You are on page 1of 50

Chapter 3

The Mathematics Lecture


3.1 The Purpose of Lectures in Mathematics
In Chapter 2 we looked at curriculum design based on eleven basic principles of teaching underpinned
by accepted theories of learning of mathematics. We considered expression of content in terms of aims
and objectives, design of teaching strategies to meet the objectives, and nally the basics of preparation of
learning materials for a range of teaching activities. This really served to assemble the fundamental ideas
that we need to consider when engaging in any sort of teaching activity. In mathematics the primary form
of teaching activity is currently the lecture supplemented by various types of tutorial and independent
work set for the student. In this chapter we look in particular at the mathematics lecture. Of course the
greater part of the effectiveness of a lecture resides in its planning and preparation. In Chapter 2, under
curriculum design we covered most of the general issues surrounding such preparation, whether it be
for a whole module or a particular lecture. So we do not need to repeat this for planning and preparing a
lecture, but will simply summarize the main points in this chapter. This chapter concentrates mainly on
the delivery and evaluation of the lecture. This is underpinned by the same basic principles of Section 2.4
and we will occasionally highlight where these are particularly relevant.
Denition of a lecture
The lecture is still the most common form of contact teaching in HE mathematics for anything but small
groups (See Chapter 4). And of course there is a large range of approaches to such lectures. To accommo-
date this range we will take a fairly precise denition of a lecture:
A mathematics lecture is a contact period during which the main activity is a carefully paced presen-
tation, by the lecturer, of a particular mathematical topic, with the intention that when the students
leave they will have the means, materials and incentive to study and to learn that topic to the desired
level.
Note that this denition does not imply a monologue from the lecturer, or absence of interaction with the
students.
We are here concerned with how we can give good mathematics lectures that efciently and effectively
support student learning. For this we will put together advice from experienced practitioners, results of
educational research, commonly accepted good practice, and the views of students. Throughout, many
examples are drawn from experienced practitioners, giving a range of perspectives.
53
54
The purpose of lectures
Of course, the purpose of any lecture depends on the topic, the students and more besides. Properly
designed a lecture can achieve a wide range of educational purposes. The really important thing is that
the lecturer is clear about what they wish to achieve, and that they communicate this to the students.
This is expressed in terms of the aims and objectives (i.e. learning objectives) of the lecture (Section 2.6).
Normally the object of the lecture is to convey some important concept or technique, but it can also be
used to good effect to inspire and motivate students, to impart values and change attitudes, encourage
learning skills and develop transferable skills.
The purpose of a particular lecture may perhaps be best appreciated when it is realised that lecturing time
is actually prime time. It is the one occasion on which the lecturer can convey ideas, concepts, attitudes,
etc to all of the students at once and to get across the key messages in the subject. At other times students
will usually be learning independently or in small groups, supported by materials that we provide for
them. Lecture time should therefore be quality time. Lectures may most productively be aimed at the
key messages in the subject, at those areas where students are likely to experience the most difculty, at
establishing and encouraging higher order thinking skills, at developing attitudes and approaches that
characterise the subject, and at motivating the students. We might summarise this as:
The key purpose of a lecture is to get across the most important messages about our subject to the
maximum number of students
The importance of lectures
Not only is the lecture prime time for conveying messages of all kinds to the students, but it is also a shop
window for our subject and our community. So, lectures are important, because they can:
be key learning experiences for the students
motivate and inspire them
inuence students feelings about the subject
show students what mathematicians do and how they work
promote student condence and self esteem
convey the values of mathematics to students
now cost students a great deal of money.
At the same time, we know that lectures are not absolutely essential in mathematics teaching, and many
successful mathematicians have managed very well without them. However, even the best and most
independent student can benet from a good lecture.
Exercise
Discuss the advantages and disadvantages of lectures. Which lectures did you learn most/least from
as a student, and why? What do you think your students get from your lectures?
55
3.2 Content and Objectives of the Lecture
In Sections 2.5, 2.6 we discussed content of the curriculum in general, and how it is expressed in terms of
learning objectives. The ideas and examples given there are applicable to planning and preparing for a
specic lecture, so here we will simply summarise the points as they would apply to giving a particular
lecture, and for more detail you can refer back to Chapter 2. We have retained similar exercises, but
applied to the individual lecture, and these might be used as preparation for forthcoming lectures. As
discussed in Section 2.5 for the curriculum in general, the content and objectives of a lecture should:
be made clear to the students (P3)
t coherently within the context of the course (P7)
progress at a reasonable pace in both time scale and intellectual demands.(P5)
take account of the mathematical background of the students (P5)
For more on these see Section 2.5, or work though the following exercises.
Exercises
1. Think of a lecture you are soon to give and write a few lines that will inform the students of
the content and purpose of the lecture.
2. Choose a particular topic from one of your lectures. What is the content and purpose of that
topic in the overall course/module? What does it rely on and where does it lead?
3. Break down one of your modules into say 25 lectures, each covering a specic topic(s)or
objectives. Write a concise description of the content and objective of each lecture, and sketch
out how the different lectures are linked together.
4. Choose a particular lecture you are soon to give and analyse its pre-requisites. What
knowledge and skills will the students need to cope with this, and what level of facility do they
need in these skills? Are there likely to be any students who lack this facility, and how can you
address this?
3.3 The Teaching and Learning Strategy for the Lecture
Again, designing teaching and learning strategies to meet the aims and objectives of a particular learning
activity are considered in Section 2.7. In designing the teaching and learning strategy and the learning
activities for a given lecture we might need to think about:
the resources available (P1)
the students abilities and background (P5)
alignment with the learning objectives (P4)
ways of engaging the students in active learning behaviour (P10).
56
Refer to Section 2.7 for more details. The following example is perhaps an extreme example that at least
shows what is possible.
Example
Millet [55] describes a very involved teaching and learning strategy designed for lectures on
a Precalculus course for service students. In it activities were designed precisely to engage
student activity in a way that would encourage learning. The 75-minute lecture was not
what we would normally regard as such! The format, outlined below, is very interactive.
1. An initial segment of administrative information followed by an invitation for student
questions
2. A segment devoted to discussion of problems that had arisen from previous weekly
coursework
3. A group surprise quiz came next, based on reading previously assigned for the lecture,
culminating in a class discussion of results which moved the class onto new material
(about 20 minutes into the lecture)
4. The fourth and longest segment concerned new concepts and methods growing out of
the surprise quiz, and developed in a manner based on the previous discussion. This
comprised no more than one or two new key elements as the focus of the work
5. Time permitting any secondary new material was incorporated in the presentation and
discussion of a closing issue
6. In closing students were reminded of homework, quizzes, material to be reviewed, etc.
Of course, we do not advocate that all lectures are planned in such meticulous detail, but in fact as
you develop experience you nd yourself automatically thinking ahead about such things. Again, the
exercises below can provide you with practice in this area.
Exercises
1. Itemise the resources necessary for a few of your lectures. Include your contact time,
preparation time (including thinking time!), costs of any materials, student time,
accommodation facilities, equipment, etc. Now consider any efciencies that can be made,
without compromising the learning objectives.
2. Think of a particular lecture you are about to give. Are the students ready for the proposed
teaching method and associated activities? When you cover a particular point, will they be
able to cope with your delivery method - will a straight lecturing mode sufce, or will they
need more interactive and discursive explanation? Are they used to doing exercises unaided in
class?
3. Consider a number of lectures you might be giving soon and design appropriate teaching and
learning activities to meet the various objectives of the lectures.
4. Think about your next lecture and plan how you might introduce one or two short periods of
activity to emphasize particular points. Aim to deepen the students understanding, without
necessarily sacricing coverage of material.
57
3.4 Preparation of the Learning Material for a Lecture
Media and message in the lecture
In Sections 2.8 and 2.9 we considered the issues behind preparing learning materials in general, splitting
the discussion into the media and the message. The application to a particular lecture is straightforward,
and the exercises will provide practice if needed. In a typical lecture the main materials needed will be:
the students notes, either transcribed from your presentation or in a handout, or electronic le
problem/exercise sheets
occasional administrative stuff such as timetables
ongoing work such as coursework and associated solutions and feedback.
As emphasised in Chapter 2 remember that preparation takes longer than you expect, especially for the
rst delivery of a lecture, the writing style you use is important, and the volume of learning materials
should be reasonable for the length of the lecture. Also the real bulk of preparation time lies in the
intellectual construction of what you are going to say to the students, how you will express it, what
exercises you will give, how to phrase things, what depth of proof to use - the message.
Design of learning materials for the lecture - media
The list below summarises the sorts of things we have to think about when preparing teaching materials
for a lecture. See Section 2.8 for details, or work through the exercises below.
The alignment of the materials to the teaching and learning activities of the course
The clarity and conciseness of the materials and the style of presentation
The volume of learning materials.
Exercises
1. Outline the learning materials you will use for a selection of your lectures and explain how
these will support the achievement of the learning objectives
2. Check your lecture materials for clarity and conciseness. Can they be improved? Can you
reduce the number of pieces of paper?
3. Consider one of your lectures. At the end of the lecture, how much material will the students
have? Does it seem about right? Compare with other modules.
Design of learning materials for a lecture - message
Again, this is simply an application of Section 2.9 to the case of a particular lecture. Also note that you
might not be able to incorporate all the suggestions here in your rst lectures, but gradually they can be
absorbed until you do them automatically as the need arises. Based on the eleven principles of Section
2.4 we suggest the following guidelines for the learning materials for a lecture. They should:
58
be appropriate to the students needs and level of understanding of the topic
support the achievement of the objectives
highlight/reinforce the key ideas
provide hooks for students to hang ideas on - aide memoires, mnenomics, etc, to make it as easy as
possible to remember and internalize key results and ideas
anticipate the difculties students might encounter and support them in dealing with these
highlight the essence of complicated ideas, theorems, methods, etc and support the development of
students intuitive views of formal arguments
provide roadmaps and overviews of difcult sequences of arguments
include devices for developing rapport with the students
For more details see Section 2.9, and if needed work through the following exercises.
Exercises
1. Check your lecture material for each lecture. Do the notes, exercise sheets, etc take account of
the students background - will the language and notation used be familiar to them, both in
terms of content and mathematical maturity?
2. Choose a typical rst year topic such as partial fractions and prepare a succinct non-technical
(i.e. accessible to those not yet familiar with the topic) summary of the topic and its relevance.
3. In lectures you are about to give what are some of the most difcult ideas to get across - where
do students have the most difculty? Why is it a difculty and how can it be addressed?
4. Choosing examples from some of your own lectures, bring out the essential messages of
particular ideas, expressing them in the simplest, most incisive way with a minimum of
technical language and possible use of analogies.
5. Choose a number of key topics from some of your lectures and write succinct overviews and
directions through difcult sequences of arguments. Dont forget that while they must be
concise and succinct, they must also be in terms that the students can understand.
3.5 MATHEMATICS for the Lecture
It is useful to have a checklist to run through before the lecture to avoid last minute hiccups and get you
in the mood. If this is the rst time you have given this particular lecture then you may spend some
time rehearsing it and getting your ideas together, even if everything is fully prepared. Most of us nd
it stressful in front of a class, and do not always function at our best, so we really should leave nothing
to chance. Indeed a part of your last minute run through may be just to calm your nerves and bolster
your condence. For a run through of what you might consider when about to give a lecture, we can use
MATHEMATICS (Section 2.2):
59
Mathematical content
Aims and objectives of the curriculum
Teaching and learning activities to meet the aims and objectives
Help to be provided to the students - support and guidance
Evaluation, management and administration of the curriculum and its delivery
Materials to support the curriculum
Assessment of the students
Time considerations and scheduling
Initial position of the students - where we are starting from
Coherence of the curriculum - how the different topics t together
Students.
Obviously, the mathematical content of the lecture is important, and although we dont usually have
lesson plans we do have to think about the aims and objectives of the lecture. Read through your notes
to remind yourself. You should of course know the subject matter inside out by now, but refresh your
thoughts on it. Usually the teaching and learning activity in a lecture consists of the students listening
to what you have to say, or watching you work, but of course there are many ways of engaging them in
fruitful activity and maybe today you might try something new. Practice your delivery, if only mentally.
Help and support for the students may simply amount to directing them to a section of a text, or giving
a handout on a difcult topic, or taking particular care in explaining a particular point. Maybe it is time
to give out a student questionnaire, but in any case you should be continually evaluating how the class
is going. And what materials do you need for this lecture - a problem sheet, a few OHPs? Organise
them for easy access? And in materials include the accommodation and facilities of the lecture room -
particularly chalk/pens/eraser! What about temperature, noise levels, over/under crowding, boards,
screen, and other AV - order equipment needed in good time. In terms of assessment, do you need to
hand out/collect in coursework or tell them something about the examination to come? Do you have
time to cover the next topic today, have you got to tell them the examination dates? Normally, if you are
following on from previous lectures, you will know the initial position of (most of) the students, but it
is a good idea to remind them where you got to last time. And if it is the rst class of the session then
you may even need some sort of review. In fact does it follow on from the last lecture, or is it a new
topic altogether? If it is then you have to explain the links to the students, to ensure the coherence of the
module. Finally, remember why you are there - for the students. Think about what is best for them, see
any students that you need to, work to establish a rapport with them.
Of course, we may have to think about other things as well, but this provides at least a basic minimal set.
In fact, if you construct a table with the MATHEMATICS letters as row headings then you can scribble
in relevant notes for the lecture which effectively becomes your lesson plan, as well as a record of how
the lecture went for later. Of course, if none of this works for you, then devise your own checklist.
Exercise
Work through the MATHEMATICS checklist for your next lecture. Did it cover most eventualities?
If there is anything missing, make a note of it for next time.
3.6 Presenting the Lecture - an Overview
The next four sections are about actually giving the lecture. For many of us, particularly in our early years,
this is something of a trial. While we hope this book will help you to prepare for this, it will probably be
a long time, if ever, before you feel entirely relaxed when you go into the classroom. It will help in your
60
early lectures if you can discuss your experiences and problems with an experienced colleague. Your
department will normally provide a mentor for this purpose.
In giving a lecture we have a wide range of things to consider, even when we have prepared all the
materials we need. First there is the basic classroom technique including establishing a professional
presence in the classroom, running the lecture to ensure the objectives are met, and structuring it for
maximum impact, starting and nishing appropriately. There are the mundane technical issues of using
the various resources in ways that will aid the learning process. Then there is the very important task of
producing the right atmosphere in the classroom- creating a warm, relaxed, but industrious environment
in which people want to learn, enjoy the activity and feel free to get involved. Last, but most important,
there is the job of ensuring that students are engaged in effective learning. So the next four sections cover
the following.
Basic classroom teaching technique (Section 3.7)
professional and personal issues
starting the lecture
ensuring satisfactory delivery
closing the lecture.
Use of resources (Section 3.8)
use of accommodation - media and message
use of black (white) board technique - media and message
use of OHPs - media and message
use of handouts - media and message.
Maintaining a conducive learning environment during the lecture (Section 3.9)
establishing good relations with the students
classroom management
motivating and challenging students in the lecture
encouraging student engagement and interaction in a lecture
helping students to get the best from the lecture.
Lecturers facilitation and support for the students learning (Section 3.10)
ensuring that students learn during lectures
the lecturers appreciation of students understanding, needs and level
careful explanation of complicated ideas.
A great deal of this material is generic, applicable to the teaching of any subject, so you might see some
of it in your generic institutional staff development courses, and here we will focus on the mathematical
aspects. There is no harm in repeating such material in the context of mathematics, for mathematicians
of all people can appreciate the power of particular examples of general principles!
61
3.7 Basic Classroom Teaching Technique
In this section we look at the basic mechanics of giving lectures, focusing on running the class in a pro-
fessional, organised and interesting way, establishing effective communication and relations with the
students. Specically, we will be looking at:
professional and personal issues
starting the lecture
ensuring satisfactory delivery
closing the lecture.
