Untitled
Untitled
CALCULUS
With a Supplement on Mathematical Understanding
Other World Scientific Titles by the Author
Terrance Quinn
Middle Tennessee State University, USA
Zine Boudhraa
Montgomery College, Maryland, USA
Sanjay Rai
Montgomery College, Maryland, USA
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is
not required from the publisher.
Printed in Singapore
Terrance Quinn: In memory of my parents, George and Bernice, and two of
my brothers, Patrick and John, all of whom encouraged me in my interest in
mathematics and science, from an early age.
Zineddine Boudhraa: In memory of my mother, Habiba, as well as the
memory of all of those who lost their battles to cancer.
Sanjay Rai: To my parents, Shri Sarva Deo Rai, my late father, and Smt.
Munni Rai, my mother, neither of whom had the opportunity for education,
but through their own experiences in life, understood the need and potential
of education in changing the lives of a generation.
We also dedicate this book to students and teachers, in the hope that it will be
enjoyable and helpful in your growth in mathematics.
Preface
This book was developed from a series of lecture notes for teachers of
undergraduate vector calculus, and students who want to understand the
essentials of the classical theorems. With that said, we believe that graduate
students in mathematics and mathematical sciences generally will also find
the material helpful. In order to reach some level of mastery of, say,
differential forms, one needs to have source insights for the classical vector
calculus theorems. Promoting those source insights is the main purpose of
our book. To help set the stage, we begin by providing some context.
where
Note, however, that what happens to paddle wheels in water flow is only
settled by experiment. As it happens, if a handle with a paddle wheel is
placed horizontally, then in some cases the paddle wheel rotates, while in
other cases it does not. Whether or not a paddle wheel rotates also depends on
orientation of the handle relative to the direction of flow. The integral
definition given above comes from classical fluid dynamics and, with some
work, can lead to the 3-d vector curl formula given in Section P.2. Discussion
of “imagined push and pull of water flow” notwithstanding, without
computing the link from the integral to the vector form, the student is
directed away from the possibility of basic insight and here, too, is instead
moved along into a flow of exercises for symbolic techniques.
While descriptions of water flow can invite beginnings, it took physics
decades to formulate key quantities and vector formulas. There are sequences
of insights needed, to go from imagining and describing water motion to
understanding such motion in terms of vector fields, vectors normal to a
plane in which “rotational displacement” occurs, combinations of partial
derivatives and line integrals.
P.8 Acknowledgments
We are indebted to colleagues of Sanjay Rai and Zine Boudhraa at
Montgomery College, MD. Motivation for this book partly emerged in the
context of conversations regarding possible redesign of particular elements of
the undergraduate mathematics curriculum. Some of these discussions
regarded the teaching of single- and multi-variable calculus, and the
importance of understanding, and being in control of concepts. The
divergence and curl operators especially illustrated the need for insight
regarding what they convey about a vector field. We would like to thank our
students who help us grow as teachers. Sanjay Rai would like to thank his
professors from Allahabad University, Dalhousie University and the
University of Arkansas for their tutelage and inspiration. Grateful to all of his
teachers, Terrance Quinn would like to thank three, in particular, who
introduced him to the vector calculus theorems, in various contexts: Professor
Rooney (1925–2016), in Advanced Calculus (University of Toronto, Fall and
Winter of 1981/82). Rooney opened our eyes to riches of classical
mathematics; Professor Keith Johnson (Dalhousie University) helped us learn
about differential forms (Spring, 1987) from Michael Spivak’s deceptively
brief Calculus on Manifolds11; and Professor Lowell Jones who (in a
graduate course at SUNY at Stony Brook, Fall, 1988) shed light on the
significance of the vector calculus theorems in 20th century differential
geometry.12 Finally, we’d like to thank two anonymous reviewers for their
comments and suggestions which helped improve the book.
_________________________
1In this context, closed means that when parameterized in a permissible way (for instance, one-one
and differentiable) the initial and final points of r(t) = (x(t), y(t)), t0 ≤ t ≤ t1 are the same, that is, (x(t0),
y(t0)) = (x(t1), y(t1)). For the term positively-oriented, non-technically, one can imagine a
parameterization tracking counter-clockwise relative to a point in the region R bounded by the curve.
The “circle” on the integral symbol indicates that the curve for integration is closed. To compute the
line-integral , we use any permissible parameterization r(t) to get
The integral of curl(a, b) is the familiar 2-d integral from 2-variable integral calculus. In order for the
formulas to be well-defined, it is crucial that the integrals involved be invariant under permissible
changes of variables. That follows from the chain rule in finite dimensions. Details are brought out in
the book.
2One may explore the literature.
3For more detailed observations, see the Supplement.
4Rigorous definition of orientation requires a more advanced context. See Supplement, note 31, for an
indication of the importance of identifying “casual insights.”
5See Supplement, Section A.4, for more details on “proof” and “middle terms.”
6This includes “calculus reform.” See Sections P.2 and P.3. The books by Snider and Davis (2000)
and Lovrić (2007) both include some pedagogical strategies that are “on point.” See Section P.6.
7Regarding the historical origins of “Green’s theorem,” Katz observed that “[a]ll of the
mathematicians who stated and proved versions of this theorem were interested in it for specific
physical reasons” (Victor J. Katz, “The History of Stokes’ Theorem,” Mathematics Magazine, vol. 52,
no. 3 (May, 1979), pp. 146–156. See p. 149.)
8Ed Dubinksy, “Using a Theory of Learning in College Mathematics Courses,” MSOR Connections
(2000), 1.10.11120/msor.2001.01020010. This article was originally published in Newsletter 12,
TaLUM, Teaching and Learning Undergraduate Mathematics subgroup, 2001. It is available online:
www.math.wisc.edu/wil-son/Courses/Math903/UsingAPOS.pdf. Accessed July 17, 2019. See
Supplement, Part C, for discussion.
9The approach taken is guided by the historical development in mathematics and mathematical
physics. It is the approach taken in modern differential geometry, where streamlines and velocity fields
are the basic elements. Circulation is an average weighted velocity along a curve. Force is a second
derivative. Work is an average weighted force along a curve.
10See Katz, 1979, “Generalization and unification,” pp. 150–156.
11Michael Spivak, Calculus on Manifolds. A Modern Approach to Classical theorems of Advanced
Calculus. Reading, MA: Addison-Wesley, 1965.
12The textbook used was the classic: Noel J. Hicks, Notes on Differential Geometry. New York: Van
Norstrand Publishing, 1965.
Contents
Preface
P.1 General context
P.2 Beginning with the coordinate definition of the 3-d vector curl
P.3 Beginning with descriptions of imagined paddle wheels and
integral definitions of circulation
P.4 Do vector fields represent force or velocity? Or other?
P.5 Historical context and pedagogy
P.6 Prerequisites, the lessons, and follow up
P.7 Our experience
P.8 Acknowledgments
7. STOKES’ THEOREM
7.1 Stokes’ Theorem: The Question
7.2 Stokes’ Theorem: First approach: elementary derivation of
“Green’s theorem” for a two-parameter surface
Index
Chapter 1
The 2-d and 3-d vector calculus theorems are extensions of the Fundamental
Theorem of Calculus in one variable that students encounter in a first course
in calculus. To understand the vector calculus theorems, one needs a good
understanding of the one variable case. Sections 1.1 and 1.2 review some key
ideas in single variable calculus, leading up to the fundamental theorem in
one variable. Section 1.3 is to help bring out why the fundamental theorem of
calculus is called “fundamental.” (Among other things, it led to the solution
of numerous previously unsolved problems. It also opened up new lines of
inquiry.) Sections 1.4 and 1.5 foreshadow how the fundamental theorem in
one variable is both clue and key for the vector calculus theorems. In the 2-d
and 3-d cases, the theorems involve areas and angles between vectors.
Sections 1.6 and 1.7 are for developing the dot product, and the vector
product in three dimensions. The dot product and the vector product are both
solutions to problems in classical coordinate geometry. The cross-product
rule is a technical result that will be needed. Sections 1.8, 1.9 and 1.10
discuss three main cases. Section 1.11 draws attention to implications in
integration over a region in the plane.
This chapter is for development in prerequisite material. If you already have a
good understanding of one variable calculus and other topics treated here,
you could go directly to the beginning of Part 1, Fluid Motion in 2
Dimensions. You could then re-visit topics treated in this chapter on a “need
to know basis.”
Figure 1.1 Galileo’s experiments. Rolling steel balls down wooden beams. Measure distances and
times.
Straightforward computation shows that for s = t2, average speed in the fourth
second is larger than average speed in the third second, average speed in the
fifth second is larger than average speed in the fourth second, and so on. In
other words, free-fall is “accelerated motion.”
A question arises: What is the speed of a free-falling object falling near a
particular time, near, say, t = 3?
Figure 1.2 Distances, times and average speeds are conveniently represented in a graph. Average
speed for a time interval is slope of a secant. Slope of secant for t = 3 to t = 5 is .
Students who have been through a first course in calculus will be familiar
with the following computations.
For t = 3, and with the idea of improving accuracy in computing average
velocity, we use increasingly small time intervals.
In units of , for the time interval t = 3 to t = 3 + 1, average speed is
And so on!
That is, for the time interval t = 3 to t = 3 + h, average speed is
On the graph of s = t2, these are slopes of secants that get increasingly close
to being tangent to the graph at t = 3.
The algebra reveals that the smaller the time interval h, the closer [2·3+(h)] is
to 2 · 3, that is, the closer the average speed is to 2 · 3 = 6. Saying it another
way, as h approaches zero, the difference between the difference ratio
and 6 approaches zero.
Figure 1.3 For h small, the secant is close to being the tangent to the graph at (3, 9); and the slope of
the secant is close to being the slope of tangent line.
Figure 1.4 Linear approximation to y = f(t) is the tangent line given by y = f(t0) + f′(t0)(t – t0). The
time difference t – t0 = h.
More generally, the derived quantity is called the derivative at t. Since the
derivative usually is computed for more than one value of is
also called the derivative function.
Derivatives and Anti-derivatives: Once the derivative is understood, we can
study the derivative of sums, products and quotients of functions. That way
we can build up a repertoire of derivative formulas. By the same token, we
can work out tables of “anti-derivatives.” For instance, . If, then,
we start with g(x) = x and ask for the function f(x) that satisfies , the
solution is .
Exercises
Exercise 1.1.1
Suppose a ramp is at an angle so that Galileo’s result for “free-roll” is s = t2
feet in t seconds. (If you do a course in Newtonian physics, you will be able
to compute the angle at which that occurs. For now, suppose that in Galileo’s
experiment, free-roll s = 16 sin (θ) t2. Find the angle for which s = t2.)
(a) Compute the average speed for the time interval t = 6 to t = 6.5.
(b) Compute the average speed for the time interval t = 6 to t = 6.1.
(c) Compute the average speed for the time interval t = 6 to t = 6 + h, where
h is any small positive number.
(d) Compute the “instantaneous” speed at t = 6? Explain.
Exercise 1.1.2
Do Exercise 1.1.1 for a more general case. That is, as is done in the book for
s = t2, use only “first principles” to find the “instantaneous speed” at any time
t during free-roll.
Exercise 1.1.3
Suppose that a metal shuttle attached to a spring slides along a smooth
surface (friction is negligible) and that, for a time, its motion approximately
satisfies s = t3 feet in t seconds. Compute a good estimate for the
“instantaneous speed” at time t.
Exercise 1.1.4
Suppose that . Using only first principles show that
Exercise 1.1.5
Suppose that f(x) = x3. Using only first principles show that
Exercise 1.1.6
Suppose that f(x) = cos x and g(x) = sin x, where x is in radians. Using only
first principles, show that
Hint: What are the addition formulas for cos (x + h) and sin (x + h)? Looking
to the unit circle, for h ≈ 0, h ≠ 0, what are geometric approximations for
and ? Remember what radians means, a unit distance. What is that unit, in
the case of a circle of radius 1?
If, however, we start with a square of length x = 10 and again increase length
by dx = 1, then
Figure 1.5 Two rolls of paper, of width 24′′ and 40′′ respectively. Rolling forward, the rate of change
of area is (24 sq. inches)/(inch advanced) and (40 sq. inches)/(inch advanced), respectively.
In other words, for a given roll of paper, the rate of change of area is the
width of the roll!
Let’s now go back to the question about square areas. After some progress in
differential calculus, we have that
Figure 1.6 [Area of square of length ] - [Area of square of length 3] = [Areas of two
rectangles] + [Area of corner piece].
Figure 1.7 [Area of square of length (3 + h)] – [Area of square of length 3] = [Areas of two rectangles]
+ [Area of corner piece].
That is:
Relative to change in length x, the instantaneous rate of change of a square
area A(x) = x2 is the length of the advancing front line. In the case of a square
of length x, that length happens to be 2x.
Without yet having a rigorous definition of area,18 the same question can be
asked about an area A(x) between the graph of a non-negative function y =
f(x) ≥ 0 and the x axis.
Notation: The symbolism for ‘area under a graph’ was introduced by Leibniz
(1646–1716):
For the interval a ≤ t ≤ x, the area A(x) is written
Figure 1.8 Leibnitz heuristics and notation: Area A(x) under graph is a sum of areas of the form
“f(t)dt”: .
Figure 1.9 A(x + h) – A(x) ≈ f(x) · h + [area of small triangle]. Small triangle is determined by the
tangent line at (x, f(x)).
Exercises
Exercise 1.2.1
Exercise 1.2.2
Exercise 1.2.3
Exercise 1.2.4
Exercise 1.2.5
Exercise 1.2.6
Exercise 1.2.7
Exercise 1.2.8
And so on, that is, formulas can be obtained for areas of figures, when the
figures can be constructed from rectangles, parallelograms and triangles.
Partly because they all are reducible to rectangles, the answers are mutually
compatible.
However, what is the area of a circle of radius r? This doesn’t fit into the
system so easily. We can make a beginning by observing that a planar circle
is defined by a center C and its radius r. We can then use that radius to
construct 4 squares to cover the circle. By insight into the diagram, the area
of four squares each of length r is greater than the area of the circle of radius
r.
Figure 1.10 Area of circle of radius r is less than 4r2.
Observe that the argument does not depend on the radius. And so we
conclude that there is a ratio, call it π, 3 < π < 4, such that the
There are many other figures in Euclidean geometry. Archimedes solved the
problem of “quadrature of the parabola.” He obtained a formula for the area
of a parabola without needing to introduce unknown constants like ‘π.’ By
subtraction, we then also get a formula for area between a parabola y = x2 and
the x-axis. To get his result, however, Archimedes had to develop the theory
of non-terminating geometric series. He needed and reached an
understanding of “limit of a sequence of geometric sums.”22 Note that all of
this was done about 2000 years before equivalent results were (re-)
discovered by European mathematicians.23
If there had been Nobel Prizes given out in in ancient Greece, we venture that
Archimedes would have deserved several, one of which would have been for
solving the problem of the quadrature of the parabola.
What about other areas? What are areas under the graphs of y = x3, y = x4, · ·
· ; under the graphs of trigonometric functions like y = sin x, y = cos x, y = tan
x; derived trigonometric functions like y = sec2 x and y = csc2 x; and
exponential functions like y = 2x? Does each type of type of function require
its own definition and perhaps its own Nobel Prize?
Figure 1.11 The Fundamental Theorem of Calculus provides the basis for computing areas. (a) A(x) =
ex. (b) A(x) = x5/5. (c) A(x) = ln(x + 0.5).
Exactly similar arguments provide areas under the graphs of the other
functions mentioned above, and many more. That is, in each case, the
possibility of obtaining an explicit formula for area depends on having
techniques for finding anti-derivatives.
Exercises
Exercise 1.3.1
Graph y = f(t) = t2, t ≥ 0.
The problem is to find the area under the graph of the parabola. As explained
in the text, Archimedes solved the problem by developing the theory of
convergent geometric series. Using derivatives, the solution to the problem is
elegant and extends to an indefinite range of other functions and applications.
(a) Draw the area under the graph of y = f(t) = t2, for the interval t = 0 to t =
5. What is the length of the vertical front-line?
(b) Without doing “calculus”, what is an approximation for the area swept
out, if the vertical front-line moves forward by Δt = 0.1?
(c) What is an approximation for the area swept out if the front-line moves
forward by “small” Δt = h > 0?
(d) What is the instantaneous rate of change of area at t = 5?
(e) Repeat (c)-(d) for the interval t = 0 to t = x. That is, using first principles,
what is the instantaneous rate of change of A(x), the area under the graph
of y = f(t) = t2, at t = x? In other words, what is ?
(f) Using (e), what is the formula for the area A(x) under the graph of y = t2,
t ≥ 0, at t = x? This reproduces Archimedes’ result.
