathematical idea,
and (e) an image of how children who have come to think about the mathematical
idea in the specified way are empowered to learn other related mathematical ideas.
With respect to quality of knowledge, researchers have focused on identifying
what type of content knowledge is held by the child rather than focusing on the
knowledge itself. The types of content knowledge have been categorized by
researchers in different ways. For example, Skemp (1978) distinguished instrumental
understanding, which enables one to explain how to apply a rule or procedure, from
relational understanding, which enables one to explain why the rule or procedure
is mathematically valid. In distinguishing between procedural and conceptual
knowledge, Hiebert and Lefevre (1986) emphasized that conceptual knowledge is
well connected whereas procedural knowledge may not be. Ma (1999) discussed
having learners develop Profound Understandings of Fundamental Mathematics,
which includes connectedness, multiple perspectives, basic ideas, and longitudinal
coherence. The National Research Council (2001) described five strands of
mathematical proficiency:
Conceptual understanding – comprehension of mathematical concepts,
operations, and relations; procedural fluency – skill in carrying out procedures
flexibly, accurately, efficiently, and appropriately; strategic competence – ability
to formulate, represent, and solve mathematical problems; adaptive reasoning
– capacity for logical thought, reflection, explanation, and justification; and
productive disposition – habitual inclination to see mathematics as sensible,
useful, and worthwhile, coupled with a belief in diligence and one’s own
efficacy. (p. 116)
Researchers studying professional knowledge for teachers of mathematics
have also considered categories of knowledge, such as content knowledge and
pedagogical content knowledge (Shulman, 1986). Ball, Thames, and Phelps (2008)
further developed this distinction by identifying Mathematical Knowledge for
Teaching (MKT), within which they subcategorized both types. Regarding content
knowledge, which they named Subject Matter Knowledge, Ball et al. (2008) included
Common Content Knowledge (the mathematics that teachers teach to students),
Specialized Content Knowledge (the mathematics that teachers do not explicitly
teach to students, but that they need to know to effectively teach), and Horizon
Knowledge (understanding how the mathematical content being taught evolved from
simpler ideas and subsequently will lead to more sophisticated mathematics). Within
pedagogical knowledge, they included Knowledge of Content and Curriculum (how
to use existing curricula to teach mathematics), Knowledge of Content and Students
(knowing how children learn), and Knowledge of Content and Teaching (knowing
ways of teaching). Some researchers have added other aspects to their frameworks:

401

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Tatto (2013) discussed the concept of general knowledge for teaching, and Baumert
and Kunter (2013) added categories such as pedagogical or psychological knowledge,
organizational knowledge, and counseling knowledge.
Finally, researchers studying professional knowledge for teachers have focused
on situations in which mathematical knowledge surfaces in teaching. Rowland
(2013) identified four dimensions of knowledge: Foundation, Transformation,
Connection, and Contingency. Foundation includes the mathematical content
knowledge and beliefs and is the only one of the dimensions that is not situational
per se. Transformation includes how to transform that knowledge for the students
and comprises actions toward the students (in planning and enactment). Connections
includes connections among concepts and thus is relevant and observable in
sequencing for mathematical coherence. Last, contingency refers to the teacher’s
reaction to unanticipated events such as a student question or contribution. Other
researchers have also considered the mathematical thinking of children as a
separate category. For example, Ernest (1989) was one of the first in the field to
consider this dimension when he included in his model of mathematical teacher
knowledge a dimension called Knowledge of the Students Taught. Baumert and
Kunter (in COACTIV, 2013) and Tatto (in TEDS-M, 2013) separated the knowledge
of children’s mathematical thinking from content knowledge. Ball et al. (2008),
in their framework, placed knowledge of content and students in the pedagogical
part of the framework and knowledge of content in the subject-matter part of the
framework. In our work, we meld those two dimensions and leverage the focus on
Children’s Mathematical Thinking as a pathway to develop and reframe subject-
matter knowledge.

EXAMPLE 2: ANDREW

Considering mathematics through the lens of children’s mathematical thinking


opens an entire new world, a world that is important for at least two reasons.
First, how children think about mathematics, whether or not their reasoning is
valid, is important for teachers to understand so that the teachers can help connect
mathematics to the children’s current understandings. However, we see a second
reason for understanding mathematics through children’s thinking: Often, children’s
ideas are mathematically valid and lead to rich and important mathematics. Figure
14.5 presents a second example of a child’s thinking, this one by Andrew, and how
grappling with this example supports prospective elementary school teachers to see a
different mathematics. Andrew, a second-grade student,2VROYHG± ͒DVVKRZQ
in Figure 14.5. We invite the readers to think about Andrew’s work and, in particular,
to consider two questions: (a) Why does Andrew’s strategy makes mathematical
VHQVH"DQG E +RZZRXOG$QGUHZVROYH± ͒XVLQJKLVUHDVRQLQJ"
We have asked prospective elementary school teachers and practicing teachers
versions of these two questions to gain insights about their mathematical
content knowledge. In analyzing Andrew’s work, teachers address foundational

402

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

Figure 14.5. Andrew’s work in solving 63 – 25

mathematical ideas for primary-grade children (in this case, whole-number concepts
and operations). Moreover, making sense of children’s mathematical thinking is
part of what teachers naturally do in the classroom. Mathematically, to analyze
Andrew’s (likely unfamiliar) strategy, teachers need to engage in problem solving,
and to explain why the strategy makes mathematical sense, teachers need to provide
justification, which requires not only understanding of the strategy but also ability
to articulate that understanding. Applying Andrew’s approach to a new problem
requires teachers to invoke a form of procedural fluency.
Of the teachers who were able to explain a mathematically valid approach to
Andrew’s strategy, most used one of the following three ways of reasoning. In the
first approach, referred to as a static DSSURDFK í2 (we suggest you read this as
“dash two”) represents negative 2, resulting when one subtracts 5 from 3. In the
second approach, referred to as a dynamic DSSURDFK WKH í2 represents 2 more to
be subtracted after one subtracts 3 of the 5 in the subtrahend (25) from the 3 in the
minuend (63). A third approach is a compensating strategy whereby one might add
2 to the 63, subtract 25 from 65 and then recognize that either negative 2 must be
added to the 40 or 2 more must be subtracted from 40 to compensate for having
DGGHGWKHWR1RWHWKDWZKHUHDVí2 is seen as negative 2 in the first approach
and as subtracting 2 in the second approach, the third approach might be associated
with either conceptualization.
We found that most prospective and practicing teachers used the first approach
to describe Andrew’s mathematical reasoning: They believed that Andrew used his
knowledge of negative numbers to solve the problem, finding first 60 – 20 = 40, then
± í UHDGDV³QHJDWLYHWZR´E\WKHSURVSHFWLYHDQGSUDFWLFLQJWHDFKHUV DQG
finally adding the results. It was only teachers who had experienced professional
development focused on children’s mathematical thinking who used either the
dynamic approach or the compensating approach (Figure 14.6). To be clear, all
three approaches are mathematically correct ways Andrew could have potentially

403

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Figure 14.6. Recomposing the minuend or subtrahend and compensating for the change to
show Andrew’s reasoning

reasoned. However, although few second graders will apply the negative-number
reasoning, many will apply one of the dynamic or compensating methods.