3.7.1 Professional and Personal Issues
In Section 2.4 Principle 2 emphasizes the fact that teaching is a matter of human interaction, and that
is the context in which professional is used here. As well as the professional responsibility to do a
good job technically, the teacher also has to cope with the extra demands of the human side of teaching
- remaining detached in difcult situations, setting an inspiring example, treating all students equally
despite any personal feelings towards them, and so on. We will look at this generally later (Section 3.9),
but for now we are concerned with the things that arise in the typical lecture, and how the needs of our
students might inform our professional judgement and behaviour. Essentially, we need to be enthusiastic
about our subject, condent and positive during the lecture, but alert to possible problems that might
arise. We have to conduct ourselves in a proper and considerate manner and be fair, rm but generous
with the students.
Enthusiasm
Time and again students put enthusiasm high on the list of qualities they like to see in their lecturers.
But sometimes the topic we have to teach really is quite boring, and then perhaps we have to fake en-
thusiasm! Or you might be open with the students and tell them that this part of the subject is in fact
a little uninspiring, but we will need it later on. In any event, you need the self awareness to recognise
when you are feeling a bit jaded about a topic, and have the self-discipline to do something about it - you
need to stay fresh ([53]). Then you have to nd the interesting aspects of the topic - there must be some,
otherwise you perhaps shouldnt be wasting precious class contact time on it. You can stimulate your
own enthusiasm by asking yourself such questions as:
history of the topic, where did it come from and why?
where does it lead to, what areas of maths are built on it?
what sort of applications does it have?
can I derive any of the results involved by different methods?
can it be generalized?
how is it related to my own particular area of interest - if it appears not to be, why?
why do the students need this anyway?
62
what concepts are at the core of the topic, and where else do these occur - can you imagine how the
original developers of the ideas found their way through them?
The point is to be imaginative about generating interest in the topic, and at the same time this might also
provide ideas about how to teach it.
Examples
1. Denitions and notation, particularly in something like abstract algebra or analysis, are
notoriously boring to put across. But there is usually a tale behind most denitions, a
reason for the name or notation and occasionally mentioning some of these can break up
the monotony. This might also help the students to understand why precise denitions
are so important, which is something they dont always appreciate. And the sooner the
denitions are used in subsequent material the better.
2. There is a lot of low level mathematics teaching, even in the most research intensive
universities, for example foundation or rst year service teaching. If you are a keen
researcher in topology and you are assigned such a class to teach then it may be very
difcult to nd something to enthuse about in solving quadratic equations, for example.
But that is the nature of the teaching job - we have to try to be enthusiastic and convey
that to the students, even if the material is elementary. Of course, quadratics are very
interesting to any keen mathematician. They lead us into complex numbers. For many
geometrical purposes they may be used to approximate other continuous functions. They
have a very rich history. Completing the square is an example of a magical trick used
everywhere in mathematics - getting something for nothing (adding and subtracting,
multiplying and dividing, etc). Any polynomial with real coefcients, .... . The list goes
on. Any elementary topic can be similarly elaborated.
3. Sometimes it is refreshing to approach a familiar task by a new approach or in a different
order. However, this should not be at the expense of making it more difcult for the stu-
dents to learn. If, for example, the lecture is devoted to a long proof, then the presentation
can be freshened up by simply bulleting the stages of the proof and then reconstructing
the details from scratch, with the students help, rather than by transcribing them from
prepared notes.
4. For such things as techniques of integration, matrix algebra you can simply choose dif-
ferent examples each time, but be sure these are going to work out conveniently - for
example it would be a brave lecturer that chose to nd the eigenvalues of a random 3
by 3 matrix. It is amazing how unexpected complexities pop up when you try newly
invented examples, and embarrassing when you have to try and rescue the situation. So
when you do manufacture new examples, test run them rst.
5. Mason ([53]) suggests such things as changing the symbols in some piece of work (Why
must it always be x and y?). This may be a useful tactic, but again be careful because
it requires intense concentration to keep this up (which is of course the point), and will
confuse students if you keep slipping between notations.
Exercise
Think of the most boring piece of mathematics you can imagine. Now nd something interesting
about it. What are the boring areas of mathematics, why are they boring, and how can they be made
interesting?
63
Condence and awareness of the difculties and stress that may arise
Good preparation will help your condence levels, but it is still likely that problems will arise, you will
make mistakes, you will be uncomfortable with some of the students behaviour, and you may have to
change your plans. Such things may get to you and may be stressful - but usually things are not as bad
as they may feel. Try to maintain a sense of humour, and never lose your temper or control. Learn to not
let errors uster you. Being nervous is OK, but learn to control it.
You may on occasions make mistakes in the class, everyone does ([7], P181). The real question is how you
deal with them. Some common difculties that can arise include:
losing your train of thought during a long sequence of arguments
minor errors such as dropping a sign or writing the wrong symbol
saying something different to what you actually write - as we will note later, in general any given
individual will think, write and speak at different speeds
omitting a crucial step because of your high familiarity with the material
going off in the wrong direction.
Usually the students will pick you up on these, so be attentive to this. Be alert to advance warnings, such
as students suddenly starting to talk to their neighbours - sometimes they are just checking if you really
have made a mistake before raising it with you. You may sense a bit of discomfort in the class, fumbling
with their notes, looking at each other, etc. Sometimes a number of students shout out queries together,
and you cant gure out what they are saying. Take charge, quiet them down, scanning your work as
you do so, and if this doesnt locate your error (It nearly always does), ask for one student to point it out,
thanking them for that. When you reread your writing on the board you usually read what you think is
there, so if students have pointed out an error read more carefully! A few errors, corrected, at least shows
the students that you are human - but of course there must be no errors in key ideas and concepts.
Exercise
In your next class make the odd deliberate mistake, and be alert for any student response. If nobody
picks it up go back and discuss it. Are they not paying sufcient attention, are you going too fast so
that it is all they can do to keep up with the writing, are they too frightened to question you? Then
reassure them that you are going to make real mistakes on occasions, and then you welcome their
intervention - they shouldnt feel inhibited.
Setting an example of proper behaviour
To ensure that lectures run smoothly and encourage productive learning you will need to insist on good,
responsible behaviour from the students. The best way to achieve this is to set a good example yourself,
being punctual and proper in your behaviour. and treating students with respect and courtesy. Of course,
students are people rst, mathematicians second (or third, or ... ). So dont let your feelings for them as
mathematicians colour your treatment of them as people. This can be very difcult, even for experienced
lecturers. It is easy to favour the keen, capable student and to prefer to be in their company, but in fact it
is often the weaker less interesting students who need the most attention. You will occasionally get some
students that are a bit bolshy. This seems to be not uncommon amongst bright mathematicians. Usually
they are well intentioned and genuine. Try to remember this when you encounter them and dont ght
re with re.
64
Treating students fairly and generously
A common description applied to the good teacher is Hard in the head, soft in the heart. You have to set
high standards of behaviour, thinking and debate, particularly in mathematics, but you also sometimes
need to be generous and give the benet of the doubt. This does not mean compromising on quality or
standards, and it is perhaps not so much about being soft on the students, but being hard on yourself.
All of us cannot help but let our own personalities and our own problems affect the way we deal with
other people, and mostly, with our family, friends, working colleagues, we get away with this without too
many repercussions in everyday life. But we cannot afford this luxury when interacting with students -
our position is more one of the priest or counsellor, and we have to set ourselves aside. There are many
occasions on which one gets frustrated, irritated or impatient with some of the students and it is then you
have to guard against inappropriate behaviour - you should certainly never actually display any of these
emotions to the students.
Personality, disposition and attitude of the teacher
As Krantz ([49], page 42) argues, you are the most important thing in the lecture room - and you are a
person interacting with people. But you are not interacting in the ordinary sense of people interacting
with others say in a social or work context. Here, interacting to produce the best environment for learning
is actually part of the job - in some sense you are actually acting. There are few professions in which our
personality and disposition are as crucial as in teaching. There are few professions that have such a
powerful inuence over their customers. A good teacher can inuence young people in a positive way
for the rest of their lives. A bad teacher can destroy condence, alienating students from the teacher and
possibly the subject. It is part of our professionalism to be aware of this and to control our actions in the
best interests of an enlightened educational environment. Essentially you have to adapt your personality
and disposition to the needs of the job and continually monitor yourself to remain fresh, self-possessed
and alert. You can, as much as possible be yourself, so you can relax, and help your students to get to
know you. Convince students that the material is doable, but it mustnt look too easy. Avoid wafe
and dont be patronising. Be an objective honest broker. Take charge of ensuring that the students have
maximum freedom within an enlightened educational environment. Dont abuse power. Both Wankat
and Oreovicz ([72], p 36) and Baumslag ([7], p 141) emphasize that the most important attitude a lecturer
can have is to be on the side of the students, to help them progress, and to keep encouraging them.
Examples
1. One of the most trying tasks in a lecture is to explain difcult ideas to the students with
a genuine desire that they understand. This means that you have to keep questioning
them and insisting you get an answer, and treating answers in a kind and patient way. If
you are by nature impatient this is very difcult indeed, but it has to be done, and you
need to train yourself to cope with it. To show impatience or any other negative emotion
is very counter productive and never advances the learning process.
2. Sometimes, when you are discussing an area of mathematics with your students, you
may inadvertently slip into the mathematical vernacular, forgetting who you are talking
to and thinking you are talking to a mathematical colleague, using all the familiar col-
loquialisms. Try to avoid this. Also, dont be unduly dismissive or opinionated about a
particular area of mathematics, whatever your views. Before the days of the RSAencryp-
tion algorithmand other applications of abstract algebra one would often hear dismissive
comments from some lecturers about the pure mathematics parts of the curriculum and
65
this might have inuenced students against the subject. The students are not interested
in our prejudices, lets allow them to develop their own!
Exercise
Think about yourself as a person. Do you have any characteristics that might enhance, or possibly
impede your abilities as a teacher? Are you naturally impatient, prone to arrogance, etc? Ask
trusted friends and family. How can you best use these characteristics, or diminish any negative
effect on your teaching?
3.7.2 Starting the Lecture
The way you start the lecture can set the mood for the whole class. You might settle themdown with light
small talk, give time for the (slightly) late arrivals, or get the administrative stuff out of the way. When you
do actually start do it clearly and obviously - that is, command their attention and concentration. Then
start with a brief outline of the purpose of the lecture, perhaps giving them a reminder of the previous
lecture. See Wankat and Oreovicz ([72], p38) for a start up list you can adapt to your own needs. Such
things as mentioning a recent current event that relates to the subject of the class, or comments on points
arising from recent coursework help students to settle into the new class.
A recap of where you got to in previous lectures can provide reinforcement of the previous material and
set up a context for what you are about to do. Baumslag ([7], p143) for example emphasizes the need
to continually introduce, summarise, review, and revise. He suggests the use of diagrams showing the
relation of the current lecture to previous material. In the rst lecture of a module you can give a summary
and overview of the whole module, but of course in laymans terms.
Examples
1. When coming to discuss the factor theoremfor polynomials, you can remind the students
of previous examples of factorising simple polynomials such as quadratics and some
cubics by inspection, and note the limitations of these methods. Then emphasise the fact
that the factors indicate precisely where these polynomials vanish, and indeed to nd
such points is one of the purposes of factorisation. One can then point out to themthat by
reversing this process we can devise a way of factorising more complicated polynomials,
and this leads naturally into the subject of this particular lecture. In doing this we have
reinforced previous methods of factorisation, and established a context for the new work
on the factor theorem.
2. Baumslag ([7], page 144) gives an example of an introductory overviewfor an elementary
calculus course. As he explains, the idea is to teach the main aspects of the topic concisely
using little more than intuitive notions and ideas that the students will already recognise.
Exercise
For your next few lectures, think ahead about what they will contain and draft a few brief lines of
introduction that will serve to introduce the topic of the lecture and lay the foundations for the main
content.
We can also spend a few minutes at the start of the lecture giving the students a clear vision of what the
lecture will be about - essentially the learning objectives of the lecture. Then they can start to construct
66
the context for assimilating the material and they can start to think about the nature of the intellectual
demands it might make on them. Some lecturers write out the learning objectives at a convenient space
on the board or on OHPs and keep referring back to them. Note that there is a slight danger when
announcing up front what you are going to do in the lecture. Some of the students might think they have
seen it before and mentally switch off, which is the last thing you want. So a bit of thought is needed
about how we describe what the lecture is to be about. This problem is particularly common in rst year
classes.
Example
Suppose the lecture is to rst year maths students with good A-levels and is devoted to in-
tegration using partial fractions. Now many of the students will have seen this before, and
will tell you this, and may be impatient at you going over A-level stuff. They might not see
the purpose of the lecture. But the real purpose of the lecture may actually be to consolidate
and extend their skills and speed up their performance, or maybe introduce them to some
short-cuts such as the cover-up rule for partial fractions. This is what should be emphasised
to the students from the start. They may have met partial fractions before, but it is unlikely
(for the UK students anyway) that their skills will be as highly developed as they need to be.
3.7.3 Ensure Satisfactory Delivery
Speaking and writing clearly
While this really goes without saying, it is an area where new lecturers can have difculties, and it in-
volves a number of dimensions, from how you speak to how you write and how generally you put
yourself across. Firstly, we have to ensure that what the students take away with them will enable them
to achieve the learning objectives. In a mathematics lecture this usually translates into sending them off
with a good set of notes and instructions for future tasks. So organise what you write/say to make it
easy for them to get the message. Use sectioning, headings, etc skillfully ([7], p166). Krantz ([49]) advises
that you write such things as conditions for theorems on separate lines, or as bullet points, not as a single
sentence with and between them. Give each theorem/technique an easily identiable name/number.
Baumslag ([7]) also gives suggestions for dividing/arranging boards to good effect and using lines, ta-
bles, arrows, and other diagrammatic ploys to link up associated ideas, equations, etc. Link what you say
and what you write considerately, to improve the quality of the students notes (It is a good idea from
time to time to read through some students notes to see if your message really is getting across). Not all
such ideas may work for you, but it illustrates the sorts of things we might do to help students get the
best from our lectures.
Pretty obviously, you need to ensure that the students can see and hear whatever you do. This is the sort
of thing that a generic staff development course might help with. It may include such things as voice
training, as for actors ([49] and [7], page 165). If you have concerns over your voice projection, use a
microphone. You can enliven your delivery with intonation, emphasis, body language - ll the room
with yourself as Krantz says, although in a dignied and reserved way. Face and talk to the students,
establishing eye contact with them. To emphasise an important point pause, say and write it out clearly,
tell them it is important, repeat it, ask the students if they have any questions, refer to the examination
and need for the ideas later in course. So letting students know when you are making a key point is an
important skill, which means continually thinking of the key ideas that you are trying to convey. For
things such as use of boards, see Section 3.8.
Example
Such things as pronouncing and writing Greek symbols can cause students problems, and
67
you need to be very emphatic about these things, particularly with rst year classes. In fact,
fromexperience at seminars, conferences, teaching observations it seems that this sort of thing
is not uniform across mathematicians, especially of different nationalities. Even experienced
mathematicians sometimes nd themselves having to double take at some strange symbol on
the board, or word uttered - so imagine what it is like for students. And of course mathematics
is so precise, sequential and fast moving that a seconds distraction trying to decipher a word
or symbol can lead to the student losing the thread altogether.
Exercise
Make a list of all technical terms and symbols in your course that you think might be new to the
students. Make sure you can write and say these clearly enough for them. Maybe issue the list to the
students and keep referring to it when the terms and symbols rst come up.
