The Art of Problem Solving
The Art of Problem Solving
Problem
Solving
Digitized by the Internet Archive
in 2019 with funding from
Kahle/Austin Foundation
https://archive.org/details/artofproblemsolvOOOOunse
The Art of
Problem
Solving
To our children,
Lisa Joan, David Richard,
Birgit, and Beate,
who provide us with constant reinforcement
to hone our problem-solving skills
The Art of
Problem
Solving
A Resource for the
Mathematics Teacher
Editor
Alfred S. Posamentier
Associate Editor
Wolfgang Schulz
All rights reserved. Use of the handouts, resources, and sample documents in each chapter
is authorized for local schools and noncommercial entities only. No other part of this book
may be reproduced or utilized in any form or by any means, electronic or mechanical,
including photocopying, recording, or by any information storage and retrieval system,
without permission in writing from the publisher.
95 96 97 98 99 10 9 8 7 6 5 4 3 2 1
Introduction vii
From the earliest times, a cornerstone of mathematics has been problem solving.
Too often, solving a problem is seen as a means to an end rather than as an end in
itself. In the collection of ideas presented in this book, we hope to change that
perception. Here, problem solving is the main theme. It is presented from the very
practical to the theoretical, from the common to the glitzy, from the purely
mathematical (viewing the beauty of mathematics for its own sake) to the peda¬
gogical and psychological considerations surrounding the "sharing" of this impor¬
tant discipline.
This book is intended for mathematics teachers. It was conceived as an effort
to bring them a host of interesting and useful ideas, thereby raising their conscious¬
ness level and enabling an enrichment of the mathematics instruction program.
To ensure a wide variety of ideas without repeating notions commonly found
in current literature, we drew from a team of authors representing a wide array of
geographical areas as well as a broad experiential range. The contributing authors
in this book are professionals who, in a variety of ways, influence mathematics
education internationally and range in age from the mid-30s to the upper 80s.
A key stipulation for the authors, who were given free rein in terms of space
and topics, was to discuss their favorite ideas in the realm of problem solving in
the most enticing manner possible, so as to make the book enjoyable and enter¬
taining as well as useful. This inevitably has led to an occasional example being
used more than once. We chose not to change these duplications because their
replacement or removal would have altered the individual contribution. Authors
were urged to write in an informal and easily understood style.
vii
viii THE ART OF PROBLEM SOLVING
The book can be used in a number of ways. For the most part, the chapters
are independent of one another and, therefore, can be read in any order. Our order
of presentation was chosen to provide a meaningful guide through the extensive
and complex world of problem solving. Naturally, other orderings not only are
possible but may even serve some readers better, so readers should feel free to
personalize the order in which they read the chapters. To facilitate other reading
possibilities we offer the reader two marginal guides: One will indicate the
problem-solving strategies being used, and the other will allow an easy subject
matter identification.
Throughout the book, a variety of problem-solving techniques are intro¬
duced, discussed, and illustrated. Similar techniques can manifest themselves
somewhat differently within various contexts. A reader might find it helpful, after
being stimulated by a particular technique, to locate other applications of those
particular problem-solving methods by means of the circled numerals in the
margins. The numbers and their corresponding techniques are as follows:
1. Working backwards
2. Finding a pattern
3. Adopting a different point of view
4. Solving a simpler analogous problem
5. Considering extreme cases
6. Visual representation (diagram, table, chart, etc.)
7. Intelligent guessing and testing
8. Determining necessary and sufficient conditions
9. Sequencing
10. Specification without loss of generality
11. Systematically accounting for all possibilities
12. Using a computer
13. Deductive reasoning
14. Organizing data
15. Approximating
16. Determining characteristics of objects
17. Specializing
18. Generalizing
The boxed "signposts" in the margins allow the reader to search for problem¬
solving discussions by subject matter strategy. The categories were selected to
reflect broadly the mathematics taught in schools today. This may not always be
the best way to consider the topic of problem solving in mathematics, but because
we would like to make this book as reader friendly as possible, we are providing
these signposts as an additional pathfinder for those readers bent on pursuing a
subject-matter focus. They should also prove useful for those unconvinced readers
Introduction
IX
Algebra
Arithmetic
Combinatorics
Games
General application
Geometry
Logic
Number theory
Probability
Statistics
Topology
Trigonometry
Figure 1
who, despite our efforts to highlight problem solving as a central theme, still want
to investigate mathematics by topic, or for those readers who are looking for clever
applications of problem solving for a particular topic in order to enhance their
instruction. The key to these signposts follows (see Figure 1).