ANALYZING ANDREW’S SOLUTION

We analyze Andrew’s solution in three ways. First, we note that his answer, 38, was
correct. An analysis of right/wrong is quick, objective, and useful, albeit limiting.
0RUHRYHUVRPHWHDFKHUVPLJKWEHFRQIXVHGDVWRZKHWKHUíRULV$QGUHZ¶V
solution.
Second, we previously have attended to Andrew’s reasoning and noted that there
are (at least) three possible mathematical strategies he might have used in solving.
Knowing that most 7- or 8-year-old children are unlikely to use negative numbers, a
teacher attentive to children’s mathematical thinking might understand that Andrew
was more likely using either a dynamic or compensating approach.
Third, we attend to the mathematical idea that emerged from the solution: the
PHDQLQJRIíRUPRUHEURDGO\WKHPHDQLQJEHKLQGQHJDWLYHQXPEHUV)RUPXFK
of human history when numbers represented physical quantities, the idea that 3 – 5
would have an answer was absurd: How could you take away five things from a
group that only has three things in total? Indian and Islamic mathematicians had
been using the concept of negative numbers since at least the 7th century, but most
European mathematicians resisted the concept until the 17th century (Bishop et al.,
2014.) For example, English mathematician Francis Maseres wrote that negative
numbers “darken the very whole doctrines of the equations and make dark of the
things which are in their nature excessively obvious and simple” (Maseres, 1758,
S   <RXQJ FKLOGUHQ DUH OLNHO\ WR ILUVW HQFRXQWHU í DV ³ PRUH WR WDNH DZD\´
EXWWKHQDUHWDXJKWWRVHHíDVHLWKHUDSRVLWLRQRQWKHQXPEHUOLQH HJíLVWKH
position on the number line 2 units to the left of 0) or as a negative amount (e.g.,
í ZRXOG EH UHSUHVHQWHG E\ WZR UHG FKLSV ZKHUHDV  ZRXOG EH UHSUHVHQWHG E\

404

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

WZR EODFN FKLSV  0DWKHPDWLFDOO\ DOO WKHVH ZD\V RI WKLQNLQJ DERXW í DUH YDOLG
EXWIRUFKLOGUHQ DQGHYHQVRPHDGXOWV WKHVHGLIIHUHQWZD\VRIYLHZLQJíFDQIHHO
disconnected.

EXAMPLE 3: LYNN

The first example, Maria’s work, shows a common example of children’s mathematical
thinking, common because research has shown that this type of reasoning is robust
(Carpenter et al., 2015). The second example, Andrew’s work, is less common but
still robust in young children’s mathematical thinking. This third example, based on
the thinking of a child named Lynn, also arose in the context of a child’s thinking,
but we suspect that this is less common. However, when we focus on children’s
mathematical thinking, uncommon ways of reasoning can arise so often that teachers
learn to look for such creative solutions.
In February, Lynn, a second-grade student,3 was asked to solve the following
SUREOHPí͒ /\QQFRXQWHGRQKHUILQJHUVDQGWKHQJDYHDQDQVZHURI
How might Lynn have counted? (You might want to pause and think for a moment
before reading her strategy in the next paragraphs.)
We have posed this question to teachers and mathematics educators, and many
have conjectured that Lynn counted on her fingers, putting up one finger for each
number: “Negative two (put up one finger), negative one (put up a second finger),
zero (put up a third finger), one (put up a fourth finger), two (put up a fifth finger),
three (put up a sixth finger), four (put up a seventh finger),” and then she counted
the fingers that were up and answered, “seven.” This counting mistake whereby the
child starts at the first number is common.
However, that is not what Lynn did. She started by saying “negative two,” but did
not put up a finger. She first put up a finger when she said, “negative one.” The first
time she tried to solve the problem, she incorrectly counted up 4 instead of counting
until she reached 4, but she quickly realized her error and started again. On her
second attempt, she again started by saying “negative two,” and then put up a finger
for each number while she counted on, saying “negative one, negative zero, zero,
one, two, three, four.” After counting, she saw that she had raised seven fingers and
concluded that the correct solution is 7.
One might think about Lynn’s solution in several ways. Was she correct? No,
the correct answer is 6 and she answered 7. Did she use the reasoning her teacher
wanted her to use? Again, no. She used a nice counting-on strategy, but she made
the mistake of including “negative zero” in her counting, causing her solution to be
off by 1.
Finally, we might ask, “What is valid about Lynn’s reasoning and what
mathematical ideas are raised by her solution?” She understood the problem and the
numbers involved, including the negative numbers. She did not start by counting
WKHíRQKHUILQJHUV6KHXVHGRUGHUEDVHGUHDVRQLQJZKHUHE\VKHOHYHUDJHGWKH
sequential and ordered nature of numbers to reason (Bishop et al., 2014; Lamb,