Exercise 1.3.2
Repeat Exercise 1.3.1 for the areas under the graphs of the following
functions:
Figure 1.12 Linear approximation of area A(x): A(x) ≈ A(xi) + A′(xi)(x − xi).
So, does the Fundamental Theorem of Calculus really only depend on linear
approximation?
Discussion: The first two equalities,
approximates area under the graph of a function, a function that in the present
context happens to be y = A′(x).
In the Question, it is the third equality where we find what might seem to be
“sleight of hand.” Why can we write
is partly, but not merely, a consequence of linear approximation. That is, the
ingredient that provides the third equality is the fact that the derivative of area
under the graph of the function y = f(x) > 0 is the length of the front-line of
advancing area.
Exercises
Exercise 1.4.1
Using equation 1.2, find the formula for the area under the graph of y = f(x) =
x, over the interval 0 ≤ x ≤ b. If you draw the graph, you will find that you
already know the answer from classical geometry. The exercise here is to find
it using the new method. Hint: Use subintervals of width . Also, what is
1 + 2 + · · · + n?
Exercise 1.4.2
Using equation 1.2, find the formula for the area under the graph of y = f(x) =
x2, over the interval 0 ≤ x ≤ b. Hint: Use subintervals of width . Also,
what is 12 + 22 + · · · + n2? One way to find this sum is to construct the
following series of equations:
Adding, the left side is a telescoping series while the right side can be written
in terms of the unknown 12 + 22 + · · · + n2 along with two sums that are
already known.
Exercise 1.4.3
Generalize Exercises 1.4.1 and 1.4.2. That is, for k a positive integer, using
equation 1.2 of Section 1.4, find the formula for the area under the graph of y
= f(x) = xk, over the interval 0 ≤ x ≤ b. Hint: What is 1k + 2k + · · · +nk?
Imitate the computations used in Exercises 1.4.1 and 1.4.2. You will need the
binomial theorem.
and that, for each x, g(x, y) is a (linear mass density)24 function in y. This
means that g(x, y)dy approximates the mass of a length dy in the y direction,
based at (x, y).
Figure 1.14 Shaded verticals are to indicate that there is a linear mass-function in the y-direction.
In other words,
And so: A difference of two weighted line elements in the y direction (each of
which is an integrand for a one-dimensional integral) is approximated by a
weighted area element, an integrand for a two-dimensional integral.
Moreover, the weight for the area element is a derivative of the weight for the
line element.
This computation does not provide the fundamental theorems of vector
calculus. But you might keep it in mind as foreshadowing for the
development of the vector calculus theorems.
Exercises
Exercise 1.5.1
Let g(x, y) = x2 y3. Using linear approximation, show
and satsifies
Figure 1.15 Determine the length of the orthogonal projection of v onto w in terms of coordinates a, b,
c and d.
Problem 1.1.
(i) Assuming the Pythagorean Theorem and appealing only to the axioms
of Euclidean geometry, develop the law of cosines. Hint: Extend the
diagram so that you can express the problem in terms of two nested
right angle triangles.
(ii) Apply the law of cosines to the case where two sides of a triangle are
given by coordinate vectors v = (a, b) and w = (c, d) emanating from
the origin. Depending on one’s choice of initial and final point, a third
side of the triangle is u = ±(c – a, d – b).
(iii) Develop a formula for the magnitude of the projection of v = (a, b)
onto w = (c, d), in terms of the coordinates of v = (a, b) and w = (c, d).
Hint: See ii.
(iv) Observe that a key element of the projection formula obtained in iii. is
the combination ac + bd. This combination is called the dot product
and is written v · w = (a, b) · (c, d) = ac + bd.
(v) Recast your results so that you get the well-known formula for v · w in
terms of || v ||, || w || and cos θ, where θ is an angle between v = (a, b)
and w = (c, d).
Exercises
Exercise 1.6.9
u = 2i – j + 3k, v = –i + j + 3k
Problem 1.2.
The classical formula for the area of a parallelogram is (base) × (height). To
compute height, we need the sine function. Starting with the classical
formula and using both the sine function and the Pythagorean formula,
show from first principles that the area of the parallelogram determined by
v = (a1, a2) and w = (b1, b2) is |a1b2 – b1a2|. Hint: Express the sine function
in terms of the cosine function and use Problem 1.1.
Figure 1.16 Determine the area of the parallelogram determined by v = (a1, a2) and w = (b1, b2), in
terms of coordinates a1, a2, b1 and b2.
The quantity a1b2 – b1a2 is called the determinant of the pair of vectors v and
w. The determinant can be written as follows:
Volumes
In (x, y, z) space, three vectors u = (a1, a2, a3), v = (b1, b2, b3) and w = (c1, c2,
c3) determine sides of a parallelepiped.
Figure 1.17 Determine the volume of a parallelepiped determined by u = (a1, a2, a3), v = (b1, b2, b3)
and w = (c1, c2, c3), in terms of the coordinates a1, a2, a3, b1, b2, b3, c1, c2, and c3.
Invoke the previous problem as needed and so work out that, in terms of
coordinates, the volume of the parallelepiped is
The quantity
is called the determinant of the three vectors u = (a1, a2, a3), v = (b1, b2, b3)
and w = (c1, c2, c3).
Similar to the 2-d case, this is often written:
Figure 1.18 Area of parallelogram determined by 2v and w is twice the area of that determined by v
and w.
is four times the volume determined by u = (a1, a2, a3), v = (b1, b2, b3) and w
= (c1, c2, c3). And, if we substitute the coordinates of u = (a1, a2, a3), 4v =
(4b1, 4b2, 4b3) and w = (c1, c2, c3) into the 3-d determinant we get
Problem 1.3.
Make a diagram for the volumes discussed above.
If we switch the order of two sides then areas and volumes do not change.
However, the two vectors representing the sides change in relative
orientation.
Figure 1.19 In the diagram, v to w is a negative rotation, while w to v is a positive rotation.
This shows up in the determinant formula as a change in sign, but leaves the
absolute value invariant. In the 2-d case, the computation is as follows:
Two properties common to both the 2-d and 3-d determinant functions are:
they are linear in each vector component; and they alternate in sign when two
vectors are switched.
Differential forms generalize areas and volumes to higher dimensions. For
readers who plan to go on study differential forms, note that these two
properties are all that is needed to define a determinant function in n
dimensions. In other words, a generalized n-dimensional determinant
function (or n-dimensional volume function) is any real-valued function
For any n, there are many such volume functions. As the following example
illustrates, it is a matter of choosing units.
Problem 1.4.
With e1 = (1, 0) and e2 = (0, 1), define .
(i) Show that the defining properties determine for all pairs of
vectors v = (a1, a2) and w = (b1, b2).
(ii) What is the relation between this and the classical
?
Problem 1.5.
Show that a generalized determinant function is completely determined by
its value on “the unit n-cube,” that is, by the value of D(e1, e2, · · · , en),
where e1 = (1, 0, 0, · · · , 0), e2 = (0, 1, 0, · · · , 0), · · ·, en = (0, 0, 0, · · · ,
1).
Exercises
Figure 1.20 Rigid radial arm attached and orthogonal to an axle. Applied force F = (b1, b2, b3) at end-
point of r = (a1, a2, a3) and coplanar with plane of rotation of r about axle.
The problem is to determine the angular acceleration of the radial arm caused
by the applied force F = (b1, b2, b3).
The structure is rigid. Within the constraints, only rotational motion is
possible. By taking the component of F = (b1, b2, b3) in the plane Π that is
orthogonal to r = (a1, a2, a3), and using the law of the lever, we can
distinguish three cases:
(1) Angular acceleration is maximum when F = (b1, b2, b3) is orthogonal to
r = (a1, a2, a3).
(2) Angular acceleration is zero when F = (b1, b2, b3) is parallel to r = (a1,
a2, a3) (pointing either toward or away from the axis of rotation).
(3) Angular acceleration is between (i) and (ii), when F = (b1, b2, b3) is at
other angles relative to r = (a1, a2, a3). More precisely, using the law of
the lever, angular acceleration is proportional to the area of the
parallelogram determined by the vectors F = (b1, b2, b3) and r = (a1, a2,
a3).
Problem 1.6.
Make diagrams for cases (1), (2) and (3).
Angular vectors
In experiments, it is found that if two forces F1 and F2 are applied, angular
accelerations add. However, angular accelerations are not vectors whose
components represent displacement of the end-point of the radial arm. There
is the question, then, of how to represent angular accelerations by quantities
that add.
In physics, rotation is about an axle which is normal to the plane of motion.
And, in coordinate geometry, a plane Π is represented by a normal vector.
Bringing these two ideas together, “an angular vector” is defined to be a
vector along the axis of rotation normal to the plane Π in which rotational
displacement occurs. Positive and negative angular rotations in Π about the
axle are represented by vectors pointing “upward” and “downward” along the
axis (or axle) of rotation, respectively.25
This is well-known in classical physics and leads to the following
mathematical problem:
To more easily allow for other applications, let’s first replace the symbols F
and r by u and v.
Given u = (a1, a2, a3), v = (b1, b2, b3), find a vector c = (c1, c2, c3) that has the
following three properties:
(1) c = (c1, c2, c3) is orthogonal to u = (a1, a2, a3);
(2) c = (c1, c2, c3) is orthogonal to v = (b1, b2, b3); and
(3) || c || = || (c1, c2, c3) ||= (area of parallelogram determined by u and v).
The new vector is called the cross-product or vector product of u and v and is
denoted by the u × v.
Problem 1.7.
Solve the system of equations (1), (2) and (3) and obtain that
Figure 1.21 Given u = (a1, a2, a3) and v = (b1, b2, b3), the cross-product c is defined to be the vector
that is orthogonal to both u and v, whose length is equal to the area of the parallelogram determined by
u and v, and whose direction is determined by the “right-hand rule”.
Exercises
and
The question is whether or not it possible to use these given ratios to
determine the ratio
Of course, the ratio inches to feet is familiar and known. But the question
here (a special case of the chain rule) is to use the given ratios — inches to
yards, and yards to feet — to determine the ratio inches to feet.
Why is the question plausible?
The first ratio relates inches and yards. The second ratio relates yards and
feet. There is, then, a link between the two ratios, a common unit, yards.
The trick is to take advantage of that common unit.
One way to do that is as follows:
The common unit “yard” cancels and the arithmetic provides a ratio of
“inches to feet,” that is, , or in the symbolism of proportions, a 12 : 1
ratio.
Example
Suppose that a car gets highway mileage of
Clearly, the product of these two ratios will be in the required units, .
In other words, the ratio is obtained as follows:
More generally, ratios change. But that is precisely what calculus is designed
to handle.
Example
Let y = x2, x = x(t) = 5t + 4 and consider the composition
We have
and then apply derivative formulas to the resulting quadratic function. The
question here, however, is different.
While we are only using elementary functions, the problem anticipates the
case where functions involved are more sophisticated. They could be high
degree polynomials, trigonometric functions, and so on.
And so the question is: Can we use the rates of change of the component
functions to determine the rate of change of a composition function?
Since the composition function is built out of known component functions,
does it not stand to reason that this might be possible? Part of the value of
Leibniz’s symbolism is that it reminds us that while the derivative of a
function is not a ratio, it is a limit of ratios. And, indeed, using linear
approximation, that limit can be used to estimate change in function values.
And so, taking advantage of the Leibniz notation, might we not anticipate
that
But again, the derivative symbols are limits of ratios. So, let’s dig into these
limits in order to get a better grip on this.
To get at , we need to estimate how a small increment Δt impacts the
composition
(2) For a small increment Δx, there is a small increment in y = f(x) = x2:
As we did in the last two examples, we can bring these ratios together.
Algebraically, that means substituting (1) into (2). Hence,
We now have something to prove that can be sorted out by appealing to first
principles, and the definition of “limit.”
The general case is sorted out in exactly the same way. And so we get the
result anticipated.
Exercises
Exercise 1.9.1
A car gets 30 miles per gallon on the highway. On a particular trip, the car is
using up 2 gallons per hour. From the discussion in the text, the speed of the
car is 60 m.p.h. Use the chain rule to find the same result.
Exercise 1.9.2
For and x = (t + 1)2, find .
Exercise 1.9.3
For and x = t + sin t, find .
Exercise 1.9.4
For , find .
Exercise 1.9.5
For , find .
As it happens, units that are more convenient for the carpenter (both for
building furniture and for accounting) are . In those units, the selling price
of the wood is
and given that we need to purchase so many inches of wood (e.g., 120′′), how
do we re-express all of that in inches and
To meet the needs of the carpenter, we can redo this by changing the length
scale from feet to inches.
Let t = inches. The relationship between inches and feet (between x and t) is t
= 12x.
In order to calculate the integral in the new scale, we need to first convert all
terms of the integral to the new scale.
(1) The new interval is t(0) = 12(0) = 0 and t(10) = 12(10) = 120.
(2) The function C(x) is constant and so, numerically, C (x(t)) = 6 and units
are . The units here are not right. Remember, we are foot
converting to inches. However, there is no cause for concern. Much as in
the solution to the elementary case, once we put everything together,
units will “self-adjust.”
(3) Since t (inches) = 12x (feet), an increment dt in inches correlates with an
increment dx in feet by the relation dt = 12dx. In other words,
Substituting into the integral that was given in terms of feet, and tracking
units,
and
In the example just given, on a hunch that the integrand might be brought
under control, let x = t5 + t. Then,
The argument in the general case is the same. By the chain rule,
By the Fundamental Theorem of Calculus in one variable,
The new integrand might then be further analyzed. In some cases, such as in
the example above, for a known function f(x). In that case, by the
Fundamental Theorem of Calculus, we conclude with a numerical solution,
namely, f(x(d)) – f(x(c)).
Exercises
Exercise 1.11.1
Evaluate the definite integral: .
Exercise 1.11.2
Evaluate the definite integral: .
Exercise 1.11.3
Evaluate the definite integral: .
Exercise 1.11.4
Evaluate the definite integral: .
Exercise 1.11.6
The area of the region bounded by and the x-axis for 0 ≤ x ≤ 1.
Figure 1.24 The curve z(t) = f(x(t), y(t)) has a tangent line and slope at (x(t), y(t)).
The problem is to determine the slope of the tangent line of the curve z =
f(x(t), y(t)) in (x, y, z) space. In other words, what is where z = f(x, y) and
each of x and y are functions of t, x = x(t) and y = y(t)?
Figure 1.25 To compute net change, use Δx, Δy and partial derivatives of z = f(x, y) to compute Δzx-
direction and Δzy-direction. Then add.
Using the one-dimensional result together with the diagram, tracking change
in one direction at a time, the result follows:
and
(2) For the small increments in (1), there are two small increments in z = f(x,
y):
and
Dividing by Δt, and taking the limit, the result follows. That is, if z = f(x(t),
y(t)), then
Exercises
Exercise 1.12.1
Exercise 1.12.2
z = sin x cos y, x = 3t, y = t + 4
Exercise 1.12.3
z = ln (x2 + y), x = e–3t, y = te3t
Exercise 1.12.4
Suppose that a curve in space tracks along the surface of the graph of z = x2 +
y2, and that the projection of the curve onto the x-y plane is given by the
curve r(t) = (cos t, sin t, 0). Find the slope of the tangent line to the curve in
three dimensions. Graph your results and explain.
Exercise 1.12.5
Suppose that a curve in space tracks along the surface of the graph of z = x2 +
y2, and that the projection of the curve onto the x-y plane is given by the
curve r(t) = (4 cos t, 3 sin t, 0). Find the slope of the tangent line to the curve
in three dimensions. Graph your results and explain.
(1) by a pair (x, y), where x and y refer to distances along rectangular
coordinate lines relative to the selected origin; and
(2) by a pair (r, θ) , where r is the distance from the origin to the P and θ is
the angle (in radians) from the line y = 0 to the radial line containing P.
There are translation equations going both ways:
To see the “two-variable chain rule” in action, let’s look to the transformation
given by equation (1) immediately above.
Suppose that two points P and Q on the plane are relatively close to each
other. (We can do this because the plane has the Pythagorean distance
function.) We suppose also that the polar coordinates of both P and Q are
known, as are the increments dr and dθ.
As in examples of the chain rule given above, the problem is to determine the
corresponding increments in the variables (x, y). This is a valid question
because x = x(r, θ) and y = y(r, θ).
Given P = P(x, y) = P(r, θ) and Q = Q(r + dr, θ + dθ), what are the
corresponding increments dx and dy by which to (linearly) approximate Q =
Q(x + dx, y + dy) in rectangular coordinates?
Figure 1.27 Given (r, θ) and (x, y), × = r cos θ, y = r sin θ and (r + dr, θ + dθ), what is the new point in
Cartesian coordinates?
Example
Let’s start with the x coordinate. From P to Q there is a change in both r and
θ, namely, dr and dθ. As described in previous examples, through linear
approximation the increment in x is a sum of an increment determined by a
change in r and an increment determined by a change in θ. That is,
Problem 1.8.