Refreshment breaks
Of course, even if the delivery is perfect few of us can sit through an hour or so of lecturing without
needing some sort of change of activity or break (But see below). And Principle 10 reminds us that,
particularly in maths, most of us learn best if we are actively engaged in the process. So every 20 minutes
or so provide some activity, change of activity or whatever to give the students a chance to recharge
their batteries. This is very easy in mathematics - you can simply give them a short problem to do that
either emphasizes a point you have made, or primes them for something that you are about to say. Such
activities can be directed at such things as the most difcult parts of the lecture material, the intellectual
hurdles, or at some subtlety that they might otherwise miss. Maybe a perfectly routine piece of work,
just to wake them up. Essentially, you have to think about what you want to get through in the lecture
leaving say 10-20% free time, and then think of ways to get the students active to the best effect in that
time. There are plenty more structured activities you can ask the students to do for mini-breaks, some of
which are listed below ([35]):
1. rest/time out/silent reection
2. read own or others notes as a prelude to the next stage
3. write down one or two questions that have occurred to you during the lecture
4. ask your questions
5. tackle a problem that will get you thinking about the next stage of the lecture
6. discuss a question of your own or the lecturers
7. apply a particular concept that has just been introduced
8. take a short test to check progress so far
9. planning and organising your notes and study of the subject so far.
Make sure that the instructions for such activities are clear and the students have any materials they need
for them.
An important point about such refreshment breaks is that you have to build them in and keep a close
eye on the time - it can easily be frittered away in this way, Also, you have to develop the skills of bringing
68
them back on task when you need to. Actually, this can have benets elsewhere in your relation with the
students. If they get used to the idea that you can let them go a bit - provide some playtime - but at the
end of the day you call the shots, you control the activities, then that gives them rules and boundaries
to work within. Another important point about this issue is that we should perhaps question the view
that a procient, able intellectual should not be expected to maintain attention for an hour on a topic of
relevance to their interests. Certainly, one would not provide refreshment breaks in a research seminar,
but these are not always very stimulating, and are usually more interactive than the average lecture,
and in any case we rarely get past the rst ten minutes ourselves! And our students are nothing like
so intellectually disciplined (yet), and your lectures are unlikely to be anywhere near as close to their
interests as a research seminar is likely to be to yours. So the lecture is a continual balance between
keeping the attention of the students and trying to stretch their powers of concentration.
Examples
1. Try setting a small task for the students to do while the class is assembling for the start of
the lecture - specied on an OHP, for example. Writing down a question they might have
about the topic to be discussed, recalling a result that they will be using in this lecture,
etc. Then for the rst break refer back to that exercise and get them discussing what they
wrote for a few minutes.
2. Working through particular cases of theorems often provides a convenient change of
pace or intellectual activity. Having spent a while working quite abstractly through the
proof, requiring deep concentrated thought, going through a specic example of the the-
orem provides a change of pace and level of difculty, a change to more routine and
undemanding thinking and is as good as a rest.
3. Extending a proof or result to a wider range of application, which may be non-examinable
and therefore not so serious - more relaxing. Atypical example here, for rst year classes,
is complexication of results and techniques in differential equations - just asking them
to consider what happens if you allow your variables and functions to take complex val-
ues can lead to lots of adventurous mathematics that may fascinate and enliven some of
the students - it will at least encourage them to think around the subject without being
under pressure to master it.
4. Discussing in pairs open questions such as What is algebra?, What is the connection
between algebra and geometry?. Such discussions provide a complete change for the
class fromthe meticulous attention to detail required for most of the lecture, and if chosen
carefully may awaken interest in the topic of the class.
5. See the example provided by Millet [55] in Section 3.3.
Exercise
Design a selection of refreshment breaks for use in your lectures, some frivolous, some fun, some
merely a change of pace or intellectual demand. Remember to allow time for these in the scheduling
of the class.
Pace
One of the most important aspects of good delivery is the management of the pace of the presentation.
This is a frequent student complaint, particularly about new lecturers. Adopt a pace that the students
can handle. Try to form a rough estimate of how long each part of your lecture will take, leaving a small
69
amount of slack for questions and activities. Then try to maintain an unhurried even pace. If things dont
go according to plan because maybe a topic took longer than you estimated, dont accelerate to cram
things in. Shift things into your next lecture and readjust later. If you nish your plan early improvise
some additional material, or simply carry on into stuff planned for the next lecture - or nish early (See
[72], p. 42). Another point about pace is that everyone thinks, speaks and writes at different speeds.
Sometimes, when writing on the board, particularly during mathematical calculations, you are thinking
one thing, saying another and writing something else. You can for example nd yourself missing out
some steps because your are thinking ahead too far! Slow down!
One way to control your pace is to keep asking yourself Are they getting this?, with a genuine desire
to know. Keep looking for reactions in the class - puzzled looks, signs of boredom or disengagement,
etc. And never let urgency to get through the material speed you up. You should plan and prepare
your delivery so that there is plenty of time to get across the main messages without having to rush. But
remember there is also such a thing as going too slow and labouring the obvious. As always try to get
the balance right.
Exercise
Practice pacing your lectures, monitoring your speed of writing, talking, delivering generally. Get
used to pausing occasionally to let students keep pace with you. Vary the pace depending on
difculty of material.
Balance between theory and examples
Another component of satisfactory delivery is the balance between theory and examples. No one can
listen for very long to a monologue of information about any subject without very soon needing some
sort of concrete example or application - something that they can play with to check that they have the
right idea. Indeed most mathematicians usually commence a research problemby playing with examples.
On the other hand too many examples and applications will simply slow down the delivery, so a balance
needs to be struck. By an example of course we mean any specic instance of a general principle, result,
technique, etc. It doesnt have to be applied in the usual sense, but just something that illustrates a
general statement. When explaining a long sequence of mathematical arguments one is often spared
the need for examples because the logical development of the argument is self-explanatory. So one may
lecture for some time before the need for an example arises, but sooner or later you will have to give
examples of some particular point, if only to change pace and give the students opportunity to think
about what you have been saying. One thing is for sure, you will never give enough examples to satisfy
all the students, because it is these that turn the material into things they can more easily grasp - that
gives it a more familiar context. As for the sorts of examples, these should not be too difcult nor too
routine. See Mason [53] for a discussion of what makes an example exemplary.
Examples
1. Immediately following the presentation of any major new technique such as inverting
a matrix or integrating by parts, we might present a range of worked examples show-
ing the main features of the technique. For example for matrix inversion a couple of
non-singular matrices, and then illustrate what happens for a singular matrix. Or for
integration by parts illustrate by a straightforward one such as xe
x
and one that has to
be repeated such as x
2
e
x
or circulated, like e
x
sinx.
2. Following an important or subtle denition, give some specic cases, as for example
in the , denition of a limit, or the denition of a basis for a vector space. In this
70
sometimes give examples, or get the students to devise their own, that illustrate when
the denition breaks down, or stretches the extent of the denition to its limit. Mason
([53], page 14) gives a useful tactic for this, in his boundary examples.
3. Another use of theory and examples in conjunction is the parallel treatment where some
theory is presented on one side and, in parallel stages a specic example on the other.
Thus one may work through the proof of a theorem in the general case and in parallel
present a particular example. Or, one might list the general stages in some technique such
as inverting a matrix, and in parallel do a particular example. See Burn et al ([16], Page
53 for the solution of a differential equation by integrating factor, with the general case
in parallel with a specic example. Or [21], p. 387 where the solution of linear equations
by determinants is illustrated in parallel with an actual numerical example.
Exercise
Construct and rehearse examples at appropriate points in your module, making sure that these are
illustrative of key ideas and results.
3.7.4 Closing the Lecture
The old adage Tell them what you are going to do, do it, then tell them what you have done suggests
we close the lecture with quick summary of the main points of the lecture. We should also try to end on
a high note, pointing to the (hopefully) exciting aspects of the lecture. Always end appropriately - not in
mid-sentence or halfway through a point, better to nish a few minutes early. Dont rush off at the end
of the lecture, make yourself available, see any students that you need to. This is another mechanism for
developing good relations with the students.
One useful way to close a lecturer is to ask the students to spend a minute or two writing down what they
have got out of the lecture - say three main points, and collect the results as the class nishes. When you
close, you can also say what you will move onto in the next lecture, which shows them where the work of
this lecture leads and prepares them for new ideas they will be meeting next time. This is part of putting
the lecture in context and developing an overview. In closing we may also need to remind the students of
any administrative matters, coursework submission dates, etc. Ask them to bring along anything specic
they will need next time. If it is actually important information, then mention it before they switch off
and pack up.
3.8 Use of Resources
This section addresses issues of the use of the various available resources - board technique, use of OHPs,
etc. Most of this will be generic and covered in your institutional staff development provision, so we will
only touch on some issues that arise specically in mathematics. As noted in Sections 2.8,2.9 there are
two aspects to the use of learning resources, which we can summarise in terms of media and message.
The media issues relate to actually using the resources to ensure that the communication of the message is
efcient and effective. This includes such things as writing clearly on the board, large enough writing on
the OHP slides, not standing in front of the board or projector, talking loud enough or using a microphone
if necessary, and so on. Any good institutional staff development generic course will cover most of this,
although they are unlikely to cover such issues as care needed in writing and pronouncing mathematical
symbols, or in preparing mathematical OHPs.
71
Using resources to actually get across the required message in the most effective and memorable way is
as important, and far more difcult than dealing with the media issues. Try talking mathematics to a
colleague, or over the phone, without being able to write things down. This will soon convince you of the
primacy of visual representation in conveying mathematical ideas and arguments, and should persuade
you of the opportunity black/whiteboards and OHPs provide for both good and bad communication
of mathematics. Naturally, this is even more important in a teaching situation. When working through
calculations with a colleague on a board you each have the opportunity to query what is written, to
explore different interpretations of what you see, to link closely with what is said. Most of this is denied
the student in the typical lecture. Sometimes their main priority is just to get something down, that they
can decipher later, so they at least have some sort of notes.
As we have mentioned previously, Mason ([53]) makes a useful point about the use of resources. He
notes that we can treat teaching materials, such as handouts, OHP slides, etc as objects to be used in
the course of the lecture, to encourage ownership by the students and reinforce learning points of the
lecture. For example, ask students to say what they see in a particular diagram or equation. Point to
important features of an item. Ask students to add values to a table, or reproduce their own version
of a drawing. Get them to annotate the item with their own ideas and then discuss/exchange their
annotations with neighbours. Encourage them to customise their handouts, highlighting key learning
points, incorporating additions, qualications, etc in their own way. Hide what was shown and get
students to construct it in their own way. Baumslag ([7], p160) also has some interesting ideas on the
use of props - using a torch to illustrate conic sections, students themselves to illustrate permutations and
combinations, an ofce spike to illustrate level surfaces, and so on. It is just a matter of being imaginative.
Based on the above, your use of resources might be developed in two stages: rst the technical use of
the equipment/environment then the application of those skills in getting across the key mathematical
ideas you are teaching. These days there are of course many sorts of equipment and resources and ac-
commodation. Material on some of the more specialist resources, such as the various forms of e-learning
can be found on the MSOR Network website (mathstore.ac.uk). In this book we will focus on black
(white)boards, OHPs and handouts. These are probably what the majority of us use. Also, the principles
of the use and application of these will translate easily to other media. So, this section will look at the
following in the context of the lecture:
accommodation
black (white) board technique
use of OHPs
use of handouts.
Use of accommodation - media
The lecture room is actually quite important. For large classes certainly, the standard arrangement for the
lecture is of a tiered lecture theatre. This is not always satisfactory, because you cannot easily get round to
the students (unless there is room for them to use alternate rows). It is useful if you can get close enough
to assist any individual. But even failing this, you can still walk up and down the aisles, which helps in
giving them the impression of being amongst them rather than in front of them. Try to use the space in
the room constructively. If you can get a at room then you can turn the lecture into a tutorial when you
set the students some work.
Think about the conditions in the room. Too hot/cold? Poor lighting, poor acoustics, and so on. While
such things are of course important in any discipline, they are particularly so in mathematics because it
72
demands so much detailed concentration over prolonged periods, and any kind of distraction or discom-
fort will get in the way of this. And few subjects require the same close attention to writing on the board
and OHP that are necessary in mathematics, so visibility has to be excellent for all students. So, as soon
as you know the room and the number of students you have, check it out for all such issues and get any
problems dealt with in good time. And remember that more experienced members of the department
will probably already know a lot about the quirks of the accommodation they use, so talk to them too.
Use of accommodation - message
The room and its xtures can provide imaginative illustrations of some mathematical ideas. For a start it
provides a readymade three-dimensional coordinate system. The windowblinds illustrate periodicity. Or
a roller board is better if your have it because you can write on it. Once again it is a matter of imagination.
Black/whiteboard technique - media
Good board work is obviously important. There is a lot on this in the generic literature. For mathematics
specic advice see Krantz ([49]). He exhorts us to write neatly in large letters in plain longhand or print.
Work horizontally across the board, proceeding linearly, with not too much material. We can divide the
blackboard into boxes to aid organisation by both the lecturer and the students. Label equations, use
horizontal and vertical lines. Draw sketches neatly and persist in getting students to do them ([49], Page
40). Use sliding boards effectively. If right-handed write rst on the right-hand board, then move to the
left, so as not to obscure what you have already written. Do not erase until you have to and then do it
properly. Dont reuse stuff already written, write it out again. If your writing is poor, slow down, take
special care. In truth, for most of us it is difcult to maintain such disciplined practice, but we have to try.
Black/whiteboard technique - message
When putting across the mathematical message by writing material on the board we have to think care-
fully about the actual purpose of displaying it on the board. If it is fairly routine stuff that students can
pick up pretty easily as they go along, then really all they have to do is copy it as a record. Note that this
is nothing like the waste of time and resources that some might think. The mere act of copying some-
thing down can help the students to absorb it. This is particularly so with mathematics, because even
when copying it down verbatim it is usually necessary to think through the symbols and manipulations.
Copying down the previous sentence verbatim would be ludicrous, but not so something like the power
series of the exponential function.
On the other hand if the purpose of displaying material on the board is to develop deeper understanding
(which most of it should be) then as Mason ([53]) points out we need to do more than just write the
material on the board. We have to actually use it in some way. Once the material, say a proof of a
theorem, or a worked example, has been written down and the students had time to transcribe it, then
we need to study it in more detail. You can walk up to the back of the class and look at it with the
students, discussing what you all see. Get them talking about it with you, see it from their point of view.
Discuss together with them what are the important features - would it have been a good idea to highlight
or underline some things - if so do it. Point out important features, or get the students to. Indicate to them
the logical structure, the steps made and the reasons for them. Can we improve the presentation in some
way, is it clear enough? Interact with it, pointing things out, drawing connections. Repeat the key phrases
or points and keep coming back to them. Say things in different ways, and remind them of the meaning
of key words and concepts. In all this you are using the board as a tool in developing understanding,
73
rather than simply a record of information. If this serves no other purpose than providing a short break
then it will be worthwhile.
Also of course the board should be used dynamically, as we go along, to illustrate mathematical think-
ing and argumentation. The students can get a great deal from this, watching an expert develop their
arguments on the board, writing out their thinking (rather than just copying from their notes), maybe
backtracking, making the odd mistake, asking questions, thinking asides, etc [71]. This is not to encour-
age you to present messy boardwork, but to supplement the boardwork with a running commentary of
what you are trying to convey and why.
Of course, there are those of us who are not so demonstrative and might feel inhibited about some of the
suggestions above. No matter, the point is that you have to nd your own way to use what you have
on the board, rather than simply writing the proof or whatever down and moving immediately on to the
next topic.
Another important aspect of getting the message across on the board is to make sure the message is all
there for everyone to see and ponder. So, for example dont gloss over steps - put each one in along
with words of explanation. When writing out mathematics steps on the board it is very easy to simply
write the equations, and not write down the intermediate words and explanation you are saying at the
same time. This is actually storing up trouble, because what the students then get as a set of notes is a
list of equations, and this is then how they tend to write out solutions of mathematical problems. We all
know this is not how good mathematics is presented, and that prose interjections are necessary in written
mathematics such as books and articles. They articulate what the symbols and equations really mean. Of
course, writing out the prose also slows you down, which is nearly always a good thing!