405

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Bishop, Philipp, Whitacre, & Schappelle, 2018). She used one-to-one counting
principles. She recognized that the sequence of negative numbers is opposite the
sequence of positive numbers. She realized that each positive number has a mirror-
imaged negative number. She knew how to keep track on her fingers of how many
times she had increased by 1. She knew that the number of fingers she had raised
gave the solution to the problem. Most aspects of Lynn’s solution indicate that she
applied excellent mathematical reasoning.4
One particularly interesting issue with her reasoning is her use of “negative zero.”
Note that this reasoning, though incorrect, shows good mathematical reasoning.
Every number, including zero, has an additive inverse, and the inverse of any
number n is given by –n. The issue arises here because 0 is the only number that
is its own additive inverse. In other words, n±n for all real numbers except zero.
Technically, zero is neutral (neither positive nor negative) and does not have a sign
DVVRFLDWHGZLWKLWEXWEHFDXVH]HURLVLWVRZQDGGLWLYHLQYHUVHí  6RGRHV
negative zero exist? In a conversation, even several mathematics professors could
not reach consensus. Some believe that negative zero and positive zero do not exist
because zero is inherently neutral; others believe that negatives are conventionally
used to denote inverses, so negative zero is zero (as is positive zero); and still others
question what it means for any number to exist. (On a related note, a conversation
with a computer scientist provided great detail about use of both +DQGíLQZRUN
with computers.)
Remember that this entire discussion about the possible existence of negative
zero arose because we wanted to better understand what was valid about Lynn’s
mathematical reasoning. Lynn’s thinking provided a rich opportunity to grapple
with deep mathematics. Furthermore, we recognize that when thinking about
extending whole numbers to the set of integers, students have to consider ideas
of magnitude and what ‘bigger’ means when determining how to order negative
numbers (Whitacre et al., 2017). We recognize that not only is it understandable, but
we believe it reflects mathematically sophisticated thinking for a student like Lynn
WRLQFOXGHíDVSDUWRIWKDWH[WHQVLRQ,QVXPPDU\E\YLHZLQJWKLVPDWKHPDWLFDO
idea through the lens of one child’s thinking, we actually start to see the underlying
mathematics both differently and with more connections (Bishop, Lamb, Philipp,
Whitacre, & Schappelle, 2016).

ADDITIONAL SHORTER EXAMPLES OF MATHEMATICS EMERGING FROM


CHILDREN’S MATHEMATICAL THINKING

In the previous examples, we highlighted how children’s mathematical thinking


can be used to help prospective teachers better understand the mathematical
structure that children (not adults) see in mathematical tasks and how new
mathematical concepts, principles, and ideas can emerge from children’s thinking.
Those examples are, however, far from exhaustive. Across all of mathematical
content from primary and secondary education,5 one can find rich mathematics

406

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

arising from children’s ideas. For example, discussing the symmetry of positive
and negative numbers may prepare a teacher for similar issues that might arise
related to understanding the symmetry around the decimal point. Often children
know that on the left of the decimal point are values of ones, tens, hundreds, and
so on. For smaller place values, children often expect a mirror image of these
values around the decimal point, leading to the place values to the right of the
decimal being labeled as oneths, tenths, and hundredths. A teacher who focuses
on children’s mathematical thinking could then lead a discussion as to why there
is no oneths place.
Another example of mathematics emerging from children’s mathematical thinking
arose when one of the authors, Zig, working with fifth-grade children, explained
why .9 equals 1.6 One child argued that there must be a small amount between .9
and 1. When asked what that amount was, she said it might equal .01. Zig, thinking
that she was wrong, showed her other ways to justify that .9 logically should equal
1. Later, a colleague, to whom Zig reported the incident explained that the child was
not necessarily wrong and that in a field of mathematics called nonstandard analysis,
which has infinitesimals, numbers between .9 and 1 exist. This child’s question
led to a richer understanding on Zig’s part, and since then, when his prospective
elementary school teachers share that .9 does not “feel” as though it should equal
1, Zig explains that although most mathematicians adopt the standard analysis that
.9 = 1, the intuition some children (and prospective teachers) feel that leads them to
believe that .9 is less than 1 has mathematical expression in a field of mathematics.
By using the lens of children’s mathematical thinking, teachers can see new ideas
like these emerge. Moreover, they can better understand the mistakes children make
by considering how incorrect answers might have arisen from children trying to make
mathematical sense. For example, consider the use of the English word fraction. The
word fraction is often used to mean a (small) part of a whole (or an amount bigger
than 0 but less than 1); although that definition does not hold for improper fractions,
some children believe, for example, that 5/4 < 1 because all fractions are smaller than
all nonzero natural numbers. When comparing whole numbers, many children know
WKDW!EXWFRPSDULQJQHJDWLYHLQWHJHUVIRUH[DPSOHíDQGíLVPRUHGLIILFXOW
íLVLQVRPHVHQVHmore negativeWKDQíDQGLWKDVDODUJHUPDJQLWXGH OHDGLQJ
VRPHFKLOGUHQWREHOLHYHí!í )XUWKHUPRUHíKDVmore distance from zero
WKDQíEXWEHFDXVHíLVIDUWKHUWRWKHOHIWRQWKHQXPEHUOLQHWKDQíLWLVVPDOOHU
Literal symbols, such as x and n, can be frustrating to children because they can
symbolize an unknown, a constant, a variable, or a parameter. Trying to make sense
of big mathematical ideas can be confusing for children (and adults), but teachers
who are attuned to children’s mathematical thinking can help children through this
confusion by acknowledging the valid mathematics embedded within their thinking
and using it as a starting point to build better mathematical understanding.
A teacher who is open to children’s mathematical ideas encourages the children to
make conjectures, critique one another’s ideas, and, most important, ask questions.
If given a supporting environment, children can ask deep mathematical questions.

407

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

For example, part of the definition of a polygon is that it is a closed shape with
straight sides, but if one puts enough small sides on a polygon, it looks like a circle
(or other curved shape). In fact, as the number of sides on a regular n-gon goes
to infinity, the figure becomes a circle. So, is a circle a polygon? Similarly, when
children start graphing, they can draw all kinds of planar lines and find their slopes,
except for vertical lines, for which the slope is undefined. Some children, thinking of
slope as “rise over run,” might think that a vertical line has a slope of 3/0 (or n/0 for any
nonzero n), which also makes sense. So, do these lines actually have slope values, or
do we at least have a way to talk about their slopes?
We give one final example: A difficult mathematical question arose from a
first-grade student.7 The student’s teacher, Ms. Williams, stumped by the student’s
question, brought it to several mathematics educators for their advice on how to
answer it. The student said that Ms. Williams had shown him two blocks, two
pencils, two books, and two of many other objects, but now he wanted to see just
two. Could she show him two? Obviously, the student did not want to see the digit
or word for two, but we were all stumped as to what to tell the student. How does
one explain an abstract concept such as two without showing two of some concrete
object? We mathematics educators discussed this question, but none of us devised
a satisfying answer. Our best response was to use an analogy with color: A teacher
can show a student a red block, a red pencil, or a red book but cannot show that
student just red. Even though a student cannot be shown red itself, the human brain
can understand what is meant by red after seeing many examples of redness and
making the connections as to what is common among the examples. The same is
true of two: A teacher cannot show a student two by itself, but children can begin to
develop a meaning for two by comparing two blocks, two pencils, and two books
and noting what is invariant across those examples. Ms. Williams reported to us
that during a class discussion about these ideas, the student expressed wonder and
amazement that people could think about such complicated (abstract) ideas. Through
this conversation with a first-grade student, the complexity of making sense of even
small whole numbers becomes apparent.