Show that in the case of polar coordinates,
The general case
To get the general result, observe that the computations above did not depend
on the fact that r and θ were polar coordinates. The key is that there is a
functional dependence of the form x = x(r, θ) and y = y(r, θ).
For the general case, we write x = x(u, v) and y = y(u, v); and represent
coordinates (u, v) by their own coordinate plane.
The problem is: Given the pair of increments , what is the corresponding
pair of increments ?
As with the special case of polar coordinates, we use linear approximation
given by the Jacobian matrix
Figure 1.28 Given (u, v) and (x, y), x = x(u, v), y = y(u, v) and (u + du, v + dv), what is the new point in
Cartesian coordinates?
The chain rule for two dimensions and two independent variables is
Exercises
Exercise 1.13.1
Show that the mapping (x, y) → (r, θ) is one-one and onto.
In Exercises 1.13.2–1.13.4, use the Chain Rule to find and .
Exercise 1.13.2
z = x3 + xy, x = s + t, y = st
Exercise 1.13.3
z = x + xy + 4, x = s/t, y = s – 3t
Exercise 1.13.4
z = yex/y, x = st, y = t/s
In Exercises 1.13.5–1.13.6, find the Jacobian matrix .
Exercise 1.13.5
x = 2u – 3v and y = 5u + 7v
Exercise 1.13.6
x = 3u + uv and
Figure 1.29 The linear approximation maps a rectangle in (u, v) plane to a parallelogram in (x, y)
plane.
Problem 1.9.
By linear approximation to the transformation
From our previous work on area functions for the Euclidean plane, the area of
this parallelogram is
This provides the area element and we get the well-known “change of
variable” formula for integrals:
If the region S(u, v) in (u, v) corresponds to R(x, y) in (x, y), then
Example
In polar coordinates,
and so
Exercises
(a) Evaluate this integral and sketch the region D of integration in the xy-
plane.
(b) Let u = x + y and v = x – y. Find the region S in the uv-plane that
corresponds to D.
(c) Use the change of variables formula to evaluate the integral by using the
substitution u = x + y, v = x – y.
Exercise 1.14.2
This exercise concerns the iterated integral
(a) Evaluate this integral and sketch the region D of integration in the xy-
plane.
(b) Let u = x + 3y and v = x – y. Find the region S in the uv-plane that
corresponds to D.
(c) Use the change of variables formula to evaluate the integral by using the
substitution u = x + 3y, v = x – y.
Exercise 1.14.3
Evaluate ∫∫(3x − 2y)dA, where D is the region enclosed by the ellipse 9x2 +
16y2 = 144. Change the variables using and .
Exercise 1.14.4
Transform the given integral to one in polar coordinates then evaluate the
polar integral.
(a)
(b)
_________________________
13How did he measure “time”? Galileo describes using both a pendulum and a water clock.
Discovering a correlation between measured distances and times was the beginning of modern physics.
Regarding Galileo’s work, there remain unanswered questions. See, e.g. Lindberg, David C. “Galileo’s
Experiments on Falling Bodies,” Isis 56, no. 3 (1965): 352–54. http://www.jstor.org/stable/228111.
14In modern terminology, the constant k is proportional to sin(angle between beam and ground). This
Figure 2.1 Four test particles with velocities at locations (x1, y1), (x2, y2), (x3, y3), (x4, y4).
For this book, we are assuming no “turbulence”26 and that the surface
remains (more or less) level. In fact, there are usually waves of at least some
amplitude. However, at low speeds, and if measurements need not be
extremely refined, then vertical motion can be eliminated from
approximations. Note also that by using submerged test particles and
transparent walls in a laboratory chute, it is found that velocities below and
near the surface are approximately equal to surface velocities. For now, then,
we focus only on horizontal components of motion of a shallow water flow.
Note, however, that the velocities of the surface are also understood to
represent the velocities of water below and near the surface.
To get started, then, what we need is a 2-component velocity vector field for
the horizontal velocity of a test particle at location (x, y) on the surface at
time t. That is, the velocity vector is of the form
Remember, however, that velocity is also the velocity of test particles below
and near the surface. That is, locally, the velocity is independent of depth z
which leads to the notation
And actual water has breadth and depth.
As in classical physics, we make use of the idealization wherein we speak of
“points” that are “locations without breadth or depth.”
Problem 2.1.
In this book, we will not be studying the case where there is time-
dependence. Instead, we will focus on cases where v is of the form
The velocity field is constant in time t but need not be constant in space
coordinates (x, y).
(1) Can you imagine elementary cases where the velocity vector field is
independent of time but is not constant in its space coordinates? Hint:
Imagine a chute or ramp down which water flows due to gravity, and
the chute then levelling out further downstream.
(2) Can you produce explicit velocity fields
Problem 2.2.
By using small time increments and first-order terms, work out that speed
of a surface near (x, y) is given by
Hint: With projections, treat one component at a time and use the Euclidean
distance function on the 2-dimensional surface.
Exercises
In Exercises 2.1.1–2.1.10, graph the velocity field, for x ≥ 0. Try this without
using a graphing calculator or graphing program. Determine the speed and
describe the fluid motion.
Exercise 2.1.1
v(x, y) = (3, 0)
Exercise 2.1.2
v(x, y) = (x, 0)
Exercise 2.1.3
v(x, y) = (3x, x)
Exercise 2.1.4
v(x, y) = (x, y)
Exercise 2.1.5
v(x, y) = (−y, x)
Exercise 2.1.6
v(x, y) = (3x+ y, x + y)
Exercise 2.1.7
v(x, y) = (1, cos x)
Exercise 2.1.8
v(x, y) = (x, cos x)
Exercise 2.1.9
v(x, y) = (cos2 x, sin x)
Exercise 2.1.10
v(x, y) = (sin2 x, cos2 x)
2.2 Water flows, streamlines and integral curves
Test particles follow trajectories called streamlines. The velocity vector of a
test particle is tangent to the streamline. On the other hand, given a velocity
field v and an initial point, a curve r(u) = (x(u), y(u)) is a streamline or
integral curve when at every point along the curve its velocity vector is v.
That is, we require
Exercises
In Exercises 2.2.1–2.2.8, compute the integral curve for the given velocity
field.
Exercise 2.2.1
v(x, y) = (3, 0)
Exercise 2.2.2
v(x, y) = (x, 0)
Exercise 2.2.3
v(x, y) = (3x, x)
Exercise 2.2.4
v(x, y) = (x, y)
Exercise 2.2.5
v(x, y) = (−y, x)
Exercise 2.2.6
v(x, y) = (3x + y, x + y)
Exercise 2.2.7
v(x, y) = (1 , cos x)
Exercise 2.2.8
v(x, y) = (x, cos x)
Keep in mind that our inquiry begins with water and other fluid flows. And
so, we can take advantage of what is known in elementary physics. A key
quantity in physics is momentum, namely, “(mass) × (velocity).” Classical
physics investigates the motion of objects with centers of mass, but we are
investigating water flow. We can adapt the definition of momentum to water
flow.
In the situation just described, what is the rate at which mass flows in the x-
direction across a 1 cm cross-section parallel to the y-axis? In other words,
how many grams of water flow in the x-direction, per second, when the
velocity of the flow is 10 cm per second in the x-direction and the cross-
section (y direction) is 1 cm?
In one second, 10 · 1 square cms of water pass the 1 cm cross-section. In a
water flow, below each square cm of the surface is a volume of water with so
many grams per cubic cm. Remember that, in our notation, we have
suppressed the depth component, but depth is there. This means that we can
introduce a linear mass density term, μ. In other words, μ represents the
number of grams per linear cm. This can be defined and measured because,
while the vector field has only two components, we are referring to volumes
of water in 3 dimensions, and volumes have mass.
In our example, then, the mass-flow per second that passes a 1 cm cross-
section is
Problem 2.4.
If time is some fraction of a second, if the cross-section in the y-direction is
some fraction of a cm, or if the depth is something different than 1 cm (see
hypotheses given in the second paragraph, above), then mass-flow rates
need to be adjusted. Explore different cases. What are the appropriate
formulas?
_________________________
26For the present context, we are not providing a definition of turbulence. That would be done in a
follow-up course in fluid dynamics.
Chapter 3
Note that, numerically, the surface area of a 5 × 1 surface also is 5, but units
are cm2.
If we first look at the 1 × 1 cm2 immediately adjacent to the y-axis in the x-
direction, then in one second, 10 cm flows through that square. But the
segment along the x-axis that we are imagining is 5 cm in length and 1 cm
wide.
Imagine the flow originating well to left (from somewhere in the negative x-
direction) and continuing well to the right (in the positive x-direction).
Figure 3.2 Velocity with x-component = 10 cm/sec and linear mass density μ g/cm. Total mass-flow
10 · μ g in 1 second.
Through each of the five 1 × 1 cm2 along that length, there is a flow rate of
10 centimeters per second. So, with a 1-cm cross section, the total mass-flow
rate through the entire length in the x-direction is
This rate is called the circulation (of mass) along the 5-cm length.
(Note that we are not replacing 5 · 10 by 50. This is intentional, in order to
help us keep track of units and sources of terms. This will be helpful as we go
on to other cases.)
Problem 3.1.
In the example described, what is the total mass in motion, over a 3 second
time period?
The purpose of Problem 3.1 is to emphasize that units for circulation are
rates, a total of “grams per second along a curve.” If an engineer needs to
determine not mass flow rate along a curve, but total mass in motion along a
curve over a period of time, then they would need to also compute (or
approximate) a time integral.
Suppose that a line segment of length L has a unit tangent vector T = (c, d),
(that is c2 + d2 = 1).
Note: The fact that components of velocity vectors are relevant is a physical
result. That is, in applications, velocity and momenta vectors add component-
wise.
And so, making use of the dot product, we can compute the mass flow rate
along the length of L. For then the component of v = (a, b) in the direction of
L (parallel to L) is v · T. Hence, the mass-flow rate along all of L is
Substituting these into the defining integral, we get that the total circulation
of v along the curve C is
Exercises
In Exercises 3.3.1–3.3.13, use the formula in Section 3.3 to compute the total
circulation for each velocity v(x, y) along the curve C. Graph and explain
your results.
Exercise 3.3.1
v(x, y) = (3, 0), C is the line-segment along the x-axis that starts at (0, 0) and
ends at (5, 0).
Exercise 3.3.2
v(x, y) = (3, 0), C is the line-segment along the x-axis that starts at (0, 0) and
ends at (0, 5).
Exercise 3.3.3
v(x, y) = (x, 0), C is the line-segment along the x-axis that starts at (0, 0) and
ends at (5, 0).
Exercise 3.3.4
v(x, y) = (x, ex+y), C is the line-segment along the x-axis that starts at (0, 0)
and ends at (5, 0).
Exercise 3.3.5
v(x, y) = (x, b(x, y)), where b(x, y) is any function for which the integrals exist
(e.g., continuous), and C is the line-segment along the x-axis that starts at (0,
0) and ends at (5, 0).
Exercise 3.3.6
v(x, y) = (3, 0), C is the boundary of the square with vertices A(0, 0), B(1, 0),
C(1, 1), D(0, 1), parameterized counter-clockwise and starts at A(0, 0).
Exercise 3.3.7
v(x, y) = (x, 0), C is the boundary of the square with vertices A(0, 0), B(1, 0),
C(1, 1), D(0, 1), parameterized counter-clockwise and starts at A(0, 0).
Exercise 3.3.8
v(x, y) = (x, y), C is the unit circle centered at the origin (0, 0), initial point
A(1, 0), parameterized counter-clockwise.
Exercise 3.3.9
v(x, y) = (x, x), C is the unit circle centered at the origin (0, 0), initial point
A(1, 0), parameterized clockwise.
Exercise 3.3.10
v(x, y) = (−y, x), C is the unit circle centered at the origin (0, 0), initial point
A(1, 0), parameterized counter-clockwise.
Exercise 3.3.11
v(x, y) = (−y, x), C is the circle of radius r > 0 centered at the origin (0, 0),
initial point A(1, 0), parameterized counter-clockwise.
Exercise 3.3.12
v(x, y) = (3x, x), C is the boundary of the triangle with vertices P(0, 0),
Q(10/3, 0), R(3, 0), initial point P(0, 0), parameterized counter-clockwise.
Exercise 3.3.13
v(x, y) = (3x + y, x + y), C is the boundary of the triangle with vertices P(0,
0), Q(10/3, 0), R(3, 0), initial point P(0, 0), parameterized counterclockwise.
In other words, the chain rule is what allows us to evaluate an integral using a
parameter that happens to be convenient. So long as the criteria of the chain
rule are met, the integral has the same value, namely,
and y(u0) and y(u1) are, by definition, the same for all parameterizations,
because these represent the y-coordinates of the end points of the curve C in
the ambient coordinate plane.
Problem 3.2.
A good exercise is sort this out in the general case, without appealing to a
textbook. It is a good way to shore up one’s understanding of the chain rule.
In other words, show, by computation, that the circulation integral does not
depend on parameterization.
and that (dx, dy) is an increment in the plane that can be written as dr. That is,
r := (x, y) and dr := (dx, dy).
Using this more concise notation, we can also write circulation of v along C
is
and dy becomes
Figure 3.5 Two pipes collecting river water that flows due east.
Let’s suppose that the river flows west to east. Town A is on an island in the
middle of the river and town B is on the north shore, somewhat east, and so
downstream, of the island. A first pipe will connect a location O on the
southern shoreline to town A. That will provide the island-town with water. A
second pipe will connect the island to the northern shoreline. That pipe will
provide water for town B. There will be collection holes in both pipes into
which river water flows. We are ignoring numerous aspects of the
engineering problem. Still, the rate at which water flows into and then along a
collection pipe, while not identical to, is proportional to a circulation integral.
The part of the problem that is new here is that we will need to compute two
integrals. Let v be the vector field for the river velocity, C1 is the curve across
the river that the first pipe will follow from the southern shore to the island;
and C2 is the curve across the river that the second pipe will follow,
connecting the island to the northern shore.
At a moment in time, we consider the sum of all mass-flow rates at all
locations along both pipes. So, the total mass-flow rate of water along the two
pipes will be the sum of two circulation integrals. However, some care is
needed in choosing parameterizations.
Suppose that, for the pipe connecting the southern shore to the island, the
circulation integral is parameterized so that the initial point is on the southern
shore and the final point is on the island.
Then, one contribution to total circulation is
You may remember that they are “the same” but opposite in sign.28
If you trace back through how to compute circulation, you will find the same
result. That is, if we switch end points, the sign is reversed, in the original
circulation. To see that, introduce a parameterization and you will have a
Calculus II problem.
What does that tell us about the engineering problem for the two towns?
If the engineering firm uses the island as the end point of both C1 and C2,
then one of the circulations will be negative in sign to the other. We leave it
as an exercise for the reader to explore why that would cause problems. One
obvious problem is that adding two circulations where the connecting curves
Figure 3.6 Elementary integration. Reversing order of integration changes sign of integral.
Definition 3.2. Suppose that the end point of a curve C1 is the initial point
of curve C2 . Then, the sum of the two curves is denoted by C = C1 + C2.
By definition, the “sum” means: join the two curves, as described above, to
get a new “larger” curve. The circulation along the sum of curves is then
defined to be
Note: This result extends to adding any finite number of curves. It follows
that
so long as the end point of Ci is the initial point of Ci+1, for i = 1, · · · , m–1.
Problem 3.3.
Prove that under the hypotheses, the definition for adding circulation
integrals is independent of parameterizations.
Exercises
Exercise 3.6.1
Let v(x, y) = (x, 0) and C be the boundary of the square with vertices P(0, 0),
Q(1, 0), R(1, 1), S(0, 1), parameterized counter-clockwise and starts at P(0,
0). Using the four sides of the square as curves, write C as a sum C = C1 + C2
+ C3 + C4. Show that
Exercise 3.6.2
Let C1 be the unit circle centered at the origin (0, 0), initial point A(1, 0),
parameterized counter-clockwise, and let C2 be the unit circle centered at the
origin (0, 0), initial point A(1, 0), parameterized clockwise.
Figure 3.7 Vector field v = (2, 0) and rectangle dx × dy centered at (0, 0) with sided I, II, III and IV.
Circulation is a mass-flow rate along a curve. However, let’s explore the idea
of “circulation about a point” which, in this case, is (0, 0).
Sides I and III are parallel to the x-axis, while sides II and IV are parallel to
the y-axis. The net mass flow rate is, then, a combination of circulations
along I and III, respectively. However, what combination should that be?
We need to compute along curves that are appropriately connected. Start with
I parameterized in the positive x-direction. Tracking around the rectangle to
compute circulation along all segments that surround the origin, we get
with the stipulation that these are parameterized so that the end point of I is
the initial point of II, the end point of II is the initial point of III and the end
point of III is the initial point of IV.