Use of OHPs - media
Here we include most kinds of projection facilities, including such media as Powerpoint. While such
things may differ in their technical facility they all basically project the nished material in toto on a
screen for viewing in polished form, rather than being put up sequentially by the lecturer with accom-
panying explanation and commentary, as they might do on a board for example. There is a danger in
such projections being used simply as an exhibit of the nal material, particularly in mathematics. The
students are denied seeing the material being developed, and simply reading off what the slide says is
an insult to their intelligence. A compromise is to conceal the slide and reveal it sequentially, but that is a
poor replacement for seeing the live development of the ideas that we talked about under the use of the
black/white board. So in consideration of such media the primary concern is how we use it to present
the messages we want to send in the most effective way.
Much of the use of projectors as a media for transmission of information is standard generic stuff that
will be covered in any institutional staff development course. For example, dont stand between audience
and screen, dont obscure projector with shoulder while writing on slide. In other words make sure every
student can see everything they need to see when they need to see it. The best place for the screen is in the
corner of the room at a slight angle, if possible, and so on. Use a ne pen to better present mathematical
symbols, etc. See Mason ([53], Page 44) for more on this is in the mathematical context.
Use of OHPs - message
Once projected of course the slide on the screen is no different to the board in terms of display (except
of course they can be prepared in much better quality, it is easier to use colours, etc), so everything said
above for boards applies equally for OHPs. However, slides are better for presenting summaries and
overviews of what you are saying. You can give students copies of the slides, whereas you cant for the
74
board, but dont allow this to overload the students or effectively increase the pace of the lecture beyond
the assimilation threshold. Multiple slides can be used, one showing a summary, one more details. OHP
and boards can be combined to good effect ([49], p. 40, [53]). And to reiterate the discussion above, the
live development of material, by the lecturer, on a board, or on an OHP, is far superior, particularly in
mathematics, to projection of the nal product on a screen.
Use of handouts - media
Again, you will get plenty of advice on preparation and use of handout materials in standard institutional
staff development generic courses. So far as mathematics is concerned you can probably get all you need
by looking at the various examples used by your colleagues. As mentioned elsewhere, dont use handouts
to rush the course or overload students and, especially for large classes, think of the costs.
Use of handouts - message
Everything we said about boards and OHPs applies equally here, but in addition the students can now
take these away and therefore have the opportunity to customize them and make them into reference
materials, revision aids, etc. See Mason ([53], P64) for different types of handouts and their use. Such
things as handouts may vary in form from level to level to reect the students developing independent
learning skills. Thus, in the rst year students need quite a lot of support, with materials such as handouts
that are pretty self explanatory and packed with examples. The level of guidance may be relatively high
and demand little from the student in terms of their own original input. By the second year they need
to be encouraged to become more independent and take more responsibility for their own learning. The
handouts can be quite concise, brief and have some white space that they are expected to complete with
their own work, notes from the lectures, etc. In the third year their independent learning skills should be
almost fully developed, and they can be expected to make much more use of books and their own reading.
One option here is to either use a set book, or provide book quality handouts, a little more advanced than
they actually need. Then you help them through it in your lectures, covering the important parts, leaving
them to read up the details themselves. In the lecture your job is not so much to give them the material as
to help them to assimilate it themselves. Incidentally, this move to getting students to rely on materials
to learn for themselves can be a bit worrying, because you may fear they wont cope and will end up
learning too little, but in fact they invariably rise to the challenge, and often surprise you.
Putting it all together for maximum effect
It is (too) often said that lectures cannot develop deep learning, by which is meant, presumably, the
exercise of the full range of higher order cognitive skills such as synthesis, analysis, evaluation and dis-
crimination (i.e. the full MATHKIT, Section 2.6). This is plain wrong. By combining boards, OHPs and
handouts and possibly demonstration models you can engage the students in all of these. For example,
we can provide the students with book quality detailed handouts which they are expected to read either
before or after the lecture. In the lecture we display the pages on OHPs, and work through them with
the students, just as we have previously described in studying a completed board. You pick out the key
points and ideas, which if they want they can highlight on their handouts. You can go through any awk-
ward steps or arguments in detail, going through extra examples or details on the board, which they can
copy or use to annotate their handouts. This way they are continually active and are having to think and
organise the material coming from different sources, putting it all together (with your help of course).
They practice how to identify the essence of things and not to be intimidated by complicated looking
75
mathematical details. Much of mathematics is repetitive, we keep using the same arguments over and
over. You can show this by getting them to scan though the handouts or book - pointing out how often
the same argument is used over and over. You can indicate what is important and not so important but
maybe interesting. Essentially, you and the students are working though a set of learning resources to-
gether, and they are being trained in most of the cognitive skills, and even in the transferable skills such
as communication, team working, etc.
Example
Complex analysis is a daunting subject for the novice, because there appears to be so much
intricate detail that is essential in actually implementing the methods of contour integration,
for example. By using a range of resources we can help the students to understand the subject
while at the same time exercising all the skills required to read, assimilate, learn and use
advanced mathematics.
The students are issued with detailed handout notes, which they could probably use indepen-
dently, without attending the lectures. Full proofs of the main theorems are given for example,
but only a few key outline worked examples. Copious exercises are provided (With answers,
but not complete solutions). As the course progresses they are expected to work through the
handouts and exercises, rereading as much as is necessary. In the lectures they are helped in
this by focused in-depth coverage of key points and key skills. The appropriate page is dis-
played on the OHP (breaking all, the rules about density of words per slide, but of course they
have the handout in front of them!), and the lecturer and students together work through the
particular topic(s) of the day.
In reality there are only a few major results in undergraduate complex analysis, one of the
rst of which they meet is the triangle inequality. The proof of this is given in the handout,
but in outline. The details are worked out carefully and interactively on the board, constantly
referring to the slide and handout. The students are free to copy the detailed proof from the
board as a supplementary note to the handout, or as most do, simply annotate their copy of
the handout. By the end of the course most students handouts have been fully personalized
by detailed scribbles and annotations. As the course progresses the triangle inequality is of
course used time and time again, and this can be illustrated by going back through the notes.
They soon realise the importance of the result, which further motivates them to understand it.
They see for example that the integral estimation theorem is not a new complicated result, but
just the triangle inequality in disguise. They see lots of different methods applied to the same
or similar integrals and get used to judging which one to use in different situations because
they can scan back through the annotated book and see what has gone before.
A key subtle argument repeatedly used in complex analysis is the demonstration that a com-
plex integral vanishes by showing that its modulus is less than an arbitrary positive quantity.
The rst time this occurs in the book it is treated in great detail with careful explanation and
examples. The next time it is done more quickly, and they can ick back through the pages
of the handout to rework it and remind themselves. By the end of the course it passes with-
out comment, except to point out to them how many times it has been used by icking back
through the notes. At appropriate points detailed examples are inserted and worked through
on the board, giving students the chance to practice them. This approach is independent of
the size of the class, although students do need some space to work when trawling through
the notes, for example.
76
3.9 Maintaining a Conducive Learning Environment in the Lecture
Some environments are better than others for learning - and these can differ from person to person. In
the lecture we have to ensure that the environment is optimum for learning for most of the students. This
means so many things - understanding which approaches and learning methods best suit your students,
getting them relaxed and able to ask questions, keeping appropriate discipline so that the class runs
smoothly. It is perhaps one of the most difcult things for the new lecturer to do, and even experienced
lecturers nd difculty, particularly if they are suddenly put in charge of a novel type of class for them
(this effect has been very noticeable with widening participation).
Much of this section is really generic in nature, devoted to skills that all teachers need, in any subject,
at any level. But the best place to learn about the issues involved here is in the department, with ones
colleagues and students. Your colleagues are likely to know the students and the particular challenges
you face. They can help mediate in difcult situations, and provide shoulders to cry on when the going
gets tough! In this book we have assembled a wide range of comment, advice, experience from across the
HE mathematics sector, so there is some practitioner validity and relevance in what you read, but that
wont address all the situations you meet in your career. Every student in every one of your classes is a
new experience who you may not have read about or come across before. In this section we will look at:
establishing good relations with the students
classroom management
motivating and challenging students in the lecture
encouraging student engagement and interaction in a lecture
helping students to get the best from the lecture.
3.9.1 Establishing good Relations with the Students
Easier said than done. It may require qualities you might not have - a friendly nature, sense of humour,
innite patience, etc. But there is one overriding quality that will take you a very long way - you must
really care for your students. Failing that, be able to pretend you care for them. Students will forgive a
lot if you have their interests at heart.
Krantz ([49]) holds as one of his key maxims for relations with students the need to respect them. This is
absolutely right. You have to treat them as people, not only as mathematicians - and as in most environ-
ments until you are given cause to behave otherwise, it pays to be nice to people. While you are the one
in charge, extend courtesy, consideration and respect to your students. Give value and do your best for
them. Be reliable, do what you say you are going to do. Be rm and fair and ultra-tolerant. Never lose
your temper or become impatient with your students, be utterly objective and detached when dealing
with them, especially in difcult situations (See 3.9.2 below). Find an ultimate deterrent or escape route
for those situations when things do start to get out of hand.
Baumslag ([7], p. 140) promotes a similarly student-friendly approach, advising the lecturer that the most
important attitude they can have is to be on the students side and to feel that it is important for them to
succeed. And let them know you want them to succeed. Encourage them and note improvements in their
understanding as the course progresses. Sprinkle the lecture with the occasional complement on their
work - I remember in the last coursework you did quite well on that point. As Baumslag emphasizes,
there is no need to pretend or be patronising about this - there will usually be improvements, but these
wont be so clear to the students.
77
Examples
1. Particularly if you are teaching science or engineering students, it is highly likely that
some will not be greatly interested in your subject - it is just something they have to do.
It is all too easy to let this affect your relationship with the students, but it is part of the
job to avoid this - to keep good relationships with the students even though you have
little of interest in common. You will be constantly asking the students to do things they
dont really want to do - so if you have to also do some things you dont want to do, then
you at least have this in common!!
2. Sometimes a student will ask what you may consider a silly question. You have a very
important power relation with your students, and the way you respond to even silly
questions can have a lasting demoralising effect on the students condence and ability
to consult you or other lecturers again if you get it wrong. There is plenty of evidence
that even top academics can ask extremely silly questions in subjects outside their ex-
pertise (and even within it), and when they do, they are no more immune to dismissive
responses than anyone else.
Exercise
Think back to your student, or school, days. Was there a teacher/lecturer who particularly inspired
or encouraged you, one you would wish to emulate? Why? Would they have the same effect on other
students? Note that it is not so much the individual reasons that are important here, but their long
term effect.
Another important aspect of building good staff-student relations is to treat them as individuals. From
the front of a class of 200 students it is difcult to see themas individuals (It may be difcult to see themat
all!), but we have to try. Elsewhere you will see advice relevant to this issue, such as learning some student
names, learning about their backgrounds, etc. Here we are more concerned with our overall attitude to
them. Never categorise students unjustly. It is tempting to fall for such generalizations as These are
engineers, they are not really interested in mathematics, These students are not interested in learning,
they just want to pass the exam, etc. But as Baumslag ([7], P 140) notes their motives might not be as you
would like, but as individuals they are entitled to their motives - if they can pass the examination without
learning that is your fault, not theirs! As we say elsewhere, the students are unlikely to be as dedicated to
mathematics as you are, and much as you might wish it otherwise the nature of the job cannot be based
on that assumption - you have to work with what you have.
Examples
1. While writing on the board you may sometimes hear some chattering which you feel
is beginning to escalate and become a nuisance. There is a tendency to think This is
a noisy class and to berate them generally (and therefore discomfort them all) with a
telling off, which may in turn escalate into threats of removal (we will look at this in
more detail below). Now in fact you dont have a noisy class - what you have is a few
individuals who are chattering. You therefore must identify these - or some of them -
and deal with them directly. You must not penalise the whole class because you have
difculty identifying the individuals concerned.
2. If a student asks a question in class you may have a pre-formed impression of what their
difculty is, perhaps from similar experiences elsewhere. But this student may well have
a different slant on things, may have their own specic difculties. So, concentrate on
78
them and try to determine precisely what their problem is. Their is no harm in establish-
ing a mini-dialogue for a few minutes, because the other students will probably benet
from this, even if it is only to see that you are prepared to talk with students in a friendly
and considerate manner.
Exercise
In a class you are currently teaching, how many students names do you know? What is the
background of the students? Sketch out their likely mathematical knowledge. Is there likely to be
wide variation about this? What are their likely motives for studying mathematics in your class?
3.9.2 Classroom Management
By this we mean maintaining an orderly environment, ensuring all the teaching and learning activities
proceed expeditiously in a disciplined and productive manner. Of course, you should always be aiming
to develop a productive, community working atmosphere within the class that is conducive to learning,
and relies little on formal rules. However, circumstances do sometimes arise where it is necessary to deal
with some kind of troublesome behaviour, such as noiseness during lectures or disruptive late arrival.
Advising lecturers how to deal with such instances is like advising parents how to bring up their children
- the advice may be unwelcome, and they may not be in a position to take it anyway. But we will try.
Of course such problems can occur in any subject and are not specic to mathematics. They might there-
fore be addressed within your institutional staff development department, or by departmental policies.
Nevertheless, you might nd the advice that follows useful, as it is distilled froma number of experienced
mathematics lecturers who have to deal with the issues on a daily basis, across a wide range of students.
Also, there are actually some specically mathematics aspects to these issues. For example a lot of mathe-
matics teaching is to service students, whose main subject and interest may not be in mathematics. Their
motivation and commitment to mathematics may therefore be less than that of mathematics students,
and you may have to work harder to accommodate this. Another problem specic to mathematics is that,
even for mathematics students, such a precise and efcient language can leave them behind faster than
most subjects. Your students are therefore likely to be particularly sensitive to the pace of lecturing, and
the class can become very restless if this becomes too challenging, or too pedantic.
The rst issue to address is when a problem actually becomes a problem. This is a delicate balance
between your tolerance in encouraging a relaxed, friendly atmosphere and the students need for a con-
ducive, controlled learning environment. It may be tempting, particularly in your rst few lectures, to
turn a blind eye to unwelcome behaviour. You may not think it is worth making a fuss. You may think
the behaviour is your fault, because you are doing something wrong. Or you simply may not know what
to do about it anyway. The way out of this is to ask yourself not whether it bothers you, but whether it
interferes with the other students, or the running of the class. If it does, then you must act to deal with it,
the students will, quite rightly, expect that of you. In general it pays to be quite tough in the early stages
of a course, then gradually lighten up as you and the students all get to know each other.
So let us assume you have got to do something about the problem. Of course, prevention is better than
cure. Set the ground rules in the very rst class, clearly and concisely. Try to get the students on side
by involving them in maintaining a good learning environment. Students are less inclined to misbehave
if they feel you know something about them, so learn some names and get to know as much as you
can reasonably manage. Keep seeking feedback from them, informally and by frequent measures of
progress. Keep them active. Be everywhere - wandering about the class rather than standing at the
front. Any cure of unwelcome behaviour must be systematic and fair - a rst warning in pleasant mode,
a second warning including statement of consequences, at the third offence impose the consequences
79
without embarrassment or compromise. If the circumstances are extreme and it cannot be resolved by
removal of some students, leave (with dignity!!) rather than lose it. Note that leaving the classroom is an
absolute last resort. To do so penalizes the well behaved students not involved in the incident. In all my
career as a university lecturer I have never found it necessary to abandon a class, although a few students
have been removed along the way! This is probably the case for most lecturers.