USING CHILDREN’S MATHEMATICAL THINKING TO MATHEMATICALLY


MOTIVATE PROSPECTIVE ELEMENTARY SCHOOL TEACHERS

Helping prospective elementary school teachers to understand the elementary


mathematics they will teach is a difficult task (Ball, 1990; Ma, 1999; Sowder,
Philipp, Armstrong, & Schappelle, 1998). Even when a course is designed to engage
prospective elementary school teachers with the mathematics conceptually and
in meaningful ways, many of them resist, believing that the course is introducing
more complicated and confusing ways for them to teach. A common belief among
prospective elementary school teachers is that if they already know some specific
content, why do they need to relearn it? And if they do not already know that content,
then they cannot imagine needing to teach it to children (Philipp, 2008). “Knowing

408

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

the content” for many of them means being able to carry out a procedure in a step-by-
step manner to attain the correct result. How then can we support prospective teachers
in changing their beliefs about learning and teaching mathematics so that they will
recognize the need to understand the mathematics in a deeper and richer way?
One effective way to get prospective elementary school teachers to care about the
mathematics they will teach is to start with what they already care about: the children
(Philipp, 2008; Philipp et al., 2007). When they first start to think about teaching
children mathematics, they worry that the mathematics is too challenging. They
believe that focusing on conceptual understanding instead of simply memorizing
procedures makes the mathematics more difficult for the children. However, after
they see children engage in problem solving, either through selected videos, student
work, or actually working with children, they begin to understand the variation and
complexity in children’s solutions. They then start to care about better understanding
children’s mathematical thinking. Here, the prospective elementary school teachers
want to understand why certain mathematical strategies or incorrect solutions are
common with children and what they can do as teachers to help the children grapple
meaningfully with the mathematics. They then realize that to be prepared to support
children’s learning, they must themselves be willing to grapple with the mathematics
more deeply as well, which leads them to care about the mathematics.
Philipp et al. (2007) described these levels as Circles of Caring (Figure 14.7)
whereby prospective elementary school teachers care first about children in a
general, phenomenological way. After they begin to talk to children, their caring
expands to include children’s mathematical thinking. And finally, when they become
aware of the rich mathematics embedded in children’s mathematical thinking, their
circles of caring expand to include mathematics. The authors then tested their theory
using a large-scale, randomized, experimental study called Integrating Mathematics
and Pedagogy (IMAP). The results of IMAP showed that the prospective elementary
school teachers whose content course had a focus on children’s mathematical thinking
developed more sophisticated beliefs about teaching and learning mathematics.
Furthermore, the mathematical knowledge of the prospective elementary school
teachers whose course focused on children’s mathematical thinking improved more
than the knowledge of those whose course did not (Philipp et al., 2007).
Building on the idea that prospective elementary school teachers care first and
foremost about children, we consider another approach to increasing their valuing
the mathematics they will teach: Family Math Nights. At a Family Math Night,
prospective teachers interact with children, their parents, and their teachers in an
authentic environment. Although approaches to Family Math Night vary, in one
format, each group of prospective teachers develops and implements an inquiry-
based activity for the children (Kurz, 2011; Lachance, 2007). Before the event, the
prospective teachers get feedback from their peers, their professor, or local teachers
and refine their activity. The Family Math Night itself is held at a local school with
children, their parents, and school teachers attending. The prospective teachers
interact with the children and see how they think about mathematics; the children

409

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Figure 14.7. Circles of Caring. A model of growth, by way of children’s mathematical


thinking, from prospective elementary school teachers’ caring about children to their caring
about mathematics (Philipp et al., 2007, p. 441)

engage with enjoyable mathematical games, and the families learn mathematical
activities in which they can engage at home with the children. After the Family Math
Night, the prospective teachers reflect on the mathematical strategies the children
used and on the mathematics inherent in the activities.
Although in some teacher preparation programs prospective elementary school
teachers participate in Family Math Nights in their methods courses (e.g., Jacobbe,
Ross, & Hensberry, 2012; Kurz & Kokic, 2011), we believe that Family Math
Nights can and should be incorporated with the prospective teachers’ content
courses. Through Family Math Nights, the prospective elementary school teachers
see children authentically engaging with mathematical activities and learn the
strategies the children use to solve mathematical problems (Bofferding, Hoffman, &
Kastberg, 2016). Not only do they observe how children do and learn mathematics,
but they themselves also learn mathematics through the activities (Thanheiser,
Philipp, & Fasteen, 2014). Through working with the children and their families,
the prospective elementary school teachers can realize how creative children’s
mathematical strategies can be, how challenging understanding their strategies can
be for a teacher, and how enjoyable mathematical activities can be (Freiberg, 2004).

410

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

Thus, by observing children’s mathematical thinking in action, prospective teachers


can appreciate their need to develop deeper understanding of mathematics.