Problem 3.4.
Show that this yields the following sum:
Why are there zeros? It is because the vector field is orthogonal to increments
II and IV, respectively. This is consistent with the graphed vector field.
There is no circulation about the origin because the flow is constant in one
direction. This is only descriptive, but it is a clue.
Example
In this example, let v = (y, 0).
Figure 3.8 Vector field v = (y, 0). Rectangle in the first quadrant with vertices A(x, y), B(x + dx, y), C(x
+ dx, y + dy) and D(x, y + dy).
Draw a rectangle in the first quadrant of the (x, y) plane with the following
vertices: A(x, y), B(x + dx, y), C(x + dx, y + dy) and D(x, y + dy).
Appropriately parameterized, AB is I, BC is II, DC is III and DA is IV.
Let’s now compute circulation along the boundary of the rectangle, using the
curve I + II + III + IV, as in the diagram.
Figure 3.9 Boundary of rectangle parameterized: I + II + III + IV, where I = AB, II = BC, III = CD
and IV = DA.
Figure 3.10 Small time increment dt > 0, A flows to A′, and D flows to D′. The vector field is v = (y, 0)
implies |DD′| > |AA′|.
Example
For this example, v = (−y, x) is the familiar vector field whose streamlines
orbit the origin (0, 0).
As in previous examples, we can introduce a small rectangle which, in this
example, will be centered about (0, 0).
Figure 3.11 Vector field v = (−y, x). Small rectangle dx × dy centered at (0, 0).
The next step is elementary and is the key to discovering Green’s Theorem. It
leads to the main ingredient for the theorem. See Section 3.8.
Grouping terms, there are differences in function values. And linear
approximation tells us how to approximate those differences.
In two variables, linear approximation involves two partial derivatives and a
reference point. In the present application, the reference point is (x, y).
Figure 3.13 v(x, y) = (a(x, y), b(x, y)), evaluated at vertices A(x, y), B(x+dx, y), C(x + dx, y + dy) and
D(x, y + dy).
But in the two differences
only one of the terms involves the reference point (x, y). A common trick
(which survives into higher analysis) is to compare by getting from one
location to the next in steps that, strategically, “go through” the reference
point (x, y). For instance, one can do this by following the edges of the
rectangle.
In the difference
the “final point” is (x + dx, y) while the “initial point” is (x, y + dy). To
compare these points, first compare the function values at (x + dx, y) and (x,
y); and then compare the function values at (x, y) and (x, y + dy).
This means that we compute
This gives
Remember that this is an approximation only for the first term in brackets in
the expression
We will need to sort out the second bracket too. But for the moment, let’s
continue with our focus on the first square bracket, and substitute what we
just computed.
The first square bracket becomes
In the present context, then, circulation is found to be an integral relative to
an area of a small rectangle, relative to which the term
Problem 3.5.
Work out the details and show that, in our linear approximation, the other
square bracket gives a similar result, namely:
If we combine the two sets of computations, what is the result? (See (2)
below.)
Figure 3.14 Closed curve, P = r(u0) = r(u1), with positive orientation, that is, parameterization tracks
boundary curve C = ∂R counter-clockwise, with dθ > 0.
In other words,
The term
where each (xij, yij) is, say, the bottom left corner of each rectangle in the
sum.
Where does circulation come into play?
Remember that in their original form (before linear approximation), each of
those summands is an approximation to
Problem 3.6.
As the diagram suggests, contributions to circulation along common edges
internal to the four rectangles all cancel. That is because they are each
computed twice, with opposite signs.
Hence . . . . !
Do you see it?
The sum of circulations around the four rectangles reduces to the
circulation along the outer edge of the large rectangle that is the union of
the four small rectangles.
Hence . . . . !!
And, we can keep going!!!
That is, if we add up circulations around all rectangles in the partition within
the region R, common internal terms cancel. And this leaves circulation along
the closed curve that is determined by the union of the outer edges of
rectangles of the partition that lie within R.
In the limit, these outer edges converge to the boundary curve of the region
and the sums therefore converge to the contour integral along the closed
curve,
Figure 3.16 (See Figure 3.15.) Interior circulations cancel. Remaining terms are along boundary of
grid that is interior to the region R.
On the other hand, linear approximation of each summand means that we also
end up with sums that converge to an area integral, namely,
The two computations are two different ways of computing the same
quantity.
And so we get the following:
Exercises
For R the unit disk centered at (0, 0), compute the circulation along ∂R,
counter-clockwise. Does this contradict the previous exercise? Graph and
think about streamlines.
We suppose a small time interval dt. Let’s see what happens to A = (0, 1) and
B = (0, 2) along their respective integral curves, reaching A′ and B′,
respectively; and then compare AB with A′B′.
At A = (0, 1), the velocity vector is v = (1, 0) and at B = (0, 2), the velocity
vector is v = (2, 0).
So for small time interval dt we get
But recall that AB = (0, 1). So, for instance, for small dt > 0, the
corresponding angle differential between AB and A′B′ is dθ < 0, clockwise, or
negative.
Remember that circulation density of v is –1 < 0.
Example
Let v = (–y, x) be the vector field. At time t = 0, start with reference points A
= (1, 1) and B = (2, 2). In other words, A and B are along the same radius
emanating from the origin (0, 0).
The formula for the vector field is linear and so linear approximation methods
are redundant.
We start with the line segment AB = (1, 1) and we suppose a small time
interval dt.
Figure 3.18 Vector field v = (–y, x). Two initial points are A = (1, 1) and B = (2, 2).
At A = (1, 1), the velocity vector is v = (–1, 1) and at B = (2, 2), the velocity
vector is v = (–2, 2).
So for small time interval dt we get
But recall that AB = (1, 1). For small dt > 0, the corresponding angle between
AB and A′B′, is dθ counterclockwise, or positive.
As is evident from the geometry,
For small angle differentials, sine of an angle differential is close to the angle
differential (all angles, of course, in radians). Hence, for the small time
interval, we get that
We get circulation density of v = (–y, x) is 1 – (–1) = 2 > 0.
Figure 3.19 Vector field v = (–y, x). Increment AB flows to increment A′B′. Relative change in angle is
dθ > 0.
Problem 3.7.
From the geometry sin .
But for small angle differentials, the sine is close to the angle (all angles in
radians). And so, for small time interval, we get that dθ ≈ dt > 0.
and
Next, we compute and compare the two line segments, AB and A′B′. We have
AB = (dx, dy) and from the above computations,
Remember that the objective is to compare AB and A′B′.
The line segment AB = (dx, dy) is a small increment connecting two nearby
points A and B. Based on our computations so far, we have that AB = (dx, dy)
flows to a new line segment A′B′.
It can be helpful to display this in matrix notation:
Note: In this book, all matrices are real. Also, henceforth, the superscript
‘transpose’ is abbreviated to ‘t’.
Problem 3.8.
The decomposition is unique in the sense that if
is a sum of a symmetric and an anti-symmetric matrix, then
Problem 3.9.
In the case of 2 × 2 matrices, prove the diagonalization result for symmetric
matrices. (In higher dimensions, we need the real spectral theorem. That
would require more background than is being supposed for this book.)
Figure 3.22 Using, principle coordinates of (Jv)S, in linear approximation, rectangles flow to
rectangles.
This means that we now have a precisely identified geometric meaning for
the quantity bx – ay.
That is,
One way to see that is to observe that
where
Hence, the term bx − ay appropriately is called the (scalar32) “curl” of the 2-d
vector field v(x, y) = (a(x, y), b(x, y)). And so we define (scalar)
We have solved two problems, but we are getting the same answer in
both cases
(1) The quantity (bx – ba) is circulation density per unit area.
(Recall that “circulation” is defined to be a line integral of (components
of) mass-flow rates along a curve. In the present context, circulation
density per unit area is a limit of line integrals, each of which is along a
closed curve with (x, y) in the region interior to the curve.)
(2) Near the same reference point (x, y) as in 1., the term (bx – ay) is twice
the angular velocity of A′B′ relative to AB. Because of this, it is
appropriate to call (bx – ay) the (scalar) “curl” of the velocity field v(x, y)
= (a(x, y), b(x, y)).
Question: Why are the two answers the same? Sorting that out will require
digging into subtleties. We will take this up at the end of Part 1.
Let’s return to some of our familiar velocity vector fields, to see our results in
action.
Example
Let v = (–y, x). Then curl v = curl(–y, x) = bx – ay = 2.
Calculating the matrix quantities, we get
Here,
is a pure rotation with unit angular velocity, while the angular velocity
obtained from the anti-symmetric
is 2.
Example
Let v = (x, y). This is a vector field where the origin is a source and
streamlines are directed along radial arms emanating from the origin.
Looking at the graph of the vector field, one might correctly anticipate that
relative rotation is minimal or perhaps zero.
Then curl v = curl (x, y) = bx – ay = 1 – 1 = 0. Calculating the matrix
quantities, we get
Exercises
Exercise 3.10.1
Let v(x, y) = (1, 1).
(a) Graph the velocity field and its streamlines.
(b) Find explicit formulas for the streamlines, in the form r(u) = (x(u), y(u)),
where u is a real variable.
(c) Compute curl v(x, y). Explain.
Exercise 3.10.2
Let v(x, y) = (x, x).
_________________________
27In a more advanced context (“higher viewpoint” — see Part A of the Supplement), terms of the form
w = adx + bdy are called “1-forms” and C is a “one-dimensional manifold.” Higher dimensional objects
like surfaces, volumes, and so on are “two-dimensional manifolds,” “three-dimensional manifolds,” and
so on. The “k-forms” formalize the notion of “line element,” “area element” and “volume element,”
respectively. A one-form w = adx + bdy provides “line elements” along a curve C in the sense that, at
each point p on C, it defines a linear functional on the space of tangent vectors at p. Computation
reduces to what we have described above. That is, one needs to introduce a parameterization and
integrate in the usual way.
28Hint: An original notion of integral is “area under a curve, over an interval.” When intervals change,
how do areas add or subtract?
29One way to sort out what the relationship is to invoke Green’s Theorem. But that would be making
use of a result that in this book has not yet been established.
30Steps need proof including rigorous limit arguments. See Preface.
31It is also the matrix of components of acceleration of the velocity flow. Further aspects of the
Jacobian would be studied in a course in fluid dynamics.
32A “vector curl” in 3-d is obtained in Part 2 of this book.
Chapter 4
As in previous sections, the small rectangle dx × dy has four sides, I, II, III
and IV.
We estimate “mass-flow rate out minus mass-flow rate in.”
To do that, we use the components of v = (a, b) that are orthogonal to each
side, and subtract appropriately.
That is, for the small rectangle, the net mass-flow rate is
This is similar to the computation for mass-flow rate along a curve which, as
you may recall, was a line integral.
Question: But is this last equation a line integral?
Figure 4.1 Net mass-flow rate through region dx × dy is net mass-flow rate in y-direction minus net
mass-flow rate in x-direction. Use normal components to sides I, II, III and IV, respectively. That is,
mass-flow rate is [b(x + dx, y + dy)dx – b(x, y)dx] + [a(x + dx, y)dx – a(x, y + dy)dx].
Problem 4.1.
Provide a similar identification for the other two terms.
Observation: In each case we get the right terms, so long as we compute the
projection (or dot product) onto what, relative to the positively oriented
boundary is the “outward directed unit normal vector N.”
Therefore, the net mass-flow rate across the region dx × dy can indeed be
written as a line integral in the following way:
The rectangle has sides I, II, III and IV. Link them up to obtain a positively
oriented boundary curve for the region
A technical issue regarding the integral for net mass-flow rate across a
small rectangular region
Can the integral that we just defined be expressed as a line integral of the
form
where f and g are functions defined on the plane?
To compute
In other words, the net mass-flow rate is a line integral of the form
Problem 4.2.
The 2-d flux integral is invariant under change of parameterization. Provide
details.
Consequently, near a point (x, y), the net mass-flow rate through the small
rectangular region dx × dy is (ax + by)(dxdy).
Dividing by the area term, we get that the net mass-flow rate per unit area is
ax + by.
Just as in Green’s theorem, we obtain a density function which in this case is:
The quantity ax + by is also called the divergence of the vector field and is
often written as div v.
Just as in the development of Green’s theorem, by linear approximation we
get two ways of approximating the same quantity. And the resulting theorem
can be left as an exercise.
Exercises
In Exercises 4.1.1–4.1.7, for each vector field v(x, y) = (a(x, y), b(x, y)),
where possible, compute the flux across the boundary ∂R of the region R,
oriented counter-clockwise: (i) directly by integrating along ∂R; and (ii) by
using the 2-d divergence theorem (Problem 4.4). Graph the vector field and
its streamlines and, if possible, find explicit formulas for the streamlines by
solving an appropriate system of differential equations. Graph and explain
your results.
Exercise 4.1.1
v(x, y) = (1, 0), R is the square with vertices A(0, 0), B(1, 0), C(1, 1), D(0, 1).
Exercise 4.1.2
v(x, y) = (1, 0), R is the triangle with vertices A(0, 0), B(1, 1), C(0, 1).
Exercise 4.1.3
v(x, y) = (x, y), R is the unit circle centered at the origin.
Exercise 4.1.4
v(x, y) = (1, 1), R is the unit circle centered at the origin.
Exercise 4.1.5
v(x, y) = (x, x), R is the square with vertices A(0, 0), B(1, 0), C(1, 1), D(0, 1).
Exercise 4.1.6
v(x, y) = (x, x), R is the square with vertices A(1, 0), B(2, 1), C(1, 2), D(0, 1).
Exercise 4.1.7
v(x, y) = (–y, x), R is the unit circle centered at the origin.
Exercise 4.1.8
Let and R be the unit disk centered at (0, 0). Show
that both sides of the divergence theorem are zero. But why does this not
provide an illustration of the divergence theorem? Hint: Consider the
domains.
Exercise 4.1.9
This exercise relates to Exercises 4.1.3–4.1.4:
Problem 4.5. (The 2-d divergence theorem from Green’s theorem; and
vice versa)
We have two integral formulas:
Green’s theorem:
and the
2-d divergence theorem: .
The formulas are obtained by answering different questions about a vector
field v = (a, b). The questions and answers are different. At the same time,
there are evident similarities. Are the two formulas related? Or, in symbols,
Clue: Under our hypotheses, the vector field v = (a, b) was arbitrary.
To follow up on that clue, let’s use different symbols for each formula.
Rewriting, we get
Green’s theorem:
and
2-d divergence theorem:
Example
For v = (2, 0), div v = div(2, 0) = 0 + 0 = 0.
Example
For v = (y, 0), there is a source term in the first component that increases with
y.
The divergence is div (y, 0) = 1 + 0 = 1. In other words, as y increases, so
does the divergence.
Example
Let v = (–y, x) be the familiar vector field whose streamlines orbit the origin
(0, 0). As you might expect, the divergence is zero. Explicitly, div(–y, x) = 0
+ 0.
Example
Imagine a round fountain with a conical surface, and a steady source of water
that flows from its top. Water flows down the conical surface in all
directions. It has (approximately) constant acceleration due to gravity until at
the bottom of the cone’s edge, water falls into a surrounding channel that
collects water for recirculation through the fountain.
Figure 4.3 Water fountain and aerial view of water-flow velocity field.
Problem 4.7. Explain why the depth of water decreases as it spreads out
along the surface of the cone. Hints: Polar coordinates, increments in
radius, and water volume. From physics, assume that water is not
destroyed.
With an aerial view (that is, focusing on x and y coordinates and ignoring
height of cone and total depth of water at locations along the surface of the
cone), streamlines flow radially, away from the origin; and the magnitude of
the vectors (speed) increase with distance from the origin.
A simplistic mathematical model that has similar features is v(x, y) = (x, y).
For this vector field, divergence is div (x, y) = 1 + 1 = 2.
The flux density per unit area is 2.
Let’s review what this means in this case.
Figure 4.4 Approximating difference in flux using difference area.
Warmup problems
Warmup 1:
Let v(x, y) = (x2, 0). In this example, there is acceleration in the first
component of the velocity field, relative to change in the first coordinate. Let
A = (1, 0), B (2, 0) and B2(1, 1). In this case, initial [Area] = 1.
Motion is in the direction of the x-axis. Suppose that the initial value of the
parameter for the integral curves is t = 0 and that time increases.
At t = 0, v(1, 0) = (1, 0), v(2, 0) = (4, 0), and v(1, 1) = (1, 0).
For a small time increment dt > 0, A = (1, 0) and B2 = (1, 1) both advance in
the x direction by 1.
Therefore,
Figure 4.8 A flows to A′, B flows to B′. Compare lengths of L = ||AB|| and L′ = L||A′B′||.
and
Because
these are
and
This gives
At B1 = (x + dx, y),
At B2 = (x, y + dy),
Therefore,
and
Also,
Problem 4.8.