However, whatever preventative action you take, you are bound to get some problems,. Perhaps the
most common, particularly in a large lecture theatre, is talking and inattention. Usually the best time to
handle this is when it rst occurs and nip it in the bud. While everyone must see that this behaviour
is unacceptable, the class as a whole must not be chastised if disruptive talking is taking place, but the
culprits must be identied and warned. If students are chatting, make direct eye contact with themso that
they know you see them. Sometimes stopping the lecture, looking directly at the students, and resuming
the lecture when talking stops is enough to resolve the problem. However, this is allowing these students
to dictate the progress of the class and should not be allowed to delay proceedings for too long. Physically
move toward that part of the room, again making eye contact with the students. If you cannot identify
the source of the noise simply ask who is talking. If there is no answer, then their embarrassment should
be a sufcient lesson.
Having given fair warning, if you nally do identify someone talking then make an example of them and
impose the threatened sanction rmly, immediately and without malice - for example ask them to leave.
Speak to the student (s) privately after class or before the next session. Explain that their behaviour dis-
tracts the other students. Above all, make it clear that you bear no grudges, and that once the issue with
them is resolved it is forgotten completely and will have no further consequences. This is particularly
important, because you may shortly be marking their work, and you should not worry that your mark-
ing will be affected by a spat over their behaviour. In fact students are not normally being deliberately
troublesome, but have possibly simply not learnt the rules of reasonable behaviour, or have previously
been set a bad example by being allowed to talk in classes. Also, of course there are times when you want
them to talk - during tutorial sessions or time out periods, for example, or to tell you that you have made
a mistake, so you have to draw a balance between keeping a lid on them and allowing some chattering.
Another common problem is lateness and inattendance (of the students, hopefully not the lecturer!).
Students shouldnt skip lectures or be late/leave early, particularly if this is disruptive to others. It is best
to be rm on this, especially in the rst few lectures. Many lecturers leave the question of attendance to
individual students. If you require attendance, be sure to have a system for reliably recording it and a
policy to follow-up on absences. If you feel that a students absences are excessive and are jeopardising
academic performance, inform the students personal tutor and/or discuss it with the student. If a large
proportion of students dont come to the lecture, consider the possibility that they do not nd the sessions
useful. On the other hand, students are adults and whether or not they attend is really their business.
As for lateness, that is a different because it can disrupt the class, and should not be tolerated, except
possibly early in the rst year when students are nding their feet, or when it really does not constitute a
disruption.
These days there also seems to be a growing problem with eating and drinking in lectures and resultant
litter in classrooms. Usually there are regulations about this, and if so then they should be rigorously
enforced - it is not unreasonable to expect a healthy youngster to last throughout an hours lecture without
refreshment, and it should be made clear that litter in the lecture room(or anywhere else) is unacceptable.
Some of todays students do not appear to have learnt this.
At some point in your career you may have to face a student who is resentful, hostile, or challenging in a
lecture. The following are a few suggestions for dealing with this:
dont become defensive and take the confrontation personally. Respond honestly to challenges,
explaining - not defending - your objectives in the class.
80
at all costs, avoid arguments with students in class.
when talking to a disruptive student, be objective about expressing your concerns.
be honest when something doesnt work as you had planned.
in the highly unlikely event that a student becomes hostile or threatening, contact Security, inform
their personal tutor or head of department. Most campuses have disciplinary procedures that pro-
tect staff as well as students.
as a last resort leave and report the matter to the appropriate authority.
Despite the perhaps daunting nature of the issues discussed in this section take comfort from the fact
that they are generally of an exceptional nature. If you genuinely care for your students and do your
best to give them good value, then most of the time you will be able to create a lively but disciplined
environment. On the odd occasion when things go amiss then the bulk of the students will be behind
you in putting it right.
Exercise
Discuss the issues raised above with as many experienced colleagues (not necessarily in mathematics
only) as possible.
3.9.3 Motivating and Challenging Students in the Lecture
Mathematics is one of the most difcult subjects to concentrate on when you are simply listening to
someone talking about it. If you are already very interested in mathematics then the last thing you want
to do is watch someone else doing it - you want to get on with it yourself, occasionally asking for help
when you need it. If you are not that interested in mathematics anyway, then it is even more difcult to
concentrate on lengthy mathematical arguments rolled out by someone else.
In a post-it exercise at the Plymouth SEDA (Staff and Educational Development Agency) Conference,
April 1997 participants (probably very few of whom were mathematicians) were asked to identify sub-
ject matter that damages students motivation. Some 80 items were produced, mainly generic such as
too much information. The only specic discipline related items were all related to mathematics or
statistics - numbers, equations, graphs, mathematics, statistics and formulae, etc. (10% of items). So in
mathematics we seem to have something of a challenge!
In your generic staff development courses you may be told about such things as Maslows hierarchy of
needs, and advised how to motivate students in general terms. Useful though this may be, the especially
difcult task, for example, of motivating biologists in mathematics requires particularly imaginative tech-
niques. Here we are focusing on the task of motivating and challenging students in the actual lecture.
This is extremely difcult, because in order to get through the material you have limited time to spend
with motivational ploys. Also, you cannot get much feedback fromthe students to measure howeffective
you are at motivating them. In a tutorial situation (See Chapter 4) you can use more time and ingenuity
in motivating students, relating it more closely to the work they are doing.
Of course, the cynical amongst us will realise that there is one very potent extrinsic motivator - the as-
sessment. And if all else fails you can remind them about that. Try this game (only once!). Pick on a
topic, preferably a quite boring one. Wax lyrical about the interesting applications of this topic, its great
importance in the history of mathematics, and so on. The students will listen politely, grateful for the
chance to put their pens down and relax for a moment, and indulgent of your misguided interests. Then
81
quietly mention that this topic is a favourite examination question, regularly cropping up in the past.
Suddenly the room will erupt into activity as everyone scribbles reminders to that effect, annotates their
notes, etc, and you will have their full attention. It is up to you whether or not you do actually have
any intention of setting the question this year! The sad truth is that although the examination is indeed
a powerful motivator, it motivates to the wrong ends. It encourages students not to take an interest in
the course and its content but in the strategic task of passing the examination. Of course, we like to think
that students will be interested in what we are saying for its own intrinsic value and relevance, but as
ex-students ourselves we know the reality - and should use it. And of course, as it is such a fact of life,
we have to ensure we set the right sort of examinations to capitalize on it. So, make (not too) frequent
reference to the examination relevance of what you are doing, perhaps referring back to past examination
papers. This may not score highly in the educational respectability domain, but it works.
So how do we motivate students and sustain their interest through a fty minutes lecture? There is one
absolute, invariable, necessary requirement, above all else, you yourself have to be infectiously enthusi-
astic. We discussed the need for the lecturer to be enthusiastic in Section 3.7, here we will look at how
to get the students interested. Of course, you have to get all the technicalities of teaching right rst - the
students have you be able to hear, see and follow what you are doing. And of course, you need to make
it easy for them to appreciate the key points, and then have mental time to play with them - they can only
afford interest when they are spared the chore of trying to keep up. And the better your rapport with
them, the more likely they are to be infected by your enthusiasm.
We can motivate the students by giving examples, applications, asides etc that may be of direct relevance
to them. This does not necessarily mean venturing into their other subjects (For example, into electricity
when teaching electrical engineers, etc). In fact, this sometimes puts students off. So try to think widely
about relevance. It might be something topical, completely divorced from their academic subjects. Rela-
tively recent examples might be:
optimisation /A Beautiful Mind
waves, sine functions/The Tsunami
trajectories and rotation/England and the Ashes
prime number theory/Recent developments in computer or business security
Black-Scholes theory/The Credit Crunch
Lorenz equations/Climate Change.
With Google anything is possible!
Exercise
Think about some recent events from the world news that might have a mathematical context and
could t into your teaching.
We can invoke their curiosity by asking the students stimulating questions (not rhetorical, and not rou-
tine) at key points in the lecture. You are trying to unsettle them slightly, so they may feel a bit challenged
to respond and engage. And always listen and respond positively to students comments and questions,
taking every opportunity to open them out and draw in other students. Sometimes, you can ask for
their help - you be the novice, they the experts (e.g., something from their own subject or experience as
students, something relating to new technology, etc) - not teacher-learners, but learner-learners
82
Justify results, theorems, etc sensibly, as well as logically. This is key to maintaining student interest. No
matter how riveting the subject material, a fewlong sentences of dry impeccable logic (Not uncommon in
mathematics!) are guaranteed to set eyes glazing over. Soften such things by simple, sensible explanation
that gives the gist in a nutshell. This is actually a difcult skill and requires that you have a thorough
knowledge of the topic yourself. There is no mathematical topic, especially at undergraduate level, for
which such elementary heuristic explanations are not possible (discuss!). An occasional short amusing
story related to the topic in hand can often help them remember it and make it more palatable. If you
think back to your own student days you may well remember a particular result more because of some
quip by the lecturer than the intrinsic interest of the result itself.
Above we have emphasised the need to motivate students. There is, however, a caveat with the desire to
motivate students. When at all possible it is much better to explain where and why you are going in a
certain direction, or doing a particular step. But doing this to excess can sometimes slowthings down and
also be very difcult in the context of the students current knowledge. Sometimes it may be necessary
to simply ask students to have a little patience or faith in what you are doing, to take a leap in the dark,
assuring them that things will be clearer later on.
Examples
1. Teaching mathematics to non-specialists is notoriously difcult. They are unlikely to
have the same motivation as mathematics students, and indeed sometimes see mathe-
matics as a boring or a daunting subject. Getting them interested in what you are trying
to say is sometimes an uphill task. There is however an abundance of advice on this in
the literature. For example in the context of teaching physicists and engineers K ummerer
([50], p 330) recommends keeping in touch with parallel material in their other engi-
neering courses, and occasionally taking a page from their notes in such a subject and
illustrating the use of some mathematics you have recently covered with them. And as
K ummerer says students may sometimes ask you mathematical questions arising from
their other subjects - seize on these and explain how they can understand it using the
material you have taught them. Baumslag ([7], p151) has similar suggestions for moti-
vating engineers as well as ideas such as including quotations from famous engineers,
or inviting guest engineering speakers along, or historical asides, etc. However, as with
all such advice we must repeat a note of caution. An engineer or a biologist may not
necessarily thank you for doing an example from their own discipline. They may not be
that strong in the topic anyway, and it may simply mystify them further, as they try to
see the connections involved. Some may simply be happy to take your word for it that it
is useful and just want to get on with the uncluttered mathematics of the topic.
2. Helen Keller (1880-1968) was a US writer who had become deaf and blind at 19 months.
She was unable to communicate through language until she was seven, when she was
put under the tuition of Anne Sullivan. This may seem a strange example for mathemat-
ics teaching, but her story illustrates an important point about motivating students. A
lm was made about Kellers life in which Sullivan becomes increasingly frustrated with
her failure to communicate with Keller. Suddenly, Keller imitates an action by Sullivan,
who immediately recognizes this as an educational breakthrough and exclaims Thats
it - imitate rst, understand later - and that is the point. In any form of teaching it is
sometimes best to simply require imitation fromthe student, knowing that with repeated
repetition, understanding will follow later. In other words, the only motivation you can
give is to simply tell them to copy the process until they become familiar with it. This
is very relevant to mathematics, which of course is full of complicated processes. Some-
times the best motivation you can provide is that if they simply imitate enough times the
83
inversion of a matrix by the adjoint matrix, eventually they will begin to understand it.
The motivation is your assurance that it will make sense in the end.
3. The attainment of mastery is another powerful motivating force ([16] p. 83, 223). Virtu-
ally everyone enjoys learning a new skill or getting on top of an idea - watch students
in a tutorial, you can usually tell when they have mastered something. Skemp ([66],
p. 134) argues that one of the prime desires fullled by successful learning is the desire
for growth. Regular success and the attainment of mastery feeds this desire and reassures
the students that they are developing, so they gain condence. We all know the boost we
get when we nally crack something or master a new technique, suddenly understand
something (the eureka moment). We not only feel pleased at the success, but it changes
our attitude, propelling us forward more optimistically and energetically. So provide lots
of opportunities for students to succeed. In some circles there is currently a pedagogic
contempt for drill exercises where students plough through lots of routine problems,
but apart from establishing the basic skills this also provides long sequences of quick
successes that themselves are motivating inuences. In the classroom there is no harm in
asking the class a few rather straightforward questions (Not entirely trivial) that you are
sure someone can and will answer. This gives you the opportunity to congratulate them
on their progress.
4. Challenging to motivate. Most people, particularly students, like a challenge (Hence
Suduko, Rubiks Cube, etc), and this can be put to good use in helping them to learn.
There is evidence that mathematics problems become the most interesting when they are
just difcult enough - not too hard, not too easy. This is a matter of judgement, and
your knowledge of the class as a whole. And of course students will vary on what they
regard as the right level - a boring trivial problem for one student can be a fascinating
challenge for another. So you have to have a range of problems to meet their needs. In
a lecture you are limited to the sorts of questions you can ask anyway, but you can still
try to pitch the level right. It is judging this particular level for your class that is the
main challenge you face! K ummerer ([50], p. 330) makes the point that teaching should
be done on (at least) three levels: for the minimum level cover all the absolutely essential
points, then 95%of the time should be spent addressing the average student (the medium
level), interspersed with some additional challenges for the gifted students (the upper
level). In the same way we need to provide challenges in the lectures at these three levels,
just one or two hard examples for the class as a whole wont provide all students with
suitable challenge. They may keep the gifted students happy, but the weaker students
will simply get nowhere and become demoralised. So, provide a range of challenges.
But challenging students is not only about providing interesting questions, it is about
the whole teaching environment. When you explain something, make some of the steps
just a little bit beyond the immediate reach of the bulk of the students and tease out their
contributions to help move the explanation along. Have sessions when you stretch them,
interspersed with more easy-going periods (exactly like athletic training). The point is
that making the lecture environment just that bit challenging for the bulk of the students
is a useful motivational tool.
5. Relevant examples - wonder and mystery. We have on occasions noted that apparently
relevant examples may actually leave the intended audience cold. So what sort of ex-
amples can be used to interest and motivate students? Often it may be nothing to do
with their degree studies, but some interesting, even mysterious, examples that all the
students can appreciate whatever their background. An example I like to use with engi-
neers is that of local realism arising in quantum mechanics. The students need no more
than A-level maths and dont need to know anything about quantum mechanics. The
84
importance and correctness of quantum mechanics is evident to anyone who uses any
electronic device. But few can get to grips with its apparent contradiction of common
sense in its violation of local reality - essentially the belief that things are really there
when they are not being observed, and that no communication is possible at speeds
greater than that of light. Few can fail to be intrigued by this paradox, which has been
conrmed by many experiments in one form of which the question of local reality [5]
comes down to the value of a particular mathematical expression [45]
T = |1 + 2 cos x cos 2x|.
For any theory satisfying local reality this quantity must be less than 2 for all angles
x. If there is any angle x for which T can exceed 2 then quantum mechanics cannot
satisfy local reality [25]. There can be few more conceptually perplexing and important
questions in science or philosophy than this about local reality. And yet it is expressed
in a simple mathematical form and can be answered by elementary mathematics (No
more than A-level). The students are asked to determine all angles for which T exceeds
2, which needs only a double angle formula and completing the square. This example
is popular with the students because the problem of local reality has an almost mystical
appeal to anyone with any curiosity, it is easy to appreciate the issues, it is at the forefront
of modern science, the background is fairly easy to explain, and the mathematics is a nice
example of ideas that are normally a bit dry and uninspiring. Also, the solution illustrates
the necessity of having uency with basic principles in order to be able to tackle multi-
step problems.