CULTURALLY RELEVANT MATHEMATICAL THINKING OF CHILDREN

Children encouraged to share their thinking in the classroom may use solutions
learned outside the classroom from a parent or friend. Such strategies may be based
on culturally specific algorithms. For example, a student from Mexico or Germany
may solve a three-digit subtraction problem as his or her parents learned to do in
their native country, as shown in Figure 14.8.
Prospective elementary school teachers who did not learn this algorithm in school,
or even those who did, are typically unsure why this way of solving the problem leads
to a correct result; they are particularly unsure of the role of 1 (circled in Figure 14.8).
A child using this algorithm may face the same uncertainty about why the algorithm
can be used, but, as with any other algorithm, may be able to use it to solve any
applicable problem. Mathematically, this algorithm yields correct answers because
one adds the same quantity to both minuend, 321, and subtrahend, 80, to leave the
difference, 241, unchanged. In this case, in the tens place, 80 is larger than 20 so 100
(or 10 tens) can be added to the 20 to result in 120 (or 12 tens). Now the subtraction
120 (or 12 tens) minus 80 (or 8 tens) results in 40 (or 4 tens). However, because 100
(or 10 tens) was added to the minuend, it must also be added to the subtrahend to keep
the difference the same; thus, 1 hundred is added to the subtrahend. The 1 represents
the same value added to both the minuend and the subtrahend.
A teacher focused on children’s thinking will not only invite children to share
such diverse algorithms in the classroom but will also incorporate them into her
teaching. Children may have learned those ways with or without making sense of
them and thus may or may not be able to explain them. In either case, allowing or
inviting children to share alternative algorithms and elevating them to be discussed
as other ways of doing mathematics enables everyone in the class to make sense of

Figure 14.8. Subtraction of 321 – 80 using an equal-addition algorithm

411

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

mathematics in more than one way. Making sense of multiple ways of operating on
numbers can lead to a more in-depth understanding of the mathematics underlying
them. Moreover, focusing on cultural approaches honors different cultures and the
identities of the children in the class.

CONCLUSIONS

The central argument put forth in this chapter is that if prospective elementary school
teachers learn mathematics simultaneously while focusing on children’s mathematical
thinking, they will be more motivated and better prepared to deeply grapple with
the mathematics. The framework we present highlights that the process of seeing
mathematics through the lens of children’s mathematical thinking promotes prospective
elementary school teachers to not only dig more deeply into the underlying conceptual
mathematical ideas but also raises new mathematical issues. Our framework builds
upon the work of others; we are not presenting new categories of knowledge for
mathematics teaching, but, rather, we are drawing upon the work of others to argue
that the order in which prospective elementary school teachers learn matters. When
they learn to see the mathematics in students’ mathematical thinking, they will be on
the road to learning how to attend to students’ thinking as teachers, which is a powerful
means by which teachers may continue to learn mathematics from their daily practice
(Franke et al., 2001). Furthermore, teachers will also be prepared to notice children’s
mathematical thinking that has a cultural basis, thereby finding connections to the
children’s funds of knowledge (González, 2005).
When teachers learn to attend to children’s mathematical thinking, they will see
reasoning that is robust and common, like Maria’s counting or using derived facts,
and they will observe less common reasoning, such as Lynn’s counting “Negative
one, negative zero, zero, one….” The process of learning to focus on the thinking
of children ought to begin during teacher preparation programs, not only so that
prospective elementary school teachers begin to see the importance of children’s
thinking but also so that they see the need to grapple with richer and even different
mathematics. In particular, when they consider children’s mathematical thinking,
issues arise that require considering mathematical connections or mathematical
ideas that would not have otherwise arisen.
Teachers must develop rich and flexible mathematical knowledge for teaching, and
many approaches have been used to support the learning of prospective elementary
school teachers. For example, one approach is historical, whereby mathematics is taught
by attending to the sequential development, including where and when mathematical
ideas were created. Another approach is curricular, whereby prospective elementary
school teachers learn mathematics while using school-based textbooks from which
they might one day be expected to teach. The most common approach focuses on
a mathematics-only approach, whereby they are taught or retaught the mathematics

412

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

that they will be expected to teach. The mathematical approach takes several forms,
and whereas in the past the focus was commonly on procedural mathematics, today
mathematics-for-elementary-school-teachers courses are focused upon a richer
integration of concepts, procedures, and ways of reasoning. Our goal in this chapter
is to propose a mathematical approach that weaves within it a focus on children’s
mathematical thinking. To illustrate the path that integrates mathematics with children’s
mathematical thinking, we present the metaphor of roads to mathematical knowledge
for teaching. All paths lead to mathematical knowledge for teaching, but we suggest
that various paths lead to higher elevation lookout points that offer better views of
the mathematical landscape. We are not arguing that the path that addresses deep
mathematics together with a focus on children’s mathematical thinking is the best;
rather, we are arguing that it offers a promising means by which to prepare teachers for
the work of teaching because by seeing and learning to look for the mathematical issues
embedded in children’s thinking, teachers may develop a means by which to learn
mathematics from practice. If we consider mathematical knowledge for teaching to be
a landscape, then metaphorically, the path that integrates mathematics with children’s
mathematical thinking leads to a high elevation, providing an encompassing, flexible,
and consequently applicable view of the terrain (see Figure 14.9).
Teaching is complex and no teacher preparation program can prepare elementary
school teachers for all that they need to know to effectively teach mathematics
over a career. Consequently, teachers must be poised to learn about mathematics
and about children’s mathematical thinking during teaching, which requires both a
stance toward one’s professional responsibility and a set of skills for learning from

Figure 14.9. Many paths exist for developing mathematical knowledge for teaching,
represented in this figure by the paths ending with arrows. Some paths provide higher views
offering a more comprehensive view of the landscape, and others do not. The path that
combines mathematics with children’s mathematical thinking, represented by the solid path,
is a particularly productive path yielding a comprehensive view of the landscape

413

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

practice. No royal road to preparing teacher candidates to learn from teaching exists,
but mathematics teacher educators must never avoid a problem just because it has
no simple solution. In this chapter, we propose that one path by which teachers
may learn from practice is to learn from children’s thinking, and we contend that
we can start prospective elementary school teachers down this path to the land of
mathematical knowledge for teaching while they are still learning mathematics.
Of course, such a focus will not, in and of itself, solve the intense demands for
mathematical knowledge for teaching, but we contend that this approach will
provide a productive start.

NOTES
1
Although we believe that the ideas put forth in this chapter may apply to prospective teachers of
children of all grades between kindergarten and high school, we focus in this chapter on the preparation
of teachers of children ages 5-11 who, in the United States, are referred to as elementary school
teachers.
2
This example comes from the United States, where children in the second grade are generally 7 or 8
years of age.
3
Again, the example comes from the United States, where children in the second grade are generally 7
or 8 years of age. Note, Lynn is a pseudonym.
4
Holding a deep understanding of the mathematics is a necessary, but not sufficient, condition for a
teacher to be able to question students so as to reveal the reasoning in this and other examples in this
chapter. Expertise in interviewing is also necessary, a pedagogical skill that develops over many years
of practice.
5
In the United States, these grades would be called K–12. Using the International Standard Classification
of Education, primary and secondary education would be referred to as Education Levels 1–3.
6
Here, the repeating decimal .9999… is written as .9 . We acknowledge that other notations, such as

.9 , .9 , or .(9), are used around the world.
7
This example comes from the United States, where children in the first grade are generally 6 or 7 years
of age.