Using the absolute value of the determinant, compute the area determined
by and . Show that
This reduces to
For a discussion about why the determinant function provides area, see
Section 1.7.
That is,
Note: This is not a rate of change of area. The solution is a rate of change of
relative change in area, determined by the flow of vertices along streamlines
(integral curves) that are simultaneously parameterized.34
is called incompressible.
Start with a small element of area and examine its progression for small time,
relative to the original small element of area. Consider the relative change of
area, with respect to time. In the limit, we obtain ax + by.
Recall that ax + by is called the divergence of the vector field, where that
name originates from estimating net mass-flow rates (flux) through a small
region whose boundary is a positively oriented curve.
Just as in our inquiry about circulation, two different questions have the
same answer:
At a point (x, y):
Question: Why are these answers the same? In classical fluid dynamics this
is equivalent to a conservation law for a type of fluid. In Chapter 5, we go
into some of the mathematical details.
Flux:
(1) At a fixed time t, using a rectangle dx × dy, the flux density function per
unit area is div v = div (a, b) = ax + by. (Linear approximation)
(2) By summing both sides of (1), and taking limits as partitions are refined,
we get the 2-d divergence theorem, that is:
Or, in words, with modest hypotheses, the flux of v = (a, b) along the
closed boundary of R is equal to the area integral of div v, over the
region R.
(3) Following areas as vertices flow along streamlines, the rate of relative
change in area is div v = div (a, b) = ax + by (using common integral
curve parameter).
(4) In terms of the Jacobian Jv, the divergence
_________________________
33In fluid dynamics, this is called “strain” in 2-d motion.
34See note 8.
Chapter 5
(1) Green’s theorem and the 2-d divergence theorem tell us different things
about a vector field. Mathematically, however, the two theorems are
equivalent.
(2) In Green’s theorem, circulation density (bx – ay) is also equal to
That is,
Next, let’s have a look at equalities observed between density functions and
rates of change obtained from geometric constructions.
In the 2-d divergence theorem, flux density ax + by is equal to the rate of
relative change in areas.
The two quantities are computed and defined differently.
The flux density is a limit of ratios of line integrals to interior areas, with no
explicit reference to streamlines.
But the rate (of change) of relative change in areas is obtained by tracking
vertices of rectangles along respective streamlines, with a common
parameter.
Let’s explore a few examples.
Example
Suppose that the coordinate system is already aligned with the vector field. In
other words, with (x, y) coordinate lines, we suppose that the vector field is of
the form v(x, y) = (a(x, y), 0).
Let’s compute both quantities, that is, both the ratio of flux to area and the
ratio of relative change in areas.
We can start with a rectangle whose vertices are listed in counter-clockwise
order:
Problem 5.1.
Show that, as before, by linear approximation — ignoring high order error
terms — the sum reduces to ax(x, y)dxdy.
Dividing by the area dxdy we get that flux density is ax(x, y) (which is ax + by
because in these coordinates, b ≡ 0).
In the same example, let’s look at the rate of change of the relative change in
areas.
As in the general derivation of the result given earlier, we need to compute an
approximation to
Problem 5.2.
Show that the rate of change of relative area reduces to ax(x, y).
Notice that the coordinate axes are lined up with streamlines. Since motion is
parallel to the x-axis, the only contribution to flux density is ax(x, y); and the
rate of change of relative change of areas is ax(x, y).
Problem 5.3.
Investigate the geometry in the somewhat more general case where
Problem 5.4.
Investigate the case where
with
That is, suppose that the Jacobian represents a pure rotation. Using only
elementary geometry, show that in the case of pure rotation, areas are
preserved. That is, show that
and
We already have that M2 = rgS ⊕ rgA. (See Section 3.10.)
Explicitly,
and
Using the abbreviation ‘ker’ to stand for the kernel (or null space) of a linear
map, observe that for
we have
and
What does the analysis of linear structure mean for a fluid with a 2-d velocity
field v?
If div v = 0 then, as defined in Section 4.3, the fluid is incompressible.
From the linear structure described above, an incompressible fluid can have a
velocity field that is partly rotational.
If div v = 0, then eigenvalues of the symmetric part (Jv)S are either both zero,
or are equal but of opposite sign.
If they are both zero, then there is neither contraction nor dilation relative to
the principle coordinates.
If they are non-zero but equal and of opposite sign, then there is contraction
relative to one of the principle axes matched by a dilation relative to the other
principle axis. See Figure 3.22.
Problem 5.5.
By going back to first principles, investigate 2-d (scalar) curl and geometry
in the case where the velocity field is of the form v = (a(x, y),0).
Let’s do it together.
We are inquiring about curl. There is, however, a correspondence between
curl and divergence, namely, curl v = div v⊥. Where does that lead?
Given the vector field v = (a, b), then v⊥ = (b, –a) also is a vector field with
its own Jacobian, that is,
The Jacobian of v⊥ has its own symmetric and anti-symmetric parts:
and
It follows that
The computation also reveals that the 2-d scalar curl is unaffected by 2-d
divergence along streamlines. Again, analysis of linear structures gives
additional results.
Observe that
is obtained from
is an isomorphism of M2:
hence T is one-one.
(ii) Given C ∈ M2, then T((QP)−1C) = C, hence T is onto.
Problem 5.6.
Recall that if curl v = 0, then the vector field is irrotational. Because div=
trace(Jv⊥) = trace(T(v)), just as was done in Section 5.2, analysis of the
kernel of appropriate linear maps provides further results.
Exercises
Exercise 5.3.1
Let f(x, y) = x2 – y. Find the level curves f(x, y) = c, where c is a constant. Let
v(x, y) = ∇f = (2x, –1) and v⊥(x, y) = (–1, –2x). Show that the level curves of
f are the streamlines of v⊥(x, y) and that the orthogonal trajectories to the
level curves are the streamlines of v(x, y) = ∇f. Graph both sets of curves and
explain.
Exercise 5.3.2
Let r(u) be a flow line of a gradient vector field v = ∇f. Show that g(u) =
f(r(u)) is an increasing function of u. Thus a particle moving along a flow line
of the gradient field v = ∇f will move from lower to higher values of the
potential function f.
_________________________
35For example trace
PART 2
FLUID MOTION IN 3 DIMENSIONS
Chapter 6
6.1 Beginnings
As in the 2-d case, we start with a velocity vector field. Now, however, we
look to the three-dimensional case. The velocity field is of the form v = (a(x,
y, z), b(x, y, z), c(x, y, z)) in an (x, y, z) coordinate system. We continue to
assume that velocity fields are independent of time. We also continue to
assume that all functions are continuously differentiable everywhere. For
brevity, we write v = (a, b, c), with the understanding that the coordinate
functions are of the form a = a(x, y, z), b = b(x, y, z) and c = c(x, y, z).
Problem 6.1.
What are some examples?
Figure 6.1 Mass per unit area is used to compute mass of volumes.
As in the 2-d case, the velocity field can be integrated to give integral curves,
trajectories of “test particles.” Also as in the 2-d case, initially, density
functions are empirically verifiable. For instance, for a layer of known depth
there is a mass-density per unit area. It is called “per unit area” with the
understanding that masses are of volumes.
For a tube of known diameter, there can be a length density function. Again,
mass is of increments of volume, in this case along a tube.
Figure 6.2 Volume has mass. Mass per unit length is used to compute mass of volumes.
With these subtleties understood, we now ask about mass-flow rates across
surfaces.
Suppose that an area density function μ is defined on surface S in (x, y, z)
space.
Mass-density is a property of a fluid. As in Part 1, we restrict to the case
where that function is constant. If that constant is not unity then, by an
appropriate adjustment of units, it becomes unity. Henceforth, we assume that
μ = 1. Under these hypotheses, this means that, numerically, momentum μ =
μ(a, b, c) = (a, b, c). Although, units of momentum are mass × velocity.
Problem 6.2.
Work out details for the change of scale needed if, for instance, mass-
density per unit area is gm per square cm.
Figure 6.3 Velocity field passes through surface area.
Now that we have “increments of flux” defined, much as in the 2-d case, this
allows for the following:
Definition 6.1. For v = (a, b, c) and suppose that n is a unit normal vector
normal field defined along the surface S. Relative to a unit normal vector
field n, the flux (mass-flow rate) across the surface S is
The definition includes the word “relative to n”. As you may already have
observed, at a point on a surface in (x, y, z) space, there are two possible unit
vectors. We will need to say more about that, shortly.
As in the 2-d case, this is an average, or integral. In the 3-d case, it is an
integral of mass-flow rates across the surface S in (x, y, z) space.
Figure 6.5 Increment of flux is increment of mass-flow orthogonal to surface, that is, d(flux) = (v ·
n)dS.
Exercises
In Exercises 6.2.1–6.2.8, find the flux for the given vector field v across the
surface S.
Exercise 6.2.1
v = (z, 5, 2y), S is the part of the plane x + y + z = 1 in the first octant with the
upward orientation
Exercise 6.2.2
v = (xy, yz, zx), S is the part of the paraboloid z = 2 – x2 – y2 that lies above
the square 0 ≤ x ≤ 1, 0 ≤ y ≤ 1, and has upward orientation
Exercise 6.2.3
v = (y, xz, x), S is the hemisphere oriented in the direction of
the positive y-axis
Exercise 6.2.4
v = (x, y, z4), S is the part of the cone under the plane z = 1 with
the upward orientation
Exercise 6.2.5
v = (x, y, 2z), S is the part of the plane z = 3x + 2 that lies within the cylinder
x2 + y2 = 4 with the upward orientation
Exercise 6.2.6
v = (xz, x, y), S is the part of the part of the sphere x2 + y2 + z2 = 81 in the first
octant with the upward orientation
Exercise 6.2.7
v = (3x, y, 5), S is the part of the paraboloid z = 9 – x2 – y2 that lies above the
xy-plane with the upward orientation
Exercise 6.2.8
v = (xz, x, y), S is the hemisphere oriented in the direction of
the positive y-axis
Example
Let S2 = {(x, y, z)|x2 + y2 + z2 = 1}, the surface of the sphere of radius 1
centered at the origin (0, 0, 0). Let v = (a, b, c) be a velocity field.
The problem is to compute
Figure 6.6 A quadrant of the surface of the unit sphere. Two parameterizations partly overlap.
Problem 6.3.
Just as in the 2-d case, invariance reduces to the fact that there is a chain
rule for a change variables in a surface integral. The proof is “in the books”
but it is an excellent exercise to sort through the details for oneself, namely,
that
Let S(x0, y0, z0) be the six-sided surface of a 3-d “box” with sides parallel to the
coordinate planes, with lengths dx, dy and dz, and corner point (x0, y0, z0) (see
Figure 6.7).
Much as in the 2-d case, in 3-d, one way to use approximate “net flux” near a
point is by using linear approximation and summing differences in flux
across opposite (parallel) sides. Using that approach, net mass-flow through
the 3-d box dx × dy × dz is approximated by
Problem 6.4.
Retaining only first-order terms, the “net flux” indicated is
Note: We wrote “net flux” instead of ‘flux’ because we were appealing to the
diagram and did not use normal unit vector field n as is required by the
definition of flux.
This raises a question: Can the sum of terms just obtained be expressed in
terms of a normal vector field n defined on S(x0, y0, z0)?
Problem 6.5.
And so we get that if E(x0, y0, z0) = dx × dy × dz, the coordinate box based at
(x0, y0, z0), then by using linear approximation, the flux over the surface of
the box is the divergence of the vector field times dxdydz, the volume of the
box. That is,
Problem 6.6.
(1) Show that if we add equation 6.1 for two adjoining boxes in a partition,
common surface integrals cancel.
(2) Repeat this result for four adjoining boxes.
(3) Extend this result to the entire triply indexed sum of boxes Eijk based at
(xi, yj, zk) whose six-sided surfaces are labelled Sijk.
The results of Problem 6.6 lead to the following two ways of evaluating the
same limit:
By linear approximation, for each partition of the volume, finite sums can be
written as
with
Theorem 6.1. Let v = (a, b, c). Suppose that E is a simple solid region in
(x, y, z) and S is its boundary surface. This is written S = ∂E. Then the flux
integral of v over the boundary S = ∂E is equal to the volume integral of
the divergence of v over the volume E.
That is,
Problem 6.7.
Extend the method of the previous example to establish this result. In a
more advanced course, rigorous limit arguments would be needed.
In other words, the net mass-flow rate of w = (e, f) along C is a line integral
of the vector field, without having to appeal to an additional vector defined in
the ambient 2-d space.
Question: What about the 3-d case?
We have defined 3-d flux as v · n dS.
(1) Can we write this in a way that does not require construction of a
positively oriented normal vector n?
(2) Is the integral invariant under (some) changes of parameterization?
In the 3-d case there are additional terms. But, as it turns out, we can mimic
the 2-d argument.
As in the 2-d case, we begin by computing terms for the integrand. A local
parameterization is needed, of form
Figure 6.10 A parameterization of a 2-d surface in (x, y, z) space: r : (u, v) → (x(u, v), y(u, v), z(u, v)).
has area
Figure 6.11 Basis vectors Tu, Tv for tangent plane of S at r(u, v).
(where dy and dz are the coordinate intervals for sides of the parallelogram in
(x, y, z) coordinates).
The other components are computed similarly.
That is, the second component of Tu × Tv is –dxdz and the third component is
dxdy.
Therefore,
This implies that the flux is
That is, as hoped, we obtain an expression for flux that does not explicitly
involve a normal vector n.
Figure 6.12 By definition, projections of Tu and Tv onto (y, z)-plane are (0, yu, zu) and (0, yv, zv).
Area of projected parallelogram is therefore |yuzv – yvzu|; which is the absolute value of the first
component of Tu × Tv. Similar computations hold for the other two projections.
Problem 6.8.
Go through the details to show that the flux is invariant under change of
parameterization.
Observation: The formula
depends on the surface S and the vector field v = (a, b, c) defined along S.
The flux integral is, therefore, a function of “two variables,” where the
“variables” are surfaces and vector fields, respectively. This hints of a higher
viewpoint needed to handle integrals on surfaces. For a two-parameter
surface, “integrands” are of the form
These integrands are called “2-forms” and the surfaces are called
“manifolds.” The definition of manifold embraces the fact that
parameterizations are not unique and that, as observed above, some surfaces
(e.g., S2) require more than one parameterization to cover an entire surface.
Exercises
Exercise 6.3.1
v = (3xy2, xez, z3), S is the surface of the solid bounded by y2 + z2 = 1, x = –3,
and x = 1.
Exercise 6.3.2
v = (x2y, xy2, 2xyz), S is the surface of the solid bounded by x = 0, y = 0, z =
0, and x + 4y + z = 4.
Exercise 6.3.3
v = (–x4, x3z2, –4xy2z), S is the surface of the solid bounded by x2 + y2 = 1, z
= x + 10, and z = 0.
Exercise 6.3.4
v = (4x3, 4y3z, 3z4), S is the unit sphere.
_________________________
36See, e.g., James Stewart, Calculus: Early Transcendentals, 6th ed., (Belmont, CA: Thomson Higher
Education, 2008): 1067.
37See note 36.
Chapter 7
STOKES’ THEOREM
Figure 7.2 Given v(x, y, z) in (x, y, z)-space, w = ProjSv is a vector field along a 2-d surface S, with
boundary curve C = ∂S.
(1) One approach to the problem follows what we did in the development of
Green’s Theorem. That is, begin with the definition of circulation.
As in the 2-d case, linear approximation can be used to replace a combination
of line integrals, with surface integrals. Note that since there are three
components and two parameters, the computations become somewhat
involved but they do not require new ideas as such.
(2) Another approach is to take advantage of (*). Then apply Green’s
Theorem to quantities in the (u, v)-plane.
With some guidance provided, both approaches are good homework
exercises and are illuminating in different ways.
By making use of projections, the first approach calls forth techniques that
eventually are needed in modern differential geometry and applications. The
second approach implicitly incorporates the projection in (1) and reveals
coherence of operations in vector calculus.
Below, we provide some details on both sets of computations.
and
We can use the cross-product to get a vector field normal to the tangent
plane. That is, set
(We use the capital letter ‘N’ rather than lower case ‘n’ because the nor
vector just computed is not, in general, a unit vector as described in Sec 6.2.)
Explicitly,
The components for N = Tu × Tv will be used below. For now, let’s keep our
focus on the geometry of vectors.
Using the dot product, ProjN(v), the projection of v onto N, can be computed
directly.
The vector field that we need (ProjS(v), the projection of v onto the surface S)
then is defined by the equation
In other words,
As in Part 1, there are four sides, I, II, III and IV, each of which has its own
initial and final points in (x, y, z) space. These are obtained by tracking
counter-clockwise around the boundary of coordinate interval du × dv and
consequently producing the boundary of the parallelogram in (x, y, z) space.