Another similar example in which an intriguing, easily understood, question is explained
by nice elementary mathematics is why it is colder in February than in December, when
days are shortest and we get less heat from the Sun? The resolution of this puzzle comes
from the fact that the Newtons law of cooling modelling the heating of the earth under
the periodic temperature variations from the Sun can be expressed as a linear rst order
differential equation with a sinusoidal inhomogeneous part. Such an equation will have
a steady state solution that is a linear combination of sines and cosines with the same
frequency as the sinusoidal forcing function. By the compound angle formula this is
equivalent to a sinusoidal function with the same frequency as the forcing function, but
with a shifted phase - which accounts for the lag between December and February (Chris
Budd - private communication).
The point about such examples is that they are relevant to all of us and they challenge
our common sense and impart a sense of wonder. Then to see them explained by rela-
tively straightforward mathematics that the students can identify with generates a great
deal of interest. A similar example at a more advanced level is the RSA encryption al-
gorithm which can be used as an illustration in a rst course on abstract algebra. The
theory behind it is understandable at that level, but is rather dry, and by showing the
students how they can use it to understand and use the most powerful encryption sys-
tem known (Used millions of times a day at cash-point machines for example) intrigues
and motivates them. It would be useful to compile a compendium of similar inspiring
examples of elementary mathematics (See Exercise below).
6. The motivating power of the abstract. One of the great powers and beauties of mathe-
matics is that of abstraction. In schools there has been a move away from abstraction in
mathematics and a move towards the concrete. And many reputable teachers of math-
ematics, for good reasons, argue for concrete over abstract. But of course at university
level abstraction is the very essence of mathematics. For example, at school and in early
teaching at university matrices are a compact means of handling large amounts of data
85
or large systems of equations. Matrices are in that case a manipulative tool, comfortingly
concrete and numerical. But in more advanced mathematics they become conceptual
symbols in which the concrete properties are abstracted in order that we can, for exam-
ple, view a system of autonomous homogenous linear differential equations in exactly
the same way as we do the single equation y = y. This graduation from the concrete to
the abstract is one of the most difcult stages in intellectual development [69], and we
cannot expect students to achieve it overnight when they come to university. It has to
be gradually and proactively developed. Given sufcient time the students are perfectly
capable of appreciating the power of abstraction, and many of them will delight in the
freedom it gives them. If they fully understand it, the realisation that a system of differ-
ential equations can be treated analogously to the simple rst order equation in a single
dependent variable is liberating and exciting, and motivates and challenges the students.
Exercise
Compile a compendium of motivational methods/resources in mathematics.
3.9.4 Encouraging Student Engagement and Interaction in a Lecture
In a lecture you are limited in what you can do to get the students engaged, but it is still possible. We
need to respond sympathetically to queries and show that we expect them to interact and contribute
at appropriate times in the lecture. Concentrate not only on the academic content of the lecture, but
devices for developing rapport with the students. Most people learn more easily when they are (not
too) relaxed about showing their ignorance, and feel able to ask questions when needed. So your mode
and attitude of delivery needs to incorporate this if possible. It is not easy to do this - some people
exude empathy, warmth, helpfulness without even trying, some simply leave others cold no matter how
helpful and sympathetic they are. Being relaxed yourself, a little light-heartedness, and interest in the
students can help. Perhaps its the things you shouldnt do that make the difference - no sarcasm, no
impatience, no go and look it up, etc. Baumslag ([7], p. 149) makes the point that students will learn how
to ask sensible questions by experiencing you asking questions, so pepper the lecture with your own
considered questions. But remember that such activities can eat up time, and not always productively, so
use sparingly and strategically.
When students do pluck up the courage to ask questions in class respond to themin such a way as to help
them as much as possible, and others in the class. When a student asks a question they are at their most
ready to learn, so we need to capitalize on this. Build an environment for interaction, and build students
condence. Ask the students questions, but insist that you get answers. Question and answer in large
lectures can be particularly difcult. If you want students to answer and ask questions you need to:
explain the question clearly, possibly displaying it visually
give them time to get their thoughts together
provide privacy so that asking and answering do not feel so threatening
allow students the opportunity to prepare questions and answers before going public
never criticise or ridicule an answer - and indeed build on it
persist.
86
Some lecturers ask particular students questions, by name. So long as you have developed a good rapport
with the students, this is ne, but be careful not to embarrass a student by doing this.
Examples
1. If covering techniques of integration, for example, ask students to design a ow chart or
algorithm for tackling integration of rational functions. Encourage them to devise their
own mnemonics, for example. These days, many UK school pupils are used to being
told which method to use in a problem and to concentrate largely on how to perform
the method. So, having learned the different methods of integrating rational functions
(Substitution, partial fractions, completing the square) they need considerable practice
in dealing with unseen problems without being given the method. They need to de-
velop skills of deciding on the appropriate methods, then discriminating between differ-
ent methods that are equally valid but of varying practicality. Such ow chart construc-
tion can be applied to any number of mathematics topics that involve consideration of a
number of different cases or methods - solving polynomials, solving rst order differen-
tial equations, testing convergence of series, etc.
2. If you have just worked through a proof or argument with an involved structure, get
the students to work through particular cases in small groups - noting down the main
stages and how they link together. Then get them to generalise by for example replacing
particular numbers by symbols. We can do the same thing with a complicated technique
or algorithm.
3. Baumslag ([7], p. 149) suggests some questions the lecturer might have for the class con-
cerning a theorem that is currently being studied:
(a) Does the theorem seem reasonable?
(b) What does it mean geometrically or algebraically?
(c) Can you give an example
(d) Why does it have such peculiar conditions?
(e) Which theorem is it related to?
(f) Could we generalize the theorem?
(g) What is the converse? Is the converse true?
(h) Where can I use this in subject S?
(i) Is there an example that basically illustrates the whole theorem?
Treating the occasional theorem in this way, and encouraging the students to do so can get
them used to discussing mathematics with others.
Exercises
1. Think of a recent lecture. Could you have engaged the students more, involved them in more
interaction? How? Would you still have got through the same material, but possibly
differently? Think of a lecture to come. Can you use any of the lessons learned to make it more
interactive, without sacricing coverage?
2. Choose a theorem or standard result and frame questions such as those in Example 3. Do a few
examples where you provide answers to the questions.
87
3.9.5 Helping students to get the best from the lecture
Principle 8 (Section 2.4) reminds us that students may also need advice and training in using lectures
and other learning experiences, that is in learning how to learn. The real place to do this is the tutorial,
but there are some things we can do in the lecture to help the student to get the best out of the lecture.
One thing we can do is set a good example as we work though the material on the board. It is easy to
plough through the meticulous logical development of the material, simply vocalizing the board work
or the slides or whatever. Then all the students get is essentially a copy of your notes and what you say.
But instead, you can develop the arguments as if on the hoof, noting down the main points and actions
on the board/slides, showing how you might reason your way through the material as if your had never
seen it before. Now, I need something which when I differentiate it is going to give me this on the LHS,
any ideas? Didnt we see something like this last week? And so on. This way, you not only present
the material in a more honest and understandable way, but you are demonstrating to the students how a
mathematician actually thinks and works (As opposed to how most textbook writers write). As well as
copying just what you say, students will learn to copy what you do and how you think, they will get a
feel for the traditions and ethos of the subject.
Baumslag ([7], p. 77) compares lectures to a sandwich - the preparation by the students is the bottom
layer of bread, without which the lling falls out in a mess, the top layer the students work after the
lecture without which the lling wont be retained. So, you can advise students to prepare before the
lecture by consolidating the previous material, think about where it is leading. Then of course tell then
they should attend all lectures, and then after each they should work through the notes, test themselves
on the material, write out a summary, and work though a few problems, any work set, and so on. Of
course, as an ex-student yourself, you do not need reminding that You can take a horse ..., and many
students will ignore this excellent advice - but then you have at least done your bit. Mason ([53], page 95)
also gives some advice on helping students to study generally. We will return to this topic in Chapter 4,
in the context of tutoring.
As mentioned previously, a minute test at the end, getting themto identify the main points of the lecture,
can aid their learning and also tell you how well they have assimilated the material. You can then briey
tell them what the key points are, and advise them to spend some time before the next lecture ensuring
that they understand each of the points, and producing an example of each, and raising any questions
that still bother them at the next lecture. This is a good ploy because identifying and exemplifying the
main points usually carries the heaviest intellectual demand, and they are being focused on it and helped
to start it by you, while it is still fresh in their mind. They also know that if any problems arise you will
deal with them next time.
Dont forget that many departments have booklets such as How to study mathematics, or your univer-
sity may put on general study skills courses. So you dont want to duplicate these, but concentrate on
how students can get the best out of mathematics lectures and yours in particular.
Exercise
Does your department have such booklets of advice for students on how to study maths, how to get
the best out of lectures, etc? Do you know of anything useful in the literature on this?
88
3.10 Lecturers Facilitation and Support for the Students Learning
3.10.1 Helping Students to Learn
This is really what teaching is all about - to educate our students to the best of our (and their) ability, and
really help them to learn. In Chapter 2 we emphasised this, we summarised ideas about how students
learn and used them to underpin the eleven basic principles (Section 2.4) that guide us in the job of help-
ing students to learn. Also in Chapter 2 we based the design of teaching strategies and the preparation
of learning materials on these principles. Now we want to turn this into concrete ideas we can use in the
classroomto facilitate and support student learning. Howfor example do we help students in making the
reconstructions necessary to develop new concepts and ideas? In this section we consider such things,
and look at:
ensuring that students learn during lectures
the lecturers appreciation of students understanding, needs and level
careful explanation of complicated ideas
getting feedback from and to students.
3.10.2 Ensuring that Students Learn during Lectures
The rst thing to recognise is that making a real effort to ensure that the students learn, and that they take
the initiative in learning, usually takes a lot more time in lectures than if you simply tell them about the
topic, assuming it is going in. It also of course takes longer preparation, as discussed in Section 2.9, and
as a result of this you may nd that you are not able to cover the syllabus. This is a serious concern - if
you slow down so that the attentive students really learn, then you may not get through all the material
on which subsequent lecturers may rely. The priority has to be getting through the syllabus, until you can
negotiate a revision of the syllabus agreed by all your colleagues. Also, remember that the lecture contact
time is only a proportion of the student learning time for the course - you might for example have 30
hours lectures, but it is assumed that the students will put in twice that in their own learning time. If the
module is designed properly, then if the students really do put in that time, they have a good opportunity
to assimilate what you cover. In the lecture you provide the raw material on which they must work - it is
not the job of the lecturers to do the learning for the students, but to make it as easy as possible for them
to learn.
It requires sustained concentration to focus what you say and do in the lecture on helping the students
to learn rather than just reciting what it is you want them to know. As well as most of the skills and
training advocated in this book, it also requires a genuine desire on the part of the lecturer to interact
with the students in a range of ways that help their learning. If we are describing a technique such as
integration by parts it is so easy just to write down the formula and work though a few examples, with
hardly any interaction with the students. However, if we really want them to understand then we have
to work with them through the origin of the formula, emphasize that it is the reverse of the product rule,
repeat it in words time and again, and then work though examples drawing in the students at each step,
constantly referring back to the formula, making them do the work, while you write out the solutions as
they construct them. This is so much harder, but is essential if we want to help students to learn, rather
than just copy our notes down. The lecturer actually talks less, but then what they say has to really hit
home, and again, this takes thought and concentration.
89
3.10.3 The Lecturers Appreciation of Students Understanding, Needs and Level
In Sections 2.5 and 2.9 we emphasized the need to pitch materials at the right level for the students, and a
lot of that is equally relevant in the lecture itself. You should certainly have a good insight into the back-
ground knowledge of your students, and their facility level and be able to talk to them in that language,
but there is a little bit more to it than that. You need to empathize with your students and get on their
wavelength. This is one of the most important aspects of any form of teaching, but particularly applica-
ble in a lecture. A lecture is not about talking to quick thinking stenographers. It is about explaining to
widely varying human beings. One of the biggest problems here is getting used to the fact that few of the
students are really that interested in what you have to say, particularly in the case of service classes. This
doesnt necessarily mean that they are lazy or disrespectful, but maybe that they are busy and have other
equally important priorities. You may ask why you should make extra efforts to explain things to people
who are not interested? To illustrate this we will go through a scenario that you may be familiar with.
Most of us have had to learn some software in recent years. In this we are in a similar position to, for
example, an engineer learning mathematics. We want to be able to use it reasonably well, and will invest
some effort in learning it, but it is really a very small part of what we do and we have so many other
things crying out for priority. If we get stuck we may ask a techy for help. He talks way above your
head, berates your lack of expertise then wearily sorts it out himself. However good he is as a techy, he is
no good as a teacher.
You call in techy2. He listens patiently to your problem, reects a moment, then leans across, taps a few
keys, presto done, and lets you follow suit to check you can do it yourself - job done. But is it? He has
just got you to imitate him - what do you do if a similar but different problem arises again? This guy is
very helpful and has solved your problem by doing it for you, but he has not helped you to learn, he is
not a good teacher.
You call in techy3. She listens to your problem, probably sees the solution straight away, but doesnt
let on immediately. Instead, she asks questions to nd out what you already know, how motivated you
are, how much time you can devote to this task, what your precise needs are. Then she explains, using
language that she now thinks you will understand, and in a depth that she thinks will benet you most.
Your problem is solved - only now your understanding will be much more permanent and portable. She
is a good teacher. She appreciates that although you are willing to learn, this topic does not have a high
priority, so she has to explain in the most efcient and effective way in the context of what you already
know and your possible future needs. You dont want to be an expert, and you are not interested in the
ner details - you are not lazy and stupid, you are just busy and in an unfamiliar environment. Techy3
has had to work harder to get onto your wavelength, but she has done the job required. She has also
reduced the probability that you will have to go to her again!
The analogy with teaching mathematics to less than avid students is obvious. Being able to explain things
clearly, at the level of the student, is one of the most important skills of a teacher [12]. In a lecture to say
100 students it is of course even more important and more difcult. They are probably all at different
levels. Each may understand different fragments of what you say. You cant easily interrogate them to
nd out what they know or whether they have understood you. For this reason it is best to work near to
the lowest common denominator of what you think they know, and must continually be on the lookout
for signs that you are losing them. Lewis and White ([51]) give some ideas of how to adapt your language
to that of the students.
Examples
1. Particularly when taking a rst year class, we sometimes adopt a level of language that
really expects too much of the bulk of the students. For example we might automatically
use the term polynomial without comment. But although it is mentioned in most A-
90
level syllabuses, few rst year students are really familiar with it - the same certainly
applies also for rational function. Whats more, few students appreciate the value of
classifying the functions in this way. So, particularly in a rst year class, we need to use
such terms considerately. If we say something like Unlike polynomials a typical rational
function is not continuous everywhere to rst year engineers (or even mathematicians),
the likelihood is that we will simply not be talking the same language as the students.
They may have seen all the words before, but few of them will recall and string together
the meanings quickly enough to get the message before we move on. Instead we will
probably help the students more if we say something like A typical polynomial (Sketch
one) has no breaks in its curve because it yields a denite number for every value you
insert, whereas a typical rational function (Sketch one -
1
x
will do) will have breaks where
its denominator becomes zero, and at such points it is not dened. Most students will
get more out of this form than the previous one, which we can start to use later when
they really do understand it.
2. Another problematical object for many rst year students is the exponential function -
again it is in the A-level syllabus, but not always taken seriously in schools. Conse-
quently some students are apprehensive about it. In particular they dont always see the
connection with indices. You can help them by emphasizing that algebraically that is all
there is to the exponential function - the laws of indices. There is nothing special about
it in manipulative terms. And the apparently mysterious value of e comes about simply
because the exponential function is its own derivative, and therein lies its importance.
This is far more enlightening than starting off with the series or limit denition, or the
example of compound interest.
Exercise
Practice explaining difcult mathematical ideas to a non-mathematician, a friend, partner, relative -
anyone who will make an honest attempt to understand what you are saying. Notice the sort of
questions they ask and think about whether your students might be worried about similar questions,
but feel unable to raise them.