REFERENCES
Association of Mathematics Teacher Educators. (2017). Standards for preparing teachers of mathematics.
Retrieved from Association of Mathematics Teacher Educators (AMTE) website, members only:
https://amte.net/standards
Ball, D. L. (1990). Breaking with experience in learning to teach mathematics: The role of a preservice
methods course. For the Learning of Mathematics, 10(2), 10–16.
Ball, D. L., Thames, M. H., & Phelps, G. (2008). Content knowledge for teaching: What makes it special?
Journal of Teacher Education, 59(5), 389–407.
Baumert, J., & Kunter, M. (2013). The COACTIV model of teachers’ professional competence. In M.
Kunter, J. Baumert, W. Blum, U. Klusmann, S. Krauss, & M. Neubrand (Eds.), Cognitive activation
in the mathematics classroom and professional competence of teachers. Results from the COACTIV
Project (pp. 25–48). New York, NY: Springer.
Bishop, J. P., Lamb, L. L., Philipp, R. A., Whitacre, I., & Schappelle, B. P. (2016). Unlocking the structure
of positive and negative numbers. Mathematics Teaching in the Middle School, 22(2), 84–91.
Bishop, J. P., Lamb, L. L., Philipp, R. A., Whitacre, I., Schappelle, B. P., & Lewis, M. L. (2014).
Obstacles and affordances for integer reasoning: An analysis of children’s thinking and the history of
mathematics. Journal for Research in Mathematics Education, 45(1), 19–61.

414

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

Bofferding, L., Hoffman, A., & Kastberg, S. (2016). Family mathematics nights: An opportunity to
improve preservice teachers’ understanding of parents’ roles and expectations. School Science and
Mathematics, 116(1), 17–28.
Browning, C., Thanheiser, E., Edson, A. J., Kimani, P., Olanoff, D., Tobias, J., & Whitacre, I. (2014).
Prospective elementary teacher mathematics content knowledge: An introduction. The Mathematics
Enthusiast, 11(2), 203–216.
Carpenter, T. P., Fennema, E., Franke, M. L., Levi, L., & Empson, S. B. (2015). Children’s mathematics:
Cognitively guided instruction (2nd ed.). Portsmouth, NH: Heinemann.
Ernest, P. (1989). The knowledge, beliefs, and attitudes of the mathematics teacher: A model. Journal of
Education for Teaching, 15(1), 13–34.
Franke, M. L., Carpenter, T. P., Levi, L., & Fennema, E. (2001). Capturing teachers’ generative change:
A follow-up study of professional development in mathematics. American Educational Research
Journal, 38, 653–689.
Freiberg, M. (2004). Getting everyone involved in family math. The Mathematics Educator, 14(1), 35–41.
González, N. (2005). Beyond culture: The hybridity of funds of knowledge. In N. González, L. Moll,
& C. Amanti (Eds.), Funds of knowledge: Theorizing practices in households, communities, and
classrooms (pp. 29–46). Mahwah, NJ: Erlbaum.
Hiebert, J., & Lefevre, P. (1986). Conceptual and procedural knowledge in mathematics: An introductory
analysis. In J. Hiebert (Ed.), Conceptual and procedural knowledge: The case of mathematics (pp.
1–27). Hillsdale, NJ: Erlbaum.
Jacobbe, T., Ross, D. D., & Hensberry, K. K. R. (2012). The effects of a family math night on preservice
teachers’ perceptions of parental involvement. Urban Education, 47(6), 1160–1182.
Jacobs, V. R., Lamb, L. L. C., & Philipp, R. A. (2010). Professional noticing of children’s mathematical
thinking. Journal for Research in Mathematics Education, 41(2), 169–202.
Kunter, M., Baumert, J., Blum, W., Klusmann, U., Krauss, S., & Neubrand, M. (2013). Cognitive
activation in the mathematics classroom and professional competence of teachers. Results from the
COACTIV project. New York, NY: Springer Science & Business Media.
Kurz, T. (2011). Establishing field-based learning by incorporating family math night into a mathematics
methodology course. Problems, Resources, and Issues in Mathematics Undergraduate Studies, 21(3),
225–237.
Kurz, T. L., & Kokic, I. B. (2011). Preservice teachers’ observations of children’s learning during family
math night. Journal of Research in Education, 21(2), 24–36.
Lachance, A. (2007). Family math nights: Collaborative celebrations of mathematical learning. Teaching
Children Mathematics, 13(8), 404–408.
Lamb, L. L., Bishop, J. P., Philipp, R. A., Whitacre, I., & Schappelle, B. P. (2018). A cross-sectional
investigation of students’ reasoning about integer addition and subtraction: Ways of reasoning,
problem types, and flexibility. Journal for Research in Mathematics Education, 49(5), 575–613.
Ma, L. (1999). Knowing and teaching elementary mathematics: Teachers’ understanding of fundamental
mathematics in China and the United States. Mahwah, NJ: Erlbaum.
Maseres, F. (1758). A dissertation on the use of the negative sign in algebra: Containing a demonstration
of the rules usually given concerning it; and shewing how quadratic and cubic equations may be
explained, without the consideration of negative roots. To which is added, as an appendix, Mr.
Machin’s quadrature of the circle. London: Samuel Richardson.
National Research Council. (Ed.). (2001). Adding it up: Helping children learn mathematics. Washington,
DC: National Academy Press.
Neubrand, M. (2018). Conceptualizations of professional knowledge for teachers of mathematics. ZDM
Mathematics Education, 50(4), 601–612.
Olanoff, D., Lo, J., & Tobias, J. (2014). Mathematical content knowledge for teaching elementary
mathematics: A focus on fractions. The Mathematics Enthusiast, 11(2), 267–310.
Pesek, D. D., & Kirshner, D. (2000). Interference of instrumental instruction in subsequent relational
learning. Journal for Research in Mathematics Education, 31(5), 524–540.

415

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
RANDOLPH A. PHILIPP ET AL.