On sides I, II, III and IV, respectively, the contributions to circulation are:
side I:
side II:
side III:
side IV:
Problem 7.1.
Show that, by ignoring high order error terms, circulation around the
positively oriented boundary of the coordinate interval du × dv mapped to
(x, y, z) space is
The vector
For the first component, suppress the a and ∂x column and use the next two
columns to compute curl (y, z)(b, c).
For the second component, suppress the b and ∂y column and use the next
two columns to compute curl (z, x)(c, a).
For the third component, suppress the c and ∂z column and use the next two
columns to compute curl (x, y)(a, b).
The subscript notation for the pairs of variables is, we hope, clear.
This way of computing the 3-d curl vector emphasizes the fact that the
components of the 3-d vector curl are 2-d scalar curls.
More precisely, the components of curl (a, b, c) are the scalar curls of the
projections of v(a, b, c) onto the (y, z) plane, the (z, x) plane and the (x, y)
planes, respectively.
Note also that this ordering and naming of coordinate planes matches the
derivation above. To see that, write (x y z x y z) and move from left to right in
cyclic ordering. That is, first suppress x to get (y, z); then suppress y to get (z,
x); and finally suppress z to get (x, y).
These aspects of curl (a, b, c) = ((cy – bz), (az – cx), (bx –ay)) will be further
explored in sections below.
2. Another common way to compute curl (a, b, c) is to use symbolism
derived from determinants. That is,
where i = (1, 0, 0), j = (0, 1, 0), k = (0, 0, 1) and the determinant formula is
applied symbolically. Observe that each row has different types of terms.
Problem 7.2 (Stokes’ Theorem).
Fill in needed details. Introduce a partition of the (u, v) parameter space
with coordinate intervals du × dv. Observe that circulation around the
boundary of S can now be computed in two ways (two different limits) and
that we therefore get Stokes’ theorem, namely, with all terms defined as
above,
Note: Just as in the 2-d case, the key is to use linear approximation to replace
differences of functions by their derivatives multiplied by appropriate
coordinate interval lengths.
maps R2 to S2. The regions S1 and S2 and their boundaries ∂S1 = C1 and dS2
= C2 may overlap. Provide details on why the invariance of the integrals
makes it possible to apply Stokes’ theorem to S1 ∪ S2 with boundary ∂(S1
∪ S2).
Figure 7.4 See Problem 7.3.
where, as above,
Here, ∂R is a curve in the (u, v) plane that gets mapped to C in (x, y, z) space.
Green’s theorem applies directly to the (u, v) vector field (v · Tu, v · Tv).
That is,
Problem 7.4.
We know what we’re hunting for. So, all that we need to do is verify that
the integrand
and
are the 2-d scalar curls of the projections of v = (a, b, c) onto the (y, z) plane,
the (z, x) plane and the (x, y) plane.
The components of
are the signed areas of the projection of Tu × Tv onto the (y, z) plane, the (z,
x) plane and the (x, y) plane, respectively.
We have, then, two vectors, three projections, and four areas. What are the
four areas? They are the area A of the parallelogram determined by Tu and Tv
together with A(y,z), A(z,x) and A(x,z), the areas of the three projections of the
parallelogram of area A onto the (y, z) plane, the (z, x) plane and the (x, y)
plane, respectively. See Figure 6.12.
Is there a relation between the areas A, A(y,z), A(z,x) and A(x,z)?
If we bisect the surface area of the parallelogram determined by Tu and Tv
along its diagonal, we obtain a triangle. That triangle is of area A. It can be
projected onto the (y, z) plane, the (z, x) plane and the (x, y) plane, providing
triangles of areas A(y,z), A(z,x) and A(x,z), respectively. An elementary
computation shows that
This implies that the integrand in Stokes’ theorem can be written as follows:
where
Comment: This shows explicitly that circulation density on a 2-d surface in
3-d depends on how the 3-d vector curl v is pitched relative to the normal
vector of the surface. This is not surprising, however, since Stokes’ theorem
is about the projection of v onto S. In particular, if v is parallel to N = Tu ×
Tv, then the projection is zero, as are the integrals.
The calculation also shows that the standard Euclidean structure of the 3-d
vector dS is also a Euclidean structure of triples of projections of 2-d areas.
Figure 7.6 The 3-d vector curl v relative to the normal vector dS = Tu × Tv.
Exercises
Exercise 7.2.1
Let S1 be the portion of the paraboloid z = 4 – x2 – y2 above the xy-plane and
S2 be the hemisphere x2 + y2 + z2 = 4, z > 0. Suppose v(x, y, z) is a vector
field whose components have continuous partial derivatives. Explain why
In Exercises 7.2.2–7.2.4, verify Stokes’ Theorem for the given vector field v
and surface S.
Exercise 7.2.2
v = (y, x, xz), S is the part of the plane x + y + z = 1 that lies in the first octant,
oriented upward
Exercise 7.2.3
v = (y – z, x – z, y – x), S is the hemisphere , oriented upward
Exercise 7.2.4
v = (x2z, xy, xz2), S is the portion of the paraboloid z = 9 – x2 – y2 above the
xy-plane, oriented upward
Exercise 7.2.5
Use Stokes’ Theorem to evaluate v · dr, where v = (x2 + 2xyz, x2y2, x2y3)
and C is the intersection of x2 + y2 = 1 and z = x – y oriented clockwise
Exercise 7.2.6
Let , R is the unit circle x2 + y2 ≤ 1, z = 0.
Compute both sides of the formula in Stokes’ Theorem. Does this example
contradict Stokes’ Theorem? Hint: Implicit in the development of the
theorem are hypotheses about domains. Also implicit is the continuity of
partial derivatives of the components of v in xyz-space. These issues are
accounted for in axiomatically rigorous presentations of the theorem.
Exercise 7.2.7
In electromagnetism, current through a surface generates a magnetic field
along the curve that is boundary to the surface; and vice versa. Let B(x, y, z)
represent the magnetic field corresponding to a steady current J(x, y, z). For a
surface S with ∂S = C, the integral form of Ampère’s law is:
Use Stokes’ Theorem to obtain the differential form of Ampere’s law, that is,
curl B = μ0J.
_________________________
38You will find all of this done in Calculus III textbooks.
39Why can we assume that the number is finite? It is because the examples of surfaces given are all
“compact.” The definition of compact is provided in more advanced texts.
40A similar derivation that uses vector product identities is: Harry F. Davis and Arthur David Snider,
Introduction to Vector Analysis, 7th ed. (Mt. Pleasant, SC: Hawkes Learning, 2000): 298.
Chapter 8
RELATIVE CHANGE IN
VOLUMES AND IN
INCREMENTS
Problem 8.1.
Using only computations from elementary geometry of vectors, and then
linear approximation, show that the rate of relative change in volume is
Definition 8.1. Since the rate is determined along streamlines, just as for 2-
d flows treated in Part 1, a 3-d fluid flow is called incompressible when
Problem 8.2.
lined up with the coordinate axes flows to rectangles also lined up with
coordinate axes.
Problem 8.3.
Show that, just as in the 2-case, linear approximation gives us that
Figure 8.2 Initial points A(x, y, z) and B(x + dx, y + dy, z + dz) flow to A′, B′. Use v(x, y, z) and v(x+dx,
y+dy, z+dz) and common time increment dt > 0 to approximate A′ and B′.
Problem 8.4.
As we did in the 2-d case, in a unique way, we can decompose the Jacobian
into a sum of the form
Observation:
and
As in 2-d, the real symmetric matrix (Jv)S can be diagonalized relative to
“principal coordinates.” See an undergraduate linear algebra book that treats
the real spectral theorem. The matrix (Jv)S is called the linear approximation
for pure strain and (Jv)A is called the linear approximation for pure rotation.
The reason for the name “strain” is that relative to the principle coordinates,
(Jv)S gives the linear approximation to relative change in increments along
coordinate axes. Why the name “rotation” is appropriate is brought out in
Section 8.3. (Recall that Section 3.9 treats the same problem in two
dimensions.)
In Section 7.2, ((cy – bz), (az – cx), (bx – ay)) was obtained not by estimating
relative changes in increments A′B′ – AB but by computing ratios of integrals
to obtain circulation density along a surface, per unit area.
This leads to the following question:
Question: Why is the same vector obtained as the answer to the two different
questions?
As we discussed above, there are three projections of v onto coordinate
planes, each of which has its own scalar curl. We have already sorted out the
connection between 2-d circulation and 2-d rotation. The problem here is to
assemble 2-d results in the 3-d setting.
The matrix (Jv)A is the anti-symmetric part of the Jacobian of a velocity field,
the components of which represent a system of differential equations for
streamlines. For now, let′s ignore the symmetric part. We will deal with that
shortly.
For (Jv)A, the system of differential equations is:
This represents pure rotation in the (y, z) plane, with angular velocity
where r(z, x) is the projection of the initial length r =|| AB ||=|| (α, β, γ) || onto
the (z, x) plane.
Each of the projections are rotational in 2-d.
What does that mean about the given flow in 3-d? Is the 3-d streamline also
rotational?
The projections are known and so
But r =|| AB || = || (α, β, γ) ||. Computing the right-hand side explicitly, we get
Problem 8.5.
The vector field curl v is independent of the relative strain.
(i) As noted at the beginning of this section, except for a constant factor
of , the components of (Jv)A are the same as the components of the
derived vector field
Problem 8.6.
We have two ways of defining the divergence of a vector field in 3-d: 1. as
a flux density; and 2. as rate of relative change in volume.
Problem 8.7.
By direct computation of derivatives, show that for any vector field w, we
get div (curl (w)) = 0. Hence, if v is of the form v = curl (w) for some
vector field w, then the flow v is incompressible.
Problem 8.8.
Suppose that v is the velocity field for a mass-flow in (x, y, z) space. For a
vector field in 2-d, we identify “orthogonal trajectories,” that is, “integral
curves” that are everywhere orthogonal to the original vector field. (See the
last paragraph of Section 5.3.) In 3-d, there is a similar construction. In 2-d,
by using a common parameter, the orthogonal trajectories are moving front-
lines in a mass-flow. In 3-d, we have a moving front-surface. We need the
equation for the surface that is everywhere orthogonal to the mass flow
given by v. In other words, the defining differential equation for such a
surface is of the form v · f = 0. At each point (x, y, z), there is a two-
dimensional vector space that solves the equation. The problem then is to
inquire into the possibility of integrating the system to produce surfaces that
are everywhere orthogonal to v.
Figure 8.3 Initial points A(x, y, z) and B(x + dx, y + dy, z + dz) flow to A′, B′. Use v(x, y, z) and v(x+dx,
y+dy, z+dz) and common time increment dt > 0 to approximate A′ and B′.
The main text was on vector calculus. Our intention is that this Supplement
will be of wider application. The purpose is to help teachers and students
make beginnings in describing (instances of your own) mathematical
development. As will become evident, this will be useful for all levels of
pedagogy.
Part of the novelty of our approach here is that, through examples and
directed questions, we invite you to grow in understanding your own
mathematical understanding and that way provide yourself with essentials for
growing as a teacher.
The meaning of the italicized statement will emerge and grow by doing
exercises such as those provided in Part A. What we are referring to will be a
personal achievement. For the teacher, however, it is rarely an altogether
private achievement. We say that because, of course, as teachers, what we
think about mathematical development (which includes one’s own
mathematical development) factors into what we attempt to promote in our
students.
Thought on teaching and learning mathematics goes back to antiquity. For
instance, although in a different context and not called “mathematics
education,” there were Plato’s reflections about the slave-boy who is helped
to solve the problem of doubling the area of a square (Meno). However, the
discipline now called “mathematics education” is mainly a 20th century
development.41 Good work has been done. For, instance, advances have been
made in “experiential learning” and “engaged learning.” However, looking at
the entire discipline, results have been mixed and opinion varies regarding
the extent to which methods have been effective, or not.42 In Part C, we
provide a few comments regarding these issues.
What we would like to do first, however (in Part A), is invite you to a few
relatively novel43 exercises in mathematics. There is no “conceptual model”
or “representational system.” Nor do we draw on or appeal to student test
scores, classroom observations, student task orientation, student cooperation,
student behavior or other evidence that mathematical learning might or might
not be happening for someone else. It’s not that such results will not
eventually contribute to progress in understanding mathematical development
in history. The focus that we invite here is more elementary. We invite you to
“go to source.” The mathematics will be familiar. Part of what is new is that
you are invited to a precise “puzzling about your own puzzling in
mathematics.” The focus, then, is you, and me, or rather, at least initially,
you-about-you and me-about-me. The invitation is to make initial progress in
being able to advert to,44 focus on, inquire about and discern orderings of
distinct events in our own inquiry and understanding, in instances, when we
are doing mathematics.45
We introduce a key diagram (Figure A.1) which could — in a sense — be
said to be a “model.” But it is not a “conceptual model” or “speculative
framework.” Think more of something analogous to what the physics
community has been getting to, namely, a “best-to-date standard model”
which, for the most part, was discovered in and continues to be verified (or
not) in instances, in experience. Or again, in chemistry, one could say that
the periodic table is a “model.” However, the periodic table emerged from
and has been established through centuries of ongoing experimental work
investigating chemical dynamics. In a somewhat similar way, we think that
you will begin to see that Figure A.1 also is no mere model. By turning
attention to our own experience, in exercises in Part A, Figure A.1 emerges
and is verified. It too is a “table of elements,” that is, “elements of dynamics
of knowing.” Note that Part A is only a first few “experiments.” We leave it
to interested readers to go further, to begin exploring the significance of
Figure A.1 in other instances in your own mathematical development (and
more).46
What all of this may have to do with improving teaching in mathematics will
be touched on briefly in part B. Part C draws attention to a few anomalies in
contemporary mathematics education — in both content and method. A
challenge for the education community will be to take advantage of good
work that has been done but to also make progress in getting a handle on the
various anomalies and misdirects. To do so effectively will not be easy. It
will need global collaboration. For this Supplement, it will be enough to tease
a few key issues “into view.”
_________________________
41Alan H. Schoenfeld, “Research in Mathematics Education,” Review of Research in Mathematics
Education 40, issue 1 (2016): 497-528. Philip s. Jones, ed., A History of Mathematics Education in the
United States and Canada, Thirty-Second Yearbook. National Council of Teachers of Mathematics.
Reston, VA: National Council of Teachers of Mathematics, 1970 (second printing 2002).
42Dawn Leslie and Heather Mendick, eds., Debates in Mathematics Education. Abingdon, Oxon:
Routledge, 2014.
43The approach is not original but will be new for the science of mathematics pedagogy. An advanced
source is: Bernard Lonergan, Insight: A Study of Human Understanding, eds. Frederick E. Crowe and
Robert M. Doran, vol. 3 in The Collected Works of Bernard Lonergan. Toronto: University of Toronto
Press, 1992. Introductory level works are: (1) John Benton and Terrance Quinn, Journeyism, 2018,
https://bentonfuturology.com/journeyism/; (2) John Benton, Alessandra Drage and Philip
McShane, Introducing Critical Thinking. Halifax, Canada: Axial Publishing, 2005 (Reprint 2006).
(This book has been translated into Spanish (Madrid: Plaza y Valdés, 2011); and (3) Philip McShane,
Wealth of Self and Wealth of Nations: Self-Axis of the Great Ascent, Hicksville, NY: Exposition Press,
1975. (This is available for free at http://www.philipmcshane.org/wp-
content/themes/philip/online_publications/books/wealth.pdf.) Compared to the present
Supplement, references (1), (2) and (3) are far broader in their coverage. The Supplement focuses on
mathematical development and pushes into further examples.
44The (intransitive) verb ‘advert’ is convenient. We are using it in its first meaning in the Merrian-
Webster: “advert, intransitive verb. 1: to turn the mind or attention – used with to.”
45Occasionally, we find it convenient to use a single expression to refer to the whole inquiry: “self-
attention.” This is meant in the precise sense that will begin to emerge by doing the exercises in Part A.
The turn “to attend to one’s own understanding” can seem strange if one has been trained to focus
mainly on models.
46See references in note 42.
PART A
Mathematical Understanding
A.1 A diagram
Much in the way an introductory level chemistry text may include a
simplified periodic table in the front cover, we begin by providing a
simplified version47 of a key diagram, a “table of elements of knowing.” The
diagram will be developed as we go.
Perhaps you have already discovered a (possible) solution and can now
continue the sequence.
If so, something happened. There was a change “in” you. A name for that
“all-of-you-event” is insight.
Let’s call this a direct insight.
Why do we need the adjective ‘direct’? Partly, it is to distinguish this insight
from what turns out to be a second kind of insight that can emerge in a
follow-up inquiry-poise discussed below.48
For this sequence-puzzle, if you’ve had a direct insight, then you’ve gotten
hold of something, a possibility.