3.10.4 Careful Explanation of Complicated Ideas
This is one of the key skills of good teaching. How well students learn depends on how effectively we
explain what we are trying to get across. There are good explainers, and there are bad, and skills in this
area can be improved by education and training ([12]). We will look at this in great detail, and break
down the task of effectively explaining ideas into the following stages. All of these derive either from
our eleven principles or the underlying educational research (Section 2.3), and some have already been
discussed in the context of producing learning materials (Section 2.9):
assist students in chunking information
provide hooks for students to hang ideas on - aide memoires, mnemonics, etc, to make it as easy as
possible to remember key ideas
provide appropriate step-laddering and lubricate arguments
anticipate the difculties students might encounter and support them in dealing with these
provide regular overviews
91
continually highlight/reinforce the (small number of) key ideas conveyed in the lecture
bring out and highlight the essence of a complicated idea, theorem, method, etc and support the
development of students intuitive views of formal arguments
emphasize the respectability of guessing and the necessity of checking
alert the students to patterns that may assist their understanding and learning
help students to build up connections by providing roadmaps and overviews of difcult sequences
of arguments - try to express proofs sensibly, as much as logically
alert students to the fact that some teaching and learning is provisional, and may have to be revised
at a later date.
Assist students in chunking information
The concept of chunking of information, ideas and processes in cognitive development is fairly well
established ([58]). But it is only a rst step on the path to deeper understanding - we all break complicated
ideas into smaller more manageable units (After all, that is what theorems, lemmas, propositions, etc are
for!). What is not so good of course is to stop there, without subsequently internalising the chunks and
assembling the whole picture from them. In the lecture environment the bite-sized chunks translates
into breaking down the topic or result into smaller more easily memorised results, and emphasising
these, continually revisiting them, hammering them home, illustrating them, etc. And from time to time
you need to test that the students have actually swallowed them!
Examples
1. In teaching introductory calculus one might regard the standard derivatives as bit-sized
chunks. The important cases of these can be thoroughly learned and understood inde-
pendently as separate entities. For mathematics students they would each be proved in
detail, for engineers maybe just hand-waving, but for all students they must be known
thoroughly, available to instant recall (absolutely not on a formula sheet!), which the lec-
turer can easily check with short verbal or written quizzes. These bite-sized chunks are
then readily available to the students as they move on to more complicated functions
such as products and compositions. Then, when they are learning the product rule, they
can concentrate on its structure without being distracted by the need to think hard or re-
fer to the formula sheet for the standard derivatives involved. Ditto, standard integrals.
2. When teaching matrix inversion by the adjoint matrix, the bite-sized chunks might com-
prise evaluating the determinant (thereby checking that the matrix is non-singular), eval-
uating each cofactor, using these to construct the adjoint matrix, dividing by the deter-
minant. In this case the bite-sized chunks are actually steps in a sequential process, and
eventually the student will have to get used to putting them together in the right order,
but initially it may help if they can focus on them independently.
Exercise
Think of a few lectures you have to give. Can you break each down into more manageable chunks in
some way. How would you do this. How would you ensure that the students appreciate the fact that
you are chunking the information? Think about how you would put them together at the end of the
lecture to form a coherent whole.
92
Provide hooks for students to hang ideas on - aide memoires, mnemonics, etc, to
make it as easy as possible to remember key ideas
Preferably funny, or otherwise memorable. Also, link them to existing knowledge and practical applica-
tions. But emphasize that eventually the student should normally internalize the idea and dispense with
the hook (See Section 2.9). This is of course a key device in mathematics, employed all the time. Acronyms
are widely used in mathematics (Yet we pour scorn on them if they are used by university administra-
tors and managers!). In Section 2.9 we have already referred to some examples in this area: BODMAS,
CAST, rings and sings, Sketch GRAPH. And indeed MATHEMATICS is a useful aide memoire for
teaching itself. Below we give some more examples. But remember the warning about over reliance on
such tricks (See Section 2.9). Also, there is a tendency to criticize such aide memoires saying they are
mindless devices that do not reect real understanding. Of course, that might be true if they are simply
left as labels. But they should not be. They should be similar to the name of a friend, where the very
name (a label!) conjures up all sorts of memories, knowledge, attitudes and feelings. A related ploy is to
name rather than number important equations and theorems whenever possible. It is usually easier to
recall something with a name than with a number!
Examples
1. Hooks can sometimes take the form of brief memorable anecdotes, maybe a little his-
torical tale, or a recent news story. For example Gausss summation of the integers, the
Pythagoreans dismay at the irrationality of

2. Many such stories can be found in the


books by David Wells, for example [74], [73].
2. Much mathematical notation and terminology can be made more memorable with the
right formof expression or explanation. Baumslag ([7], p. 171) gives some examples, such
as noting that there Exists is backward E, for All, upside down A, Z for the integers
comes from the German z ahlen, etc. Or using An r c matrix instead of An m n
matrix. However, this last example serves to signal caution in adapting widely accepted
mathematical notation to make things more palatable for the student. Clearly, such a
notation will become cumbersome and inappropriate when one starts doing summations
over subscripts, and we will then have to get the student used to a new notation. And
what sort of understanding will a student have of matrices if they have to keep being
reminded of rows - columns. There are points at which one simply has to impress upon
the students that they have to get use to the standard mathematics conventions because
that is simply what they are, popular conventions. And in this particular example the
root convention i, j, k, l, m, n, ... for integers should not be compromised to cater for a
new situation.
Exercise
Think of some topics covered in your lectures. Can you nd some hook that will make it easier for
the students to remember and retain the material? If it is a technique, use your imagination to nd
some easily remembered acronym or mnemonic. If it is an overview of a sizeable portion of work nd
some sort of analogy that might be suggestive and easily remembered.
Provide appropriate step-laddering and lubricate arguments
In sequences of arguments, as for example involved in putting together the bite-sized chunks, one im-
portant factor is the height of the steps in the arguments. Obviously, if the logical or intellectual jump
93
from one step to another is too large then most of the students will simply not be able to follow you.
This is particularly important in a lecture, where the students might not be able to go over the argument
again, and not only that but they have to keep up with subsequent arguments. On the other hand there
is clearly a limit to how many steps to put in to help the student follow. Judging the step-laddering
needed is a difcult task. It is however better to get it wrong in the students favour. See [16], p. 19.
It is not only the intellectual steps that you expect of students that can affect learning, but the way you
present the argument or explain the steps. You can lubricate the argumentation by, for example, giving
analogies, repeating a key phrase that unlocks a mental block, scribbling a quick example, asking the
class to consider a leading question.
Examples
1. The product rule in differentiation. Having ensured that the students know their stan-
dard derivatives thoroughly (see above), you can move on to the product rule. You might
prove it for mathematics students (by the way, be careful with notation - you might use
the usual u-v notation, but if your writing is sloppy then the u and v may look the same
- perhaps use f-g!) and even for engineers, because it is so important. But for engineers
try and lubricate the argument by giving the analogy of the expansion in area of a heated
rectangular plate with sides u, v. When it comes to actually applying the product for-
mula you will have to think about the steps you include. Eventually of course you dont
really want the students to use the formula, but rather to differentiate the product on the
y without introducing u and v. But for complete beginners this is probably a step too
far initially, so we include all the steps, Put u = ... etc, write down and substitute into
the formula, collect and simplify terms, etc. This level of step-laddering would certainly
have sufced for the 10% participation of a couple of decades ago, but not always for
todays 40-50%. Many students come to university clinging to the full use of the formula,
and our task in the rst year is to wean them off this, which can only be done by lots of
practice. Ditto integration methods.
2. For the inversion of a matrix example the steps will of course depend on their previous
education. For example if they havent met cofactors before then you will have to take
steps to deal with them before forming the adjoint matrix.
Exercise
Analyse some of your recent lectures. When you work though a sequence of arguments on the board
for example, is the logical spacing about right for the bulk of your students? Are you sure that they
have a reasonable chance of tting the steps together comfortably, and when you are explaining the
steps in class do you ease the logical progression by what you say and do?
Anticipate the difculties students might encounter and support them in dealing
with these
As you are lecturing you will know that some things that you are going to say/do will be quite subtle
and challenging for the average student. Let the students know this - and modify pace and presentation
accordingly. Tell them - Now we come to a very difcult step, most students have problems with this
- I remember when I saw it for the rst time I had a lot of trouble. Then prepare them for something
that is hard by waking them up, ask them pointed questions, get them to discuss with their neighbours,
map out how we are going to approach it and why, emphasis that it is hard, but it is also very important.
94
Sneak up on it, keep referring back to it, etc. Make them see that you are anxious for them to get this -
you know it is hard, but you will do all you can to help them.
An excellent source of the sorts of difculties students have is Mason ([53]). Most of his rst chapter is
devoted to this subject. However, there is a danger of seeing the students as being unequal to the task of
learning mathematics until these difculties are diagnosed and dealt with. In fact this sells students far
too short and rather defeats the object of supporting them. Much of what are characterised as student
difculties could equally be attributed to their lecturers! Denial of having seen something before, un-
certainty about precise denitions (It wouldnt take long to nd a mathematician who is insecure about
the denition of a function), calling to mind things you are condent about in preference to those that
are new to you. These and many other characteristics are common to all learners, whether they be a rst
year undergraduate or an experienced researcher - move the latter out of their specialist area and they
can err as well as the next person. So, provided one views the students difculties as simply sticking
points that anyone can encounter, and one is alert to these and willing to do something about them in a
non-judgmental way, then Mason will give you lots of food for thought.
Examples
1. As mentioned earlier, when meeting the integrating factor method in linear differential
equations students usually nd difculty in the reversal of the product rule used to con-
vert the left-hand side of the equation to the derivative of a product after multiplying
by the integrating factor. You can help considerably by revisiting the product rule and
doing a few examples of converting expressions to derivatives - i.e. using the derivative
rule backwards.
2. General angles greater than 360 degrees, especially for inverse trigonometric functions,
also give most students some problems, so it will help them if you spell out in detail
the multiple angles that can result in nding inverse trigonometric functions, or nding
general solutions to trigonometric equations. Here the step that they often nd difcult
is expressing the general result in terms of an arbitrary integer. This is the problem of
expressing a sequence of numbers in a single formula, and we can help the students to
grasp the general result by methodically working through the angles produced as we go
through multiples of 360 degrees.
3. Concepts such as convexity, linear dependence, equivalence relations, etc always cause
problems - why? Because they are difcult! Baumslag ([7], p. 172-177), provides some
useful advice on minimizing the difculties, such as only using one term rather than the
two that confuses the students (e.g. convex not concave, dependence not independence,
etc). But this masks the real problem, as we said, that these topics are intrinsically dif-
cult. What is perhaps most surprising is that some lecturers, having thoroughly mastered
a topic, are not able to appreciate that the topic is difcult for the newcomer. It is a ne
skill to be able to assess what a novice will nd difcult and then to make it easier for
them. In fact, concepts like linear independence are exceedingly difcult at rst blush.
In truth, one cannot usually say one comes to understand them. Rather, one just gets
used to them! They require extended and lengthy treatment (And that is probably the
real cause of the problem, we think we dont have time for this in our over-crowded syl-
labuses), lots of practice and discussion, lots of revisiting and lots of time to absorb and
assimilate the complex ideas involved.
4. Sometimes you may have to rethink how you would phrase things to make them more
palatable for the student. For example in partial fractions we are often tempted to say
something like Equating coefcients .... Even the best students would be hard put to
really understand this move, and they are not likely to see immediately the connection
95
with linear independence ([53], p 27). It is probably kinder, and probably more truthful
to emphasize the fact that since the result is an identity, it must therefore be true for all
values of x, which can only be the case for a polynomial if the corresponding coefcients
are equal. Even if this is presented in a vague hand waving way, it will probably go down
better than Equating coefcients...
5. Abbreviations such as Iff take a long time to get used to, as does the actual idea of
necessary and sufcient. They need to be carefully explained ([53], p 27), with lots
of examples. In fact, schools do not appear to develop pupils logical skills to a high
degree. One often nds that not only do rst year students sometimes have problems
with mathematical knowledge and facility, but they may have problems with logical
argument, especially the precise and sophisticated levels found in mathematics. This is
not the students fault and there is no point bemoaning the situation and pressing on as if
it doesnt exist. One has to face the problem and develop the required skills from scratch.
And a module in propositional calculus is not necessarily the best way to develop these
skills - perhaps better to give copious examples.
Exercise
We have given a large number of common sticking points for students. For all your teaching, scan
through the material and identify such points where you think the students might have particular
difculty. Now devise ways to ease their path!
Provide regular overviews
Even when the bite-sized chunks have been put together and students understand the connections, and
can execute techniques and proofs with some facility, they still need to get an overview of the material,
to see what the general message is, to get an intuitive feel for what is going on. This can be developed by
ensuring from the start that the students understand what the main objective is, that they are fully aware
of the purpose of the lecture. You might for example state this at the beginning, repeat it at appropriate
intervals in the lecture, and at the end, in a summary. Wankat and Oreovicz ([72], p. 42) make the point
that this presents one of the major problems to new staff, because they are often overspecialized and may
not yet have developed overviews themselves. This is why teaching, especially at the elementary level,
can be much more difcult than we might imagine. There may be many approaches to a particular topic
and the good teacher needs to know them all, to be able to judge which is most appropriate in any given
teaching situation.
Examples
1. Having developed a lot of techniques of integration - by parts, substitution, partial frac-
tions, trig identities, and so on, you can point out to the students that this plethora of
methods can broadly be summarized under three main approaches:
Manipulating the integrand into a more helpful form(partial fractions, trig identities,
etc)
Reversing the product rule (by parts)
Reversing the chain rule (substitution)
Then ask them to notice one obvious omission (reversing the quotient rule) and thus
to discover for themselves a new and unusual integration method. This gives them a
compact oversight of what integration is about. It packages it and makes it look more
manageable (hopefully!).
96
2. Laurent series are a traditionally difcult area for students, with the problems of choos-
ing the right series for the different regions of convergence, and then algebraically ma-
nipulating the expansion to give the appropriate series. When lots of examples, with
different cases of singular parts and non-singular parts, have been presented, bring the
whole lot together with an overview using diagrams. Link the different regions with the
convergence regions of the different forms of the series involved. Show how the dia-
grams can be deduced from the singularities of the function to be expanded, and how
they dictate the forms of the series in each region. It is difcult to make this sensible
to the students until they have done a few mechanical examples, but after this, such an
overview can crystallise the key ideas.
Exercise
Work through your own modules and identify where such overviews and synopses would be useful.
You will probably only need one every few lectures, but allow time for them when they are needed.
Continually highlight/reinforce the (few) key ideas conveyed in the lecture
Mathematics is particularly concept efcient, which can make it either easy or difcult to learn. Any given
undergraduate course contains only a relatively small number of really basic ideas, concepts, techniques,
proof methods, etc. But by the time these have been recycled, applied, tweaked, the resulting set of
notes or book can run to many pages of complicated looking mathematics that can be intimidating to the
novice. The job of the lecturer is to reverse this process, to root out the important core ideas and help the
student to absorb these and to recognise them in different guises throughout the course. That is, to help
them through the wood, and also to help them in developing the skills to nd their own way through
other woods. You can do this by listing the key points at the beginning of the course, agging up every
time we meet them, summarising them at the end of the course. If you learn nothing else, learn this!
Examples
1. In a one hour lecture on partial fractions to rst year engineers, say, the key points may
be: i) the reason we might need to do it, ii) what a typical resolution into partial fractions
looks like, iii) what can be done with the resolution that cannot with the compound
fraction, iv) the use and properties of an identity, v) the process of decomposition, vi)
checking by putting back over a common denominator.
2. In a lecture on separating the variables in partial differential equations the key points are
i) the form of the solution z = f(x)g(y) ii) a function of x can only be equal to a function
of an independent variable y if both are a constant iii) the boundary/initial conditions
also need to be separated iv) solution of the resulting ordinary differential equations.