Philipp, R. A. (2007). Mathematics teachers’ beliefs and affect. In F. Lester (Ed.), Second handbook
of research on mathematics teaching and learning (pp. 257–315). Reston, VA: National Council of
Teachers of Mathematics.
Philipp, R. A. (2008). Motivating prospective elementary school teachers to learn mathematics by
focusing upon children’s mathematical thinking. Issues in Teacher Education, 27(2), 195–210.
Philipp, R. A., Ambrose, R., Lamb, L. C., Sowder, J. T., Schappelle, B. P., Sowder, L., … Chauvot,
J. (2007). Effects of early field experiences on the mathematical content knowledge and beliefs of
prospective elementary school teachers: An experimental study. Journal for Research in Mathematics
Education, 38(5), 438–476.
Rowland, T. (2013). The Knowledge Quartet: The genesis and application of a framework for analyzing
mathematics teaching and deepening teachers’ mathematics knowledge. Journal of Education, 1(3),
15–43.
Schack, E. O., Fisher, M. H., & Wilhelm, J. H. (2017). Teacher noticing: Bridging and broadening
perspectives, contexts, and frameworks. New York, NY: Springer International.
Sherin, M. G., Jacobs, V. R., & Philipp, R. A. (Eds.). (2011). Mathematics teacher noticing: Seeing
through teachers’ eyes. New York, NY: Routledge.
Shulman, L. S. (1986). Those who understand: Knowledge growth in teaching. Educational Researcher,
15(2), 4–14.
Shulman, L. S. (1998). Theory, practice, and the education of professionals. The Elementary School
Journal, 98(5), 511–526.
Silverman, J., & Thompson, P. (2008). Toward a framework for the development of mathematical
knowledge for teaching. Journal of Mathematics Teacher Education, 11(6), 499–511. doi:10.1007/
s10857-008-9089-5
Simon, M. A. (2006). Key developmental understandings in mathematics: A direction for investigating
and establishing learning goals. Mathematical Thinking and Learning, 8(4), 359–371.
Skemp, R. R. (1978). Relational understanding and instrumental understanding. Arithmetic Teacher, 26,
9–15.
Smith, M. S., & Stein, M. K. (2011). 5 practices for orchestrating mathematics discussions. Reston, VA:
National Council of Teachers of Mathematics.
Sowder, J. T., Philipp, R. A., Armstrong, B. E., & Schappelle, B. P. (1998). Middle-grade teachers’
mathematical knowledge and its relationship to instruction. New York, NY: State University of New
York Press.
Stanic, G. M. A., & Kilpatrick, J. (Eds.). (2003). A history of school mathematics (Vol. 1). Reston, VA:
National Council of Teachers of Mathematics.
Stigler, J. W., & Hiebert, J. (1999). The teaching gap. New York, NY: The Free Press.
Tatto, M. T. (2013). The Teacher Education and Development Study in Mathematics (TEDS-M): Policy,
practice, and readiness to teach primary and secondary mathematics in 17 countries (Technical
Report). Amsterdam: IEA.
Thanheiser, E., & Browning, C. (2014). Preface. The Mathematics Enthusiast, 11(2), 201–202.
Thanheiser, E., Browning, C., Edson, A. J., Lo, J., Whitacre, I., Olanoff, D., & Morton, C. (2014).
Prospective elementary mathematics teacher content knowledge: What do we know, what do we not
know, and where do we go? The Mathematics Enthusiast, 11(2), 433–448.
Thanheiser, E., Philipp, R., & Fasteen, J. (2014). Motivating prospective elementary teachers to learn
mathematics via an authentic task. In P. Liljedahl, C. Nicol, S. Oesterle, & D. Allan (Eds.), Proceedings
of the 38th conference of the International Group for the Psychology of mathematics education (Vol.
5, pp. 233–240). Vancouver: PME.
Thanheiser, E., Whitacre, I., & Roy, G. (2014). Mathematical content knowledge for teaching elementary
mathematics: A focus on whole-number concepts and operations. The Mathematics Enthusiast, 11(2),
217–266.
Thompson, A. G. (1992). Teachers’ beliefs and conceptions: A synthesis of the research. In D. A. Grouws
(Ed.), Handbook of research on mathematics teaching and learning (pp. 127–146). New York, NY:
Macmillan.
Thompson, A. G., Philipp, R. A., Thompson, P. W., & Boyd, B. A. (1994). Calculational and conceptual
orientations in teaching mathematics. In D. B. Aichele & A. F. Coxford (Eds.), Professional

416

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
SEEING THROUGH CHILDREN’S MATHEMATICAL THINKING

development for teachers of mathematics (pp. 79–92). Reston, VA: National Council of Teachers of
Mathematics.
Van de Walle, J. A., Karp, K. S., Lovin, L. H., & Bay-Williams, J. M. (2014). Teaching student-centered
mathematics: Developmentally appropriate instruction for Grades K–2 (2nd ed.). New York, NY:
Pearson.
Whitacre, I., Azuz, B., Lamb, L. L., Bishop, J. P., Schappelle, B. P., & Philipp, R. A. (2017). Integer
comparisons across the grades: Students’ justifications and ways of reasoning. Journal of Mathematical
Behavior, 45, 47–62.

Randolph A. Philipp
School of Teacher Education
San Diego State University

John (Zig) Siegfried


Department of Mathematics and Statistics
James Madison University

Eva Thanheiser
The Fariborz Maseeh Department of Mathematics and Statistics
Portland State University

417

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
INDEX

A culturally sustaining, 244, 249–254,


activity theory, 8, 284, 303, 305, 258, 260, 262, 263, 265, 266
313–322
approximations of practice, 356, D
372–377, 380 decision making, 8, 114, 123, 190, 193,
authentic assessment, 3, 5, 43–72 225, 303–323
authentic intellectual quality, 43, 48, decisions in lesson planning, 303–323
51–54, 61–70
E
B equity, 7, 8, 243–266, 359, 369, 370
beliefs, vii, 1–10, 18, 21, 22, 27, 37, 57, exemplification, 2, 8, 9, 329, 330–332,
60, 61, 70, 79, 83, 114–116, 120, 334–341, 344, 348–350
131–151, 155–159, 161, 162,
165, 166, 169–173, 175–177, F
185–207, 213, 219, 232, 248, fraction division, 77, 84–88, 90–99,
253, 254, 263, 265, 282, 304, 101, 102, 196, 198, 202, 395, 396
310–312, 358, 369, 394
G
C grounded theory, 105, 110, 114, 117,
children’s mathematical thinking, 166, 315
391–414
China, 2, 5, 47, 77, 79, 85, 90–92, 95, H
98–102, 121, 137, 162, 371 helping prospective teachers, 68, 85,
Chinese lesson study, 133, 134, 143, 151 102, 372, 373, 379, 408
cognitive demand, 5, 44, 48, 50–51,
54–57, 68, 69, 83, 115, 116, 120, I
191, 194, 204, 205 identity, vii, 1–10, 45, 211–235, 252,
competency framework, 49, 57 263, 299, 360
conceptual knowledge, 5, 77–102, 401 inclusive, 7, 146, 244, 246–249, 258,
core/high leverage practices, 9, 331, 260, 262, 265, 266, 358, 371
349, 355, 359, 361–369, 373, inquiry-based practice, 8, 275–300
375, 378 in-the-moment decisions, 114, 306,
creativity, 6, 22, 155–177 310–312
cultural beliefs, 131–151
cultural transposition, 131, 138–141, J
150 Japanese lesson study, 133, 140, 142