From that (act of) insight, automatically as it were — but with no implication
here of machinery or technology — there is a “procession” in you.
Why do we say “procession”? Focusing on an image in one’s sense-ability,
being “lit up” by insight “into” an image in one’s sense-ability, not only does
a light “go on” in us in that concrete image-focus, but light immediately also
“goes on in us.” There is in us, as it were, “light from light” in the sense that
following insight, a solution “emerges in us.” There is the famous story of
Archimedes, his calling out “Eureka!” As the Greek verb says, when we
discover the solution to the sequence puzzle, it is not just an insight.
Forthwith, and forth and with, there is also an “I’ve got it.” We are not
playing a grammar game here but are inviting your attention to an event.
There is an ‘it’ in “I’ve got it.” There is something that is not merely what we
had in our sense-ability. There is something that emerges “in us” from our
insight-into-image, something that in many respects is “transferable.” For
instance, following direct insight, we can turn our attention elsewhere but
then later return to the sequence-puzzle. The same solution then comes to
mind but without labor.
Why is it without further labor? Because we already get the point. We already
“got it.” It will be convenient to introduce a name. For historical reasons, let’s
call the “something that emerges in us,” the “it” of “I’ve got it,” “inner
formulation.”49
Note that the adjective “inner” is not to suggest that what we get hold of is
somehow “spatially inside” (versus a “spatially outside”). The adjective is
merely a metaphor. Inner formulation is a fact. It is what we have in as much
as we have, in fact, discovered a possible solution. Note that inner
formulation is to be distinguished from “formulation” in the sense of
providing a “formula.” Although, evidently, in some cases, a formula can
precede insight, and also be an expression of inner formulation.
An expression for what we have just described is given by the bottom two
rows of Figure A.1:
Figure A.3 Wonder about data50 of sense, insight and inner formulation
Judgment
Note that in (A.1) and (A.2) we find the same core pattern expressed by
Figure A.1. We begin with inquiry about an “image” in sense-ability, and “a
focus inquiry.” In both cases, a What is it?’ inquiry-poise is resolved through
direct insight, following from which there is a procession “in us” to “inner
formulation.”
At the same time, there are significant differences in what is being
understood. As you might expect, grounds for judgment also turn out to be
different.
In example (A.1), direct insight is of a “pattern grasped in our sense-ability.”
Is our solution correct? Is it so? By adverting again to patterns in our sense-
ability, we find and discover sufficient grounds (or not) for assenting to our
possible solution. There is then a procession to “Yes, it is so,” or perhaps,
“No it is not so.”
In example (A.2), however, what we discover is a pattern of mutually defined
terms, and operations that, as it happens, can be both presented and expressed
in many ways.61 In that case, sufficient grounds for assenting to our possible
solution are not “patterns in sense-ability” but whether or not one can,
indeed, successfully continue the sequence by appealing to our directly
grasped (possible) pattern of mutually defined terms, and operations.
Figure A.4 Reflective inquiry and reflective insight in elementary sequence puzzle A.2.1
A.5 Proofs
My beginnings in calculus, by Terrance Quinn
In September of 1980, in the first weeks of my first-year calculus course at
the University of Toronto, the lecturer63 wrote a theorem on the blackboard
about continuity of a function y = f(x). He also provided a proof that went
something like this: “Suppose that ϵ > 0 and let . · · · ” He went on to fill
two boards with inequalities and ended with words similar to: “· · · from
which we can conclude that |f(x) – f(y)| < ϵ. Since ϵ > 0 was arbitrary, the
result follows. ”64
I trusted our excellent teacher and had no doubt that “the result did follow.”
But during that lecture I also knew that I didn’t yet see why. That evening, I
started in on trying to unpack the proof. I found some of the inequalities
within the proof (“steps” in the proof) to be a bit tricky. Before too long,
however, I could follow the proof “step by step,” inequality by inequality. I
could see that, yes, each step of the proof worked. But I remember that it was
also obvious to me that I still didn’t really have it. What I mean is, I was
aware that even though I could check all of the steps, I didn’t see how I might
have produced a similar proof myself. Up to that point, in my mathematical
education, I had been naively thinking that I already had a good
understanding of calculus. In some respects, I did. However, the first weeks
of that first semester of calculus were a shock to me and — as I learned —
for some (not all) of my classmates. Eventually, along with other “survivors,”
I started to find my way. For me, the “epsilon-delta business” crystallized in a
problem that involved a damped sine function defined over the real line. In
hindsight, I see now that my insight in that case was similar to Archimedes’
insight regarding the remainder term of a series. Problem by problem and
insight by insight, I crawled my way into modern axiomatics for calculus,
a.k.a. “elementary real analysis.”65
In that first-year class, a common type of problem was of the form: “Prove
(or disprove) X (where X was some statement).” How to begin? I soon
learned that, for me, unless it was “asily done”(in other words, unless I
already understood), it was often best to not start with the general statement.
It usually worked better for me — that is, I worked better — by starting with
examples. Of course, I kept the X in mind and oscillated back and forth
between examples and thinking about the X statement. But inevitably, the
break for me would come while working on an example. That started to take
some the mystery out of “proof writing.” In other words, “being able to write
a proof is not separate from understanding.”
I noticed that the order in which the write-up proceeded usually turned out to
be, to some extent, in a kind of reverse order to what I might call “the order
of inquiry and discovery.” I now realize that was no accident. I have been
recounting some of my early struggles in a challenging and exciting first-year
calculus course. But to begin homing in on key and core issues, it is better to
now shift to a more elementary example, essential features of which are
already described by another author.
Understanding and syllogism
Our search, through diagram, was for a middle term, and the middle term was
supplied as soon as one adverted to the significance of OP. Only then can the
syllogism be constructed. To coin an expression for this constructing, one
might say that the insight is crystallized into a syllogism. This does not mean
of, of course, that somehow insight has been pinned down on a page. What
has happened is that we have given the insight symbolic expression. Giving
all the relevant insights symbolic expression is by no means always an easy
thing to do, even when it can be done. Modern geometers have found fault
with Euclid in this matter. There are insights in the Elements which are not
explicitly acknowledged either in axioms or in the theorems, yet which were
not uncrystallizable.66
Let me note further that our simple puzzle and solution is a paradigm of how
Euclid and company may well have proceeded. They did not proceed step by
step down the page of a modern textbook, from the stated theorem to the fully
constructed diagram, to the step by step deduction beneath.67
We may return now to a brief consideration of the simple symbolic
expression of understanding which is the (mathematical) syllogism.
[In mathematics, the syllogism is a] help toward understanding.
The syllogism is not some mysterious replacement for understanding.
One may look at the syllogism as a proof of the conclusion, but this can only
mean that the structure facilitates a grasp of the implication of the conclusion
in the premises.
One might note, too, that such structures facilitate the checking, the Is-
question, relating to that grasp.68
Physical lengths can and do change. Two people with different arm lengths
will yield different cubits. A piece of wood may dry out and its length may
then change. Or again, suppose we have two metal rulers with “cms” etched
along their sides. In other words, “1 cm” is the unit length. If one of the rulers
is used at different locations on a hot wood stove then “1 cm” at one location
can be longer than “1 cm” at another, when compared with the second ruler
that is kept from direct contact with the hot stove. But if there is no fire in the
stove, and if no finer measurements are needed then, for practical purposes,
no problems arise. That is, the unit length is “invariant.” hrough the ages,
builders, carpenters, architects and engineers have relied on such invariance
when designing and building structures — from wooden shacks to pharaohs’
pyramids, from modern homes to city skyscrapers.
In classical geometry, we “ake this idea and run with it.” We suppose that an
imagined unit length is invariant in the sense that an imagined unit length at
one location is the same as an imagined unit length at another.
Let’s now return our focus to the x-axis. Let Ox be a (distinguished) point on
the x-axis. The point Ox is called “origin.” Let X be another point on the same
axis, a point that is, say, two units’ “to the right” of Ox. Notice our ongoing
reliance on imagination, e.g., “to the right,” and “to the left.”
We have an imagined x-axis which is an imagined distinguished line in the
Euclidean plane. By insight, we suppose that the line extends indefinitely in
both directions:
In the imagined length between these two imagined locations, one can
imagine fitting copies of the imagined unit length, “ _____ ” and “ _____ ”,
distinct line segments that, by hypothesis, are otherwise identical. In
particular, we suppose that they have the same length.
A handy diagram for this is:
Concepts
In Euclidean geometry, we speak of ‘points,’ ‘lines,’ ‘line segments,’
‘length,’ invariance,’ and so on, as well as the entire Euclidean plane.’
What are these?
What, for instance, is a point?
We can imagine a small dot on paper (or, say, on a computer screen). An
imagined dot has breadth and so we can imagine two smaller dots within an
imagined dot. Still, “the point” is “clear,” is it not? That is, attempting to “pin
down” a location unambiguously, we need to keep making an imagined dot
smaller. So long as a dot has any breadth or depth, further dots can be
imagined within the imagined dot.
By insight, then, we define a point on the plane to be “a location that has no
breadth and no depth.”
Or, in translation from Book I of Euclid’s Elements: “A point is that which
has no part.”
A point then, (self-) evidently is the fruit of insight! It is also a concept.73
In a similar way, that is, by adverting to experience, one can find that lines,
planes, invariant length, and so on, are also the fruit of insight and also are
concepts.
Continuing in this way, what becomes (self-) evident is that while it is
true that images and patterns in our sense-ability on which we focus
inquiry can be said to be “primitive,” points, lines and other concepts are
the fruit of understanding and, in particular, are neither reducible to
imagination nor are they primitive elements in mathematics.
_________________________
47A more nuanced version is available in Bernard J. F. Lonergan, “Appendix A, Two Diagrams,” in
Phenomenology and Logic: The Boston College Lectures on Mathematical Logic and Existentialism,
eds. Frederick E. Crowe, Robert M. Doran, vol. 18 in The Collected Works of Bernard Lonergan,
Toronto: University of Toronto Press, 2001, 319–323. A diagram for the “Dynamics of Doing” is the
second diagram in Lonergan, “Appendix A, Two Diagrams.” Evidently, our “dynamics of doing” are
similar to our dynamics of knowing. For decision, however, inquiry is in a different mode. Our
dynamics of doing subsumes our dynamics of knowing. The interweaving of the two modes of inquiry
are brought out in student exercises in John Benton and Terrance Quinn, “The Dynamics of Doing,”
Journeyism 16 https://bentonfuturology.com/journeyism16/ and Jour-neyism 17,
https://bentonfuturology.com/journeyism17/. The dynamics of doing are not an immediate focus of
this Supplement. However, it is worth noting that progress in adverting to and describing dynamics of
doing also is needed. Among other things, such progress will help resolve numerous contemporary
issues in mathematics education where, so far, dynamics of the two modes are not adverted to and
consequently, not yet adequately distinguished. For instance, it is sometimes said that a teacher’s job is
to help a student “decide” which concept or solution to accept. The plausibility of such notions emerges
from the mistaken model that mathematical understanding is a matter of connecting concepts. The fact
that that view is mistaken is revealed — becomes (self-) evident — by doing the exercises given
throughout Part A. For context in mathematics education, see Part C.
48See below: Is?-ing and reflective insight.
49Lonergan, Phenomenology and Logic, Dynamics of Knowing, “Appendix A,” 322. See also
McShane, Wealth of Self, 15.
50See, e.g., Benton and Quinn, “The Dynamics of Knowing: The First Four Boxes,”
https://bentonfuturology.com/journeyism10/; and McShane, Benton, Drage and Mc-Shane, ch.
16, “What-Questions,” 67–69.
51See note 45.
52I am using the first meaning of the word ‘radical,’ referring to ‘root.’
53That is not easy. See, e.g., note 69.
54If you haven’t already solved the puzzle, that is a clue. See also the personal anecdote at the end of
this section.
55Answer to sequence puzzle A.2.1: Symbols that are written only with straight edges go above the
line, while symbols that include curved edges go below the line.
56The Fibonacci sequence is a sequence of numbers of fertile pairs of male and female rabbits. Start
with the hypotheses given by Fibonacci (about fertile male and female pairs, gestation periods and the
numbers of progeny). The problem is within reach of the contemporary senior high school or
undergraduate student. It is interesting and well worth doing. Discovering the solution for oneself one
takes a step in a climb toward the modern theory of recursive sequences.
57Long before Fibonacci, the same sequence is found in ancient Sanskrit texts, written in a base 60
number system. See, e.g., Tia Ghose, “What is the Fibonacci Sequence?” LiveScience (October 24,
2018): https://www.livescience.com/37470-fibonacci-sequence.html; and Keith Devlin, Finding
Fibonacci: The Quest to Rediscover the Forgotten Mathematical Genius Who Changed the World.
Princeton, NJ: Princeton University Press, 2017.
58From a “higher viewpoint,” we could also identify the solution as follows: “The sequence in (A.2) is
the (unique) solution to a second-order homogeneous recurrence relation with constant coefficients
whose (non-reduced) characteristic equation is t2 – t2 – 1 = 0, and whose first two initial-values are 1
and 1.” See Section A.6.
59More generally, descriptive understanding is “grasping patterns in experience, as experience.” And
so to discern events and orderings of events in our dynamics of knowing is a beginning but it, too, is
descriptive. Progress in explanatory understanding of our dynamics of knowing will be future growth
for the academic community. Aspects of that future progress are already partly in evidence. And so,
eventually, it will include an “integration” of human biophysics, biochemistry, human zoology and
psychology, cogni-tional theory and more. To glimpse something of the challenge, precise densely
expressed heuristics can be found in Lonergan, Insight, 489 (add the word ‘self’ to the paragraph that
begins “[(Self-) S]tudy of an organism begins ...”).
60For a discussion in elementary contexts, see Terrance J. Quinn, “On Two Types of Learning (in
mathematics) and Implications for Teaching,” On Learning Problems in Mathematics, Research
Council on Mathematics Learning, Mathematical Association of America, FOCUS, Fall 2004, 31–43.
61See, e.g., note 60.
62In calculus, there are two key basic insights, basic in the sense of prior to axiomatics. A satisfactory
definition of limit did not emerge for several decades after the initial discoveries of calculus. But both
Newton and Leibniz each had those two key insights (and, of course, much more). Once shared with
the scientific communities, those two key insights soon made possible the solution of problems from
antiquity and also opened up vast ranges of new lines of inquiry and development in mathematical
sciences and engineering. See Terrance J. Quinn, “Getting Started in Calculus,” Problems, Resources
and Issues in Mathematics Undergraduate Studies (PRIMUS), vol. 13, issue 1 (March 2003): 55–74.
63I am referring to Prof. Edward Bierstone who, as I understand it, is still at the University of Toronto.
See https://www.math.toronto.edu/bierston/. Accessed July 17, 2019.
64The textbook in the 1980–81 academic year was the edition then available of Michael Spivak,
Calculus. Cambridge: Cambridge University Press.
65Historically, axiomatics for calculus were obtained after several decades of collaboration following
the initial breakthroughs of Newton and Leibniz.
66It concerns drawing a line across the interior of a triangle. One identifies the point where the line
crosses the opposite side. Why is there a point of intersection? Felix Klein discusses the problem: “Of
course, no one would doubt that, intuitively, but in the framework of axiomatic deduction we need a
special axiom, the so-called betweenness axiom for the plane. This axiom states that if a line enters a
triangle through a side, it must leave it through the other side — a trivial fact of our space perception
which requires emphasis as such, because it is independent of the other axioms. . . . If we omit this, as
Euclid does, we cannot reach the ideal of a pure logical control of geometry. We must continually refer
to the figure” (Felix Klein, Elementary Mathematics from an Advanced Standpoint: Geometry
(Mineola, NY: Dover Publications, 2004), 88. This is an unabridged republication of an earlier Dover
reprint (1949) of the translation first published by The Macmillan Company, New York, in 1939). As is
well known, there are non-Euclidean geometries for which the axiom does not hold. For further
discussion, see McShane, Wealth of Self, 68–9.
67“One should note, too, the significance of this for the teaching of geometry. Too often pupils [are
asked to] begin, not with the thrill of a puzzle but with the top of the page, and at most get a vague line-
by-line comprehension of the theorem. Memory is burdened, and examinations consists in filling out
theorems - from the bottom and top of the page! — and passing over the riders” (McShane, Wealth of
Self, 69). This provides foreshadowing to Part B.
68Philip McShane, Wealth of Self, 67-9.
69Here, ‘abstract’ is a transitive verb for what we do in thought: “abstracter/,” to draw away, remove,
derivative of abstract (or borrowed directly from Latin abstractus)” https://www.merriam-
webster.com/dictionary/abstract.
70As you might already be seeing, this has implications for teaching. See Part B.
71The task is further described in Section A.8.