3. Baumslag, ([7], p.168), gives useful guidance on how to discuss the statement of a the-
orem, which is relevant here if a key theorem occupies the bulk of the lecture. The key
ideas can often be emphasized of course with an example, and Baumslag recommends
working though this in parallel with discussing the theorem, to illustrate the various
points. Amongst other things he lists the following key points to refer to when discussing
theorems:
Comment on the conditions of the theorem, and any peculiarities
Comment on the conclusions of the theorem, and any peculiarities
Indicate the geometrical and analytical signicance of the theorem
97
Explain the intuitive content of the theorem
Discuss the applications of the theorem
Demonstrate that the theorem holds under the given conditions for an example
Show how the theorem fails for an example where the conditions do not hold
The last entry is included on the grounds that it draws students attention to the con-
clusions, and it is, as a general rule salutary to see when theorems dont hold as well as
when they do ([53]).
4. Completing the square is difcult conceptually for many students at the elementary level,
for a number of reasons - they dont always realise how important it is and so perhaps
do not take it that seriously. However, they also have difculties because amongst all
the trickery involved in completing the square it is not always clear to them what are
the absolutely essential key ideas involved. These are simply thorough facility with the
expansion (x + a)
2
and use of the 0 = A A ploy. These are the two building blocks for
completing the square. If the students are totally familiar with them, then they should
have little trouble with completing the square. If either is shaky or unfamiliar to them,
then they are likely to struggle with the method. So before embarking on the methods
ram home these ideas, keep referring to them in the course of the process, and list them
as the key ideas at the end.
Exercise
Think about your next few lectures. In each case write down the three most signicant ideas in the
lecture, the foundations of the topic. Will the students be able to pick these out and appreciate them?
How can you ensure that these key ideas really do come across? If there are more than three such big
ideas, are you sure that the students will be able to cope with them all in the time available?
Bring out and highlight the essence of a complicated idea, theorem, method, etc and
support the development of students intuitive views of formal arguments
Mathematics is full of complicated ideas which can be overwhelming to the novice, but most of them,
certainly at undergraduate level, can be explained in quite simple ways, almost accessible to the educated
layman (See Devlin [30], for convincing evidence of this). Try to give the student an intuitive feel for an
idea - what are the core points? One way to do this is to treat it as a challenge for yourself to explain
a difcult bit of mathematics to a non-mathematical friend. Mason ([53], p25) refers to a description
that captures the essence of an idea as an intensive denition. The formal denition is referred to as
extensive.
Tall ([69], p17) emphasizes the importance of helping the student, in their development from pre-formal
mathematical thinking to a more formalised approach, by aiming at developing a preliminary insight into
what is going on. This is often the reverse of the natural approach of the experienced mathematician who
tends to break down difcult and extended ideas into palatable chunks, develops these and then puts the
whole together. But when the novice sees these chunks, they are ignorant of their motivation and their
eventual destination, or the whole picture into which they have to t. At rst sight of these chunks they
may in fact, in the absence of guidance, come to think of them in ways that inhibit later inclusion in a
wider picture. By bringing out the essence of the idea from the start, and continually tting the chunks
into it, this can be avoided.
98
Examples
1. As an example of the dangers of chunking without a guiding insight into an idea Tall
([69]) considers the traditional teaching of the denition of a derivative by the following
steps:
(a) dene notion of a limit
(b) for xed x consider the limit of (f(x + h) f(x))/h as h tends to zero
(c) call the limit f

(x), then allow x to vary to give the derivative as a function


As Tall notes, the student is often stumped at step (a), with the idea of a limit plucked out
of the air for no apparent reason. Other possible cognitive obstacles in this development
are discussed, but the main point is clear. The lecturer fully understands the big picture
and knows where they are going with these steps. They knowthat essentially all they are
doing is approximating a rate of increase at a point by the average over a small distance
from the point to a neighbouring point. They are then letting the neighbouring point
approach as close as we wish to the initial point and then assuming that this will give us
a nite value that gives the derivative at the point. They also know that they can choose
any point x on the curve for this process and so the derivative will itself be a function of
x just like f(x). The last three sentences constitute all that is needed, with sketches on the
board, to convey the essence of the idea of the derivative to the students, as a preliminary
to working through the steps (a)-(c). To the uninitiated student however all this insight
is hidden behind the jargon and so they have no overall idea of where they are going.
2. The adjoint matrix method of inversion of matrices can seem daunting when displayed
as an algorithm. But with the use of one or two simple examples it is easy to give a
preliminary insight by showing how for example the cofactors simply come from the
systematic elimination of each variable, as in Cramers rule. It is essentially a notation
designed to facilitate this recipe.
3. In the solution of second order differential equations by the Frobenius method ([19]) the
details of choosing the form of the series and the strange emergence of the log term can
be mysterious to the beginner. In this case a preliminary feel for the essence of these ideas
can be obtained very simply by working through the analytical solution of the Cauchy-
Euler equation. In this case the general features that arise in the Frobenius method come
out explicitly and the log term is no surprise. So in this case the essence of a complicated
idea is conveyed by a much simpler analogy.
4. Mason ([53], p. 52) suggests crystallizing the essence of a theorem by going back over the
proof of a theorem, illustrating the structure and describing the salient features, along
with an overview of the role of the theorem in the module as a whole. After doing this
a few times one may then be able to encourage the students to do it themselves, without
prompting from the lecturer.
Exercise
Actually, there are not that many ideas per module that either need their essence exposing, or for
which it is very difcult to do so. However, there will probably be some in your modules. Identify
these and nd ways to convey the essence, using verbal explanation, board-work, analogies,
examples, etc.
99
Emphasize the respectability of guessing and the necessity of checking
Often the formal, nished, polished presentation of mathematics gives an impression that it always
moves with inexorable logic from A to B with no intervention from the imagination. Of course we know
this is not the case - but the students may not. In mathematics we have phrases for making guessing and
leaps of faith seem respectable - conjecture, by inspection, trial and error taking an ansatz, method
of undetermined coefcients, etc. All are grand clothes for nothing other than gloried guesswork -
educated though it may be. Share this secret with the students. Encourage, indeed insist, that they get
used to guessing and help them to lose the fear of being wrong. Make mistakes - but quickly!. In such
guessing, get them used to asking questions such as What would be the simplest thing to do in these cir-
cumstances?. And of course, most importantly, they must realize the imperative of checking. Baumslag
([7]) refers to this as the Do the mathematical two-step, guess-check, guess-check. In some ways this
provides a reason for not giving answers or solutions to those exercises that can be checked - for example
the solution to a differential equation does not need to be given, since the students can (should) check
it themselves by substituting back into the equation. This not only gives them practice in differentiation
and hammers home the concept of the solution to a differential equation, but also gives them condence
that they can nd their own way about mathematics.
Examples
1. The method of undetermined coefcients in nding particular integrals for inhomoge-
neous linear differential equations is a classic example of guessing made respectable.
Of course, it is so familiar to us now that it hardly seems like guessing - but that is what
it actually is, and this should be pointed out to the students. By inviting them to make
guesses themselves, before showing them the correct choice, you can give them a feel
for the whole idea of progressing in mathematics by trying out the obvious avenues. And
of course in the method of undetermined coefcients we are forced to check in order to
nd the coefcients.
2. A similar example to Example 1 is the repeated roots case of the auxiliary equation
in solving second order differential equations with constant coefcients. The auxiliary
equation gives us only one solution so we are forced to look elsewhere for another. At
the elementary level we cant justify the correct choice by for example the Jordan Normal
Form, so it just has to be treated as the obvious guess - which we later check by insertion
in the equation.
Exercise
In most areas of mathematics there are some critical stages or techniques that lend themselves to
educated guesswork - by inspection, assume a solution of the form..., etc. Identify some of these in
your course and prepare the corresponding lectures so as to highlight these to the students. Use
them, to emphasize that mathematicians often proceed in this way, by trying out simple guesses - but
of course they immediately check them.
Alert the students to patterns that may assist their understanding and learning
Some people dene mathematics as the science of patterns, and certainly patterns pervade the subject.
Students should always be alerted to patterns in whatever they are doing and they should be encouraged
to search for them themselves. Of course, the patterns can be of all sorts of different types - algebraic,
100
geometric, logical, conceptual, numerical, etc. Fortunately this is one topic that is now given considerable
prominence in UK school mathematics teaching, so students should appreciate the idea quite readily.
Examples
1. See the pattern used to summarise binomial theorem pattern in Section 2.8
2. Baumslag ([7], p. 155) notes how the right picture can illustrate the structure of an idea,
giving the well-known example of the proof of the countability of the rational numbers.
3. Another type of pattern in mathematics is the generalisation - perpetuating the present
pattern. The simplest examples are such things as sequences - like 1, 4, 9, ..., but there are
others much more powerful. For example generalizing the Pythagorean length

x
2
+ y
2
from the plane to R
n

x
2
1
+ ... + x
2
n
. Or tensor calculus as a generalisation of vector
calculus. All these are generalizations of a pattern. In this case one can carry the pattern
further into abstractions such as metric spaces or groups. See Tall ([69], p. 11) for the
distinction between generalization (easy extension of a pattern) and abstraction (hard
reconstruction of a pattern).
4. Mason ([53], p15) discusses a tactic that encourages students to develop and express
generalities.
Exercise
Identify useful patterns in your own teaching that may illustrate important points in your lecture.
Help students to build up connections by providing roadmaps and overviews of dif-
cult sequences of arguments - try to express proofs sensibly, as much as logically
This is related to the previous item about patterns, but particularly relevant with a long proof. You can
write down a summary of the steps to begin with, or provide a handout. Try to make each step sensible
(not necessarily the same as logical) - why would you do that, why did the originator do that? Sometimes
the standard proof is presented in a sanitized, perfect, logically most efcient way that often prompts the
question Why?! from any curious student. Sometimes it may be best to use a different form of the proof
that does make each step fairly obvious. And the classic rabbit out of a hat trick of pure mathematics
where a form of expression is selected, or a particular condition is stipulated, can be exposed for what it
often is - a device to make the proof work.
Examples
1. The proof of a major theorem such as Cauchys theorem cries out for some sort of road-
map linking all the steps and showing how they build up into the nal picture. You
can treat it as a group exercise within the lecture, only going through it for the students
yourself when they have had a good try. Here the basic idea is to build up the proof by
rst deriving it for triangles, using a reduction process on smaller and smaller triangles
to provide estimations for the integrals required. This can then be used to derive the
anti-derivative theorem and hence Cauchys theorem.
2. A similar example is the proof of Lagranges theorem in Group Theory. In this case the
structure is one of continual searching for cosets by a repetitive process until the group is
exhausted. The connections and structure of the proof can be lost in the technical details
if attention is not drawn to it.
101
3. Mason ([53], p. 54) advocates using tree diagrams to illustrate the structure of a proof
and to bring together the various ideas and steps.
Exercise
Choose a few of your lectures and study the major ideas in each of them. Can the complexities of
these be illustrated by appropriate roadmaps and overviews? Devise exercises that might help the
students to construct these themselves.
Alert students to the fact that some teaching and learning is provisional, and may
have to be revised at a later date
As higher levels of mathematics are reached it becomes increasingly necessary to compromise between
exactitude and economy of explanation. Due to time or other limitations we sometimes have to be a little
economical with the truth in the interests of progress through the material. We perhaps avoid a subtlety
that the students will revisit in a later course. On such occasions at least warn them, as a matter of intel-
lectual honesty, that that is what you are doing, and that you cannot at the moment give them the whole
picture. The students dont have to believe that everything they learn is the absolute last word - indeed,
informed scepticism is an essential asset for any educated person. The rst two examples below are cases
where at some stage in a childs mathematical education they are taught rmly established facts usually
without qualication, which are later overturned to reveal whole new realms of mathematics. Rules in
mathematics are made to be broken! John Bell - All no-go theorems show is a lack of imagination.
Examples
1. Students do not always realise for themselves that such things as Pythagoras, 180 degrees
in a triangle, etc are restricted to plane geometry and do not for example hold in general
on a sphere. But then, when alerted to this fact, they may protest that one would not
hesitate to calculate the diagonal of a eld using Pythagoras, despite the fact that it is
on a spherical earth. This is an example where their previous teachers have had to be
incomplete in their explanation to avoid muddying the waters. Point out that this will
often happen, because we are continually adapting to newcircumstances in mathematics.
We cannot at every stage go into all the detailed restrictions on our statements.
2. Ditto for the oft-quoted assertion in early school mathematics that we cant take the
square root of a negative number.
3. Tall [69] makes the point that we teach students a great deal about solving differential
equations analytically, and this might give them the impression that most equations can
be solved in this way. Of course, the number of differential equations that can be solved
analytically is negligible and they are the exception - thats why we need numerical meth-
ods. We can explain this to the students and perhaps show them a few differential equa-
tions that cant be solved in analytical form.
Exercise
When are you being economical with the truth in your lectures - do you need to give any health
warnings? You often feel this when you make some statement and then immediately think to
yourself Ah, but, ... At this point it is probably best to be open with the students and add a qualier
There are occasions when this breaks down, but these cases dont concern us here.
102
We have deliberately gone into great detail in this subsection on explanation, because it is so fundamental
to teaching. Good skills in explanation will not only help the students, but also save you time by enabling
you to progress more rapidly and reduce the likelihood of students coming to you for clarication.
3.11 Evaluating and Developing the Lecture
Even the most experienced lecturer sometimes nds something to improve in a lecture they may have
given a number of times. Certainly, in your rst years of teaching, every lecture is actually a learning
experience as much for you as for the students, and so each one needs to be evaluated so that you can
develop it further and improve your teaching. The Higher Education Academy UK Professional Stan-
dards Framework (http://www.heacademy.ac.uk) holds as one of its professional values a willingness
on the part of practitioners to regularly evaluate and develop their performance and teaching practice.
Your institution may have mechanisms for this as its own quality assurance and enhancement policies.
So it is likely that you will have some sort of formal evaluation of your lecturing in your rst few years of
teaching at least. This is NOT the subject of this section, although its results may of course feed into your
development. Such formal arrangements will usually follow some specied protocol that will normally
be explained to you through your institutional staff development unit. What we are referring to here is
the continual day to day process of learning and self-improvement that most academics take as a matter
of course.
In fact this area is important, because it is how we measure the effectiveness of the impact of any training
we receive. Baumslag ([7]) gives a delightful and memorable analogy to illustrate this - the difference
between a good and an indifferent carpenter. The good carpenter sharpens his tools before putting them
away after nishing a job. So the good lecturer, on leaving a lecture, spends a few minutes jotting (or
tapping!) down the main topics of the lecture, good and bad points, ideas for doing things differently
next time. This takes a little time, but is invaluable. You may put off doing this, because you think the
things are so clear that you will remember them - but you probably wont. Occasionally give the students
a very quick feedback form to see how things are going. By now you will be getting a fairly good idea
of what makes for a good lecture in theoretical terms, so construct for yourself a checklist of the sort of
things you might want to think about after the lecture.
Wankat and Oreovicz [72, p 321] recommend setting minute papers at the end of a lecture, containing
such questions as What is the most important thing you learned today? or What questions do you still
have.
And, of course there is the minute by minute micro-evaluation we do throughout the lecture, in which
we watch the students to get a feel for how the class is going. It helps if you can cultivate a couple of the
students who you can rely upon to give you feedback on progress. If students make comments, good or
bad, dont take it personally. During the lecture is by far the easiest time for a student to sort out problems
and misconceptions, and almost certainly other students will benet from the issues addressed. This is
one reason why you should identify in advance the difcult areas where you know most students will
have a problem, then you can address them there and then.
Exercise
Design a short checklist for use in evaluating how a lecture has gone, including questions for the
students and for yourself.

You might also like