419

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
INDEX

K O
knowledge package, 26, 79–82, 84 out-of-field teaching, 7, 218–224, 230,
knowledge quartet, 5–7, 78, 83, 105, 231, 234
113–124, 187, 189, 190, 213, 215,
307–309 P
pedagogical content knowledge, 4, 5,
L 15, 26, 27, 56, 78, 80, 82, 83,
lesson planning, 19, 146, 149, 303, 314 105–107, 111–114, 122–125,
lesson study, 2, 6, 37, 57, 131–135, 137, 185–207, 211, 212, 215, 224, 228,
138–151, 309, 332, 338, 341, 229, 231, 232, 234, 350, 391, 401
344, 345, 350, 356, 374 practice-based pedagogies, 2, 9,
355–380
M practicing teachers’ beliefs, 6, 186, 190,
mathematical creativity, 3, 6, 155–177 204
mathematical knowledge for teaching, professional development, 2–9, 16, 29,
5, 20, 23, 25–27, 56, 63, 67, 70, 34, 37, 43, 56, 57, 61, 69, 71, 83,
78, 107, 123, 189, 214–216, 308, 139, 155, 156, 170–177, 195,
329, 350, 391, 399, 401, 412–414 202, 207, 221, 228–231, 233,
mathematical tasks, 3, 5, 30, 43–45, 48, 248, 249, 255, 256, 258, 262,
50, 56, 57, 59, 70–72, 176, 185, 277, 284, 309, 314, 322, 329,
189, 295, 297, 339, 349, 406 330–332, 339, 341, 349, 355,
mathematical thinking, 2, 9, 43–46, 49, 356, 374, 403
50, 69, 245, 287, 288, 307, 311, professional knowledge, 7, 9, 211–235,
391–414 304, 307, 399, 400–402
mathematics classroom, 3, 6, 8, 9, 54, prospective teachers, 5, 9, 16, 17, 19,
59, 60, 67, 69, 78, 83, 107–113, 24, 55, 59–61, 63–68, 71, 78–81,
156, 159, 165, 169, 206, 215, 83–86, 88, 89, 92–102, 106–113,
225, 245, 246, 248–250, 252, 115, 117, 121, 123, 124, 134, 142,
258, 261–263, 334, 376 147–150, 159, 161, 162, 170–173,
mathematics conceptual knowledge for 176, 204, 231, 261, 306, 309,
teaching, 5, 77–102 336, 337, 361, 362, 364, 369,
mathematics knowledge for teaching, 373, 375–379, 391, 399, 406,
21, 36, 66, 67, 69, 78, 80–82, 84, 409–411, 414
106, 135, 173–175, 212–216 prospective teachers’ beliefs, 147,
mathematics teacher preparation, 44, 131–151
77, 78, 80, 83, 95, 99–102, 106,
398, 410, 412–414 S
secondary teacher education, 172, 196,
N 255, 336, 337
non-specialist teachers of mathematics, social dimensions on decision making,
7, 211–235 311–313, 317, 322
numeracy, 7, 17, 19, 22, 23, 37, 114, social justice, 8, 243, 244, 250,
211, 212, 225–235 254–260, 262, 265, 266

420

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London
INDEX

South Korea, 5, 77, 79, 90, 91, 95, 96, teacher identity, 1–10, 45, 211–235,
98–102 252, 263, 299, 360
teacher knowledge, vii, 1–10, 17, 27,
T 37, 44, 78, 84, 85, 98, 106, 109,
tasks, 3, 5–7, 9, 25, 26, 30, 43–72, 114, 122, 187, 188, 190, 193, 206,
83, 84, 87, 115, 120, 121, 123, 212–216, 224, 233, 234, 257,
136, 141, 148, 156, 163, 164, 303, 304, 307–308, 310, 321,
166, 171–177, 185, 189, 191, 336, 373, 402
192, 194, 199, 204, 205, 214, teachers’ beliefs, 3, 132, 134, 144, 147,
215, 227, 228, 232, 234, 155, 156–159, 162, 165, 169,
255–257, 259, 261, 262, 281, 170, 171, 175, 177, 185–207,
286–298, 300, 306, 308, 311, 316, 248, 253, 304, 311, 358, 394
318, 332, 334, 336–350, 359, teaching practices, 1, 4, 7–10, 25, 142,
360, 364, 369–372, 392, 394–396, 228–230, 235, 244, 247, 248,
406, 408 250, 256, 258, 262, 265, 276,
teacher decision making, 8, 114, 123, 278, 279, 282–284, 286, 289,
190, 193, 303–323 291, 295, 307, 311, 312, 315, 319,
teacher education, vii, 1–10, 19, 43, 322, 331, 332, 335, 339, 340,
44, 51, 57, 64, 68, 69, 80, 81, 355–380
95, 100, 106, 108, 113, 121,
124, 131–134, 141, 147–148, U
150, 173, 175, 185, 202, 207, university mathematics teaching, 1–10,
211, 212, 220–223, 226, 231, 275–300
233, 243, 245, 248, 263, 276,
305, 309, 322, 323, 329–350, V
355, 356, 359, 360, 363, 364, variation, 8, 9, 25, 77, 80, 84, 87, 100,
366, 372, 373–376, 379, 101, 214, 220, 329–350, 362,
380, 392 377, 409

421

- 978-90-04-41887-5
Downloaded from Brill.com 04/03/2024 05:15:13PM
via University College London

You might also like