72What is a straight line in Euclidean geometry? (Here, we avoid the extensive literature regarding
“primitive concepts,” “primitive,” and “vicious circles in logic.” Such problems need to be handled, but
not here.) In Book I of the Elements, Euclid provides what evidently are nominal definitions: “Def. 1.2.
A line is a breadthless length.” “Def. 1.4. A straight line lies equally with respect to the points on
itself.” Explanatory definition is obtained when a straight angle is defined to be the sum of two 90-
degree angles. In that way, we go beyond merely describing what we imagine. In that case, a straight
angle is defined by correlating terms in the system.
73“concept: noun, (Entry 1 of 2) 1: something conceived in the mind: thought, notion, 2: an abstract or
generic idea generalized from particular instances” https://www.merriam-
webster.com/dictionary/concept.
74Lonergan, Insight, 56.
75See note 50.
76Bernard Lonergan, Early Works on Theological Method 2, vol. 23 in the Collected Works of Bernard
Lonergan, translated by Michael G. Shields, edited by Robert M. Doran and H. Daniel Monsour
(Toronto: University of Toronto Press, 2013): 175–177. For a somewhat dated but still excellent source
on mathematical development, see, Eric Temple Bell, The Development of Mathematics. New York:
McGraw-Hill, 1940. Second Edition: New York, McGraw-Hill, 1945. Reprint: Dover Publications,
1992. University of Toronto Press, 2013): 175–177. For a somewhat dated but still excellent source on
mathematical development, see, Eric Temple Bell, The Development of Mathematics. (New York:
McGraw-Hill, 1940. Second Edition: New York, McGraw-Hill, 1945. Reprint: Dover Publications,
1992).
PART B
It is for brevity that we merely state a few results. To bring these out
pedagogically would need a few chapters. However, we hope that providing
these few results will serve as a guide and invitation to further reflection. If
you have worked through the exercises in Part A (or similar ones), you may
get to some (or all) of these yourself. As pointed to in Section A.8, further
details and implications will be discovered through one’s ongoing growth in
mathematics and in self-attention in mathematics.77 We do not comment on
curricula of any departments or colleges involved in the ongoing challenging
and creative work of meeting students’ needs and program needs in diverse
circumstances.
Part of the invitation is for a teacher to make progress in identifying and
distinguishing nominal and explanatory understandings, in one’s own
understanding.
As experience reveals, nominal understanding includes understanding by
which one reaches competence with symbolic techniques and computations.
The task of pedagogy invites teachers to know as much as possible about
where, in precise terms, the content of a course fits in the historical
development of the field. See Section A.8.
By the same token, in order to be able to help a student make progress in the
field, it will help if a teacher learns as much as possible about where a
student’s understanding is in relation to the historical development of the
field.
As expressed by Figure A.1, by adverting to one’s experience in doing
mathematics, it becomes evident that understanding emerges through inquiry
“into” images in one’s sense-ability. Consequently, in a given mathematical
context, it will help if a teacher is familiar with diagrams and symbolisms that
help raise and direct student inquiry toward reaching key insights appropriate
to a context.
To help a student reach a higher viewpoint, it will help if a teacher can
provide diagrams and symbolisms from lower viewpoints by which a student
(or group of students) can reach key insights needed to begin breaking into
the higher viewpoint. See Section A.6.
For a century and more, there has been ongoing discussion about whether or
not teachers should “use history” to teach mathematics. An advantage of the
“historical method” is that it has the potential for helping students raise
questions and reach key insights in the field. But of course not all of the
searchings from history can be included. A selection needs to be made and
tailored to an audience and to a curriculum. In that sense, a teacher needs to
“make history better than it was.” More precisely, a teacher can make
progress in identifying specific developmental sequences.78 In that way, for
instance, in two leisurely 45 minute classroom sessions one can help senior
high school mathematics students and freshmen level general calculus
students reach the two key insights had by both Newton and Leibniz from
which the entire body of calculus is developed.79
Pedagogy called engaged learning has been a fruitful idea. The importance
and relevance of modern psychology and neuroscience (so, e.g., behaviour,
environmental considerations, learning style, group work, and so on) to
pedagogy cannot be denied. But might we not also ask, engaged in what? We
are back at the need and possibility of identifying helpfully directed diagrams
and symbols by which to subtly guide student inquiry toward specific and
key mathematical insights.
Only a small percentage of students go on to become mathematics majors
who, in addition to needing basic80 key insights, also go on to (sequences of)
axiomatic systems. But students in applied mathematics, engineering,
applications and general service courses also need to be helped so that
inquiry be directed toward the emergence of initial or basic key insights.
Why? Without basic insights (which, as experience reveals, precede the
emergence of axiomatic system), understanding is mainly nominal. Even if a
student is only wanting to be able to do routine computations in a practical
career, if understanding is merely nominal, they will be unable to proceed
whenever boundary conditions do not sufficiently mesh with what is only
nominally familiar.
Modern textbooks speak of “understanding concepts” and “conceptual
understanding.”81 It is said (or assumed) that “concepts are primitive
elements” and that mathematical understanding is a matter of “connecting
concepts” (that one has prior to understanding). It is also now popular to
assert analogies between structures of computer programs and human
understanding. There is an Is it so?’ question here. By adverting to instances
in our experience (see, e.g., Section A.7), it is (self-) evident that while there
are “primitive elements” in mathematics (for instance, in Euclidean
geometry), it is not concepts that are primitive. And the experiential basis of
alleged analogies with computer programming are not found in mathematical
understanding. They are obtained, rather, by correlating structures of
computer programs with structural features of hypothetical conceptual
models of mathematical understanding. On the other hand, adverting to
experience in mathematical understanding reveals that points, lines, invariant
unit length, and other familiar concepts emerge from insight. In other words,
concepts are the fruit of understanding.
Relative to arithmetic, algebra is a higher viewpoint. As already described,
one reaches a higher viewpoint by abstracting82 from instances in a lower
viewpoint.
This sheds light for us on the problem of giving students calculators and other
computational technologies too soon. If a student does not have a good
understanding and technical competence in arithmetic, then they are lacking
most if not all of the symbolic data, diagrams and understanding needed in
order to break through to the higher viewpoint of algebra. Similarly, if a
student lacks nominal and explanatory understanding in algebra and
coordinate geometry, they will be lacking most if not all experience needed in
order to break through to the higher viewpoint of calculus. Again, there are
transitions from algebra to abstract algebra, from calculus to function theory,
and so on.
Evidently, building up an axiomatic system relies on understanding that is
beyond initial or basic understanding (see, e.g., Section A.5). This helps
reveal a common misdirect, that is, when a chapter or lesson begins with
“preliminary concepts,” “axioms” and/or “axiomatically correct definitions.”
Certainly, that approach is dull and is a way to undermine engaged learning.
Attempting to start with axioms and preliminary concepts is, in fact,
attempting to start with answers to questions from a further context. The
approach does not bring the student into an inquiry zone that promotes
emergence of initial key insights. But this is another face of the
misconception in mathematics education that is often called “conceptual
understanding.”83
Whether implicit or explicit, a teacher’s view of mathematical development
influences how and what we teach. Whatever one’s context in teaching, and
whether private or shared with colleagues, there is the possibility of reflecting
on one’s own experience in mathematics and on what one hopes one’s
students also will learn. In other words, as invited in Part A, there is the
possibility of growing in being able to advert to and precisely distinguish
elements in one’s own experience. If one is already a successful teacher, the
invitation is to accurately identify sources of one’s success and that way
become an even more effective teacher. If one is struggling as a junior
teacher, the invitation can include the task of identifying gaps in one’s
mathematical understanding, inconsistencies in what one understands, and
elements of one’s own understanding (specific combinations of nominal and
explanatory, in and across diverse contexts), and to make efforts to bring
one’s learning and teaching into better harmony with the historical field and
the needs of one’s students.
_________________________
77See note 45. “balanced method.”
78See note 76.
79See note 62.
80See note 62.
81See Part C.
82See note 69.
83See three paragraphs above. On the other hand, ‘conceptual understanding’ is just a name. Many
excellent teachers use the term in a positive sense, namely, when referring to student understanding that
goes beyond mere technical competence. Implicitly, they are referring to both nominal and explanatory
understanding in mathematics. The name “conceptual understanding” however is, in fact, inherited
from the philosophical literature and in that context refers to what does not occur in mathematics. That
is, as experience reveals, we do not understand concepts but, rather, concepts emerge from
understanding.
PART C
Note that, in the first sentence, the authors express a lack of concern for
whether or not their theory is “a statement of truth.” They then remove
themselves from the task of inquiring into “what is really happening when an
individual tries to learn”: “this is not our focus.” Their focus is, instead,
“phenomena that we can observe in students.”89 The last sentence of the
paragraph reveals that in their inquiry into learning, they presuppose a theory
of learning. That this, they presuppose a version of constructivism which, in
this context, is to the effect that students “learn one or another concept in
mathematics”; that they “construct their understanding of mathematical
concepts”; (later in the paper) that when learning mathematics one
“[perceives] mathematical problem situations”; and that, when learning, “an
individual is developing her or his understanding of a concept.”
If you have done exercises such as those in Part A (or beyond) and have
made beginnings in understanding your own understanding in mathematics,
do not the claims made in the paper cry out for correction?90
Perhaps, though, the resulting pedagogy stands. So, let’s look at what
Dubinsky and McDonald suggest for teaching cosets:
Evens and odds, “clock arithmetic” and other similar arithmetic and
geometric groupings are learned by children. In that way, students learn much
of what is needed in order to prepare the way for reaching a higher viewpoint
that includes an understanding of cosets. In our own mathematical
development, we learned mathematics, and did not make use of syntax
“similar to standard mathematical notation.” Indeed, in the historical
development of cosets, founders did not appeal to computer programming
language. Is it not evident that syntax (of computer programming) “very
similar to standard mathematical notation” is, in fact, a distraction from the
mathematical problem?
Why did Dubinksy and McDonald advocate that approach to teaching cosets?
Why attempt to use something similar to mathematics to teach mathematics
instead of helping students learn mathematics itself? Indeed, why might that
be relevant (or not) when “what is really happening” is, as they suggested,
not the focus and not known? They explain it as follows:
Their design instruction focused, not directly on mathematics, but on some
model of how the topic in question can be learned.91
Dubinsky and McDonald had a remarkable dedication to the cause of
mathematics education. As scientists, however, they chose to not focus on
“what is really happening” but rather on “phenomena that we can observe in
students.” In that approach, they also “took themselves out of the equation”
and so did not avail of experience in mathematical understanding.
Constructivist models allegedly account for mathematical understanding.
And so, to reveal the presence of flaws, all that is needed is a
counterexample. But by doing exercises in Part A (or similar ones), one can
obtain numerous counterexamples across a range of instances in elementary
mathematics.
_________________________
84Schoenfeld, 2016. For a survey of “theoretical frameworks used in the field of Calculus education,”
see David Bressoud, I. Ghedamsi, V. Martinez-Luaces, G. Torner, G. Teaching and Learning of
Calculus, ICME-13 Topical Surveys, Springer Open (2016) 1–37.
https://www.springer.com/gp/book/9783319329741.
85See, e.g., Schoenfeld, 2016.
86That will be a community achievement, coming from a new kind of collaboration on a global scale.
87E. Dubinsky and M. A. McDonald, “APOS: A constructivist theory of learning in undergraduate
mathematics education research,” in D. Holton et al. (Eds.), The Teaching and learning of Mathematics
at University Level: An ICMI Study (Dordrecht, Netherlands: Kluwer Academic Publishers, 2001):
273–280.
88Dubinsky and McDonald, 2001.
89If one’s focus is on “phenomena that we can observe in students” then inquiry will not be about
mathematical insight but, rather, patterns in “phenomena what we can observe in students” who do not
yet understand. Note, too, that if(as in APOS and its applications) those phenomena are not
explanatorily defined, statistical analysis has little or at most preliminary explanatory significance.
90Regarding concepts, see Section A.7. The confusion is historical and not unique to the work of
Dubinsky and McDonald.
91Ed Dubinksy, “Using a Theory of Learning in College Mathematics Courses,” MSOR Connections
(2000), 1.10.11120/msor.2001.01020010. This article was originally published in Newsletter 12,
TaLUM, Teaching and Learning Undergraduate Mathematics subgroup, 2001. It is available online:
http://www.math.wisc.edu/wil-son/Courses/Math903/UsingAPOS.pdf. Accessed July 17, 2019. The
tradition continues. The literature is extensive. For a point of entry into the literature see, e.g., L.
Benton, C. Hoyles, I. Kalas and R. Noss, “Designing for learning mathematics through computer
programming: A case study of puplis engaging with place value,” International Journal of Child-
Computer Interaction, vol. 16 (2018), 68–76.
92In Scotus’ speculative theory of knowing, “[i]ntellectual intuitive cognition does not require
phantasms; the cognized object somehow just causes the intellectual act by which its existence is made
present to the intellect. As Robert Pasnau notes, intellectual intuitive cognition is in effect a ‘form of
extra-sensory perception’ (Pasnau [2002])” (Williams, Thomas, “John Duns Scotus,” The Stanford
Encyclopedia of Philosophy (Spring 2016 Edition), Edward N. Zalta (ed.),
https://plato.stanford.edu/archives/spr2016/entries/duns-scotus/. See
https://plato.stanford.edu/entries/duns-scotus/#MatForBodSou.
93The method is not new but will be new in mathematics education. See note 43.
94Although, the basic structure given in Figure A.1 is invariant. If one claims that a different structure
accounts for one’s understanding, is it because one understands something about an understanding that
one does not have; or that one judges what one does not understand? In the terminology of philosophy
of science, there results what is called ‘performance-contradiction.’ Philosophical terminology is not
the main issue here. As in the text above, a fundamental question always is: “Does a model explain
instances, in detail, in my experience?” Is this me?
Index
Base, 199
Boundary
circulation along, 166, 171
curve, 95
positively oriented, 99
surface, 158
Chain rule, 38
applications to integration, 46
in one dimension, 41
in two dimensions, 50
Circulation, 73, 74, 81
2-dimensional, 121
3-dimensional, 176
along a curve, 76
circulation integral, 79, 82, 83
Coordinate geometry, 1, 26, 36, 209
Coordinates, 26, 28
polar, 53
principle coordinates, 109
rectangular, 55
Counterexamples, 224
Cross-product, 1, 34, 36
Curl, 99, 111, 132, 141, 142, 166, 175, 176
Curve
boundary, 95, 116
closed, 95
positively oriented, 95
streamline, 68
Decomposition, 108
of the Jacobian, 139
Definition
of flux, 150
of incompressible flow, 131
of irrotational flow, 112
of Jacobian, 56
of symmetric matrix, 108
of total circulation, 77
sum of two curves, 85
Density
circulation, 95, 101, 103, 112, 125, 135, 141, 166, 175, 176, 182
rotation, 101
Derivative, 7
function, 7
of area under the graph, 23
Determinant
2-d, 29, 33
3-d, 30, 33
Divergence, 115, 119, 139, 147
2-d, 129
3-d, 157, 185
Divergence Theorem
2-d, 119, 125, 135
3-d, 119, 157
Dot product, 1, 26, 27
Euclid, 26
Galileo, 2, 67
Graph, 4–6
area under, 12
Green’s Theorem, 23, 86, 88, 94, 98, 115, 121, 135
for surfaces, 166
in the plane, 132
vs Stokes’Theorem, 173
Height, 13
Nobel Prize, 18
Normal, 36
field, 150, 155, 156, 158, 167
vector, 36, 117, 137, 159, 162, 176
Oriented
oppositely, 84
positively, 94, 99, 115, 170
Rate
net mass-flow rate, 116, 117
of change, 2, 5
Rectangular coordinates, 53
Region, 24, 58
partition, 96
simple solid, 158
Rotation, 36, 88, 101, 108
pure, 111, 182
rotational, 34, 141, 184
rotational displacement, 36
Scalar product, 26
Stokes’ Theorem, 165, 167, 173
Strain, 125
pure, 182
relative, 107, 180, 185
Streamlines, 68, 89, 101, 108, 113, 124, 133, 136, 142, 183
Sum, 12, 13, 55
of two curves, 85
Surface, 65, 81, 149, 152
2-dimensional, 67, 168, 176
boundary, 158
integral, 166
orientation, 158
Symmetric, 109, 139, 141, 142, 185
anti-symmetric matrix, 108, 111, 113, 183
symmetric matrix, 108, 132, 182
Telescopes, 22
Trace, 133, 135, 136, 139, 141
Variable
change of variable, 59
independent, 7
Vector calculus theorems, 1, 25
Velocity
angular, 111, 133
field, 65, 75, 108, 141
vector, 66, 68, 106, 149
Volume, 28, 152
3-d, 71, 156
rate of change, 179
rate of relative change, 180
relative change, 179, 185
Volumes, 28