The chapters capture a broad spectrum of ideas in the area of mathematics
problem solving. We begin with "Strategies for Problem Exploration," a truly
engaging "conversation" with the reader on how to recognize problems and how
to consider appropriate solution strategies. This chapter serves as a good introduc¬
tion to the book, since it covers most of the strategies listed above. This compre¬
hensive chapter is followed by "Unconventional Problem-Solving Strategies in
Mathematics Instruction," a chapter highlighting problem-solving strategies that
exhibit surprising and often elegant solutions, particularly where they may not
necessarily be expected. Through demonstration of their use in a somewhat ideal
application, it is intended that these problem-solving methods will become part of
X THE ART OF PROBLEM SOLVING
—Wolfgang Schulz
Berlin
Strategies for Problem Exploration
IRA EWEN
1. Working Backwards
2. Finding a Pattern
3. Adopting a Different Point of View
4. Solving a Simpler Analogous Problem
1
2 THE ART OF PROBLEM SOLVING
0) 1. Working Backwards
points, there are many routes west. Where should you start? Your experience tells
you that you should start at none of them. Since there are five routes from New
York City, each leading to a network of highways, but a single small destination
point, the problem cries out for working backward. To reach Urbana, we must
have been traveling west, so let's plot the return trip. We leave Urbana and travel
to the first major east-west highway. We take it eastward, seeking major interstate
highways that are also east-west, and continue working our way east, changing
highways as necessary, toward the New York City megalopolis. No matter how
you do this, if you are careful, you are led back through one of the five access routes
to New York City. You send the professor a map indicating the route you have
drawn, and you have simply and clearly found an acceptable automobile route
from New York City to Urbana. Note how natural this is. There was nothing
wayward or blundering about the procedure. Working backward was the natural
way to go.
There are several natural fifth rules specifying the objective of the game. Let
us examine them in turn.
Rule 5A: The player who takes the last counter wins.
Rule 5B: The player who takes the last counter loses.
This time it is neither clear how to start nor how to end. If I leave 0 counters
and have an even number, I will win; but I will also win when I have an even
number of counters if I leave one. A productive backward analysis would continue
with the question, "What is the minimal number of counters I could leave to
guarantee a win if I am holding an odd number of counters?" Reflection will show
that the answer to that question is 5. Let us see why. If I hold an odd number of
counters and leave 5, my opponent has only losing choices: If he takes 1,1 take 3;
if he takes 2,1 take 3; if he takes 3,1 take 1; and if he takes 4,1 take 1. In each case,
I reverse the parity of my holding from odd to even and leave him with a fatal 0
or 1 counter. To work backward from this end position, it is necessary to realize
that there are three parts to the winning end position: two consecutive numbers if
I hold an even number of counters and a single number if I hold an odd number
Strategies for Problem Exploration 5
odd 5,11,17, 23
All is clear. I must go first and take 2. Nothing else permits access to the
pattern of winning leaves.
This is a difficult analysis. Working backward made it possible, though not
easy, to complete the analysis.
The analysis is left to the reader. The reader should be aware that none of the
games discussed in this section is fair, because the player who goes first can always
force a win. If the first player errs, his opponent can now always force a wm. The
point of giving this game as an example of working backward is to foreshadow
the next section, finding a pattern.
Games are often useful because they teach important lessons with a light
touch. Surely the lessons to any prospective player of any game include:
• Don't expect to win against a player who knows more about the game than
you do.
• Don't bet against strangers.
TRIG
The student has been correctly but deceptively instructed that no proof of an
equation can begin with the equation to be proved.
This is correct. Suppose the student was asked to prove (the false) equation
2 = 3. The student reasons correctly: If 2 = 3, then 3 = 2 by the symmetry of equality;
if 2 = 3 and 3 = 2, then 5 = 5 because when equals are added to equals the sums are
equal; 5 = 5 is known to be true. The student reasons incorrectly that since he has
deduced a true equation from 3 = 2, he has proved that 3 = 2.
The instruction that forbade beginning with 3 = 2 was deceptive. If the
student looked for reversible consequences of 3 = 2 (reversible in the sense that
each step in the logical sequence is biconditional rather than a simple consequence
of the preceding step) and if the student reached a known equation, a proof
would have been achieved. Notice that the equation 5 = 5 cannot yield either 2 = 3
or 3 = 2 as a consequence; the student's argument is irreversible and proves nothing
useful.
When analyzing complex trigonometric identities or equations or in¬
equations, working backward is extraordinarily useful because we know where
we want to end but we don't know where to begin. The need to check reversibility
at each step remains.
Let us prove the identity given above. A student might say:
A valid proof of the required identity would begin with sin2 x = sin2 x and
conclude with the required identity whenever cos x ^ 0 and 1 - cos x 0.
Strategies for Problem Exploration 7
This method of proof is efficient, practical, and eminently valid. Note that
the biconditionality of each inference is psychologically helpful because the proof
looks like a series of inferences, which it is not. It would have been valid if each line
were an implication of the line below it; that is, if "if and only if" were replaced
by "if." We would not have a proof if "if and only if" were replaced by "only if."
The proof actually follows from the bottom up, not from the top down.
2. Finding a Pattern
Two children, ages 9 and 7, are traveling in a car with their parents. The father
says, "Moon is, sun isn't." The three passengers are accustomed to this bizarre
behavior and do not say what they are thinking. There is silence for a moment. The
father continues: "Knife and fork aren't, but spoon is." The whine of the wheels
and some labored breathing are the only audible sounds. "Autumn, winter, and
spring aren't, but summer is. Breakfast isn't, and neither is lunch, but dinner is.
Hammer is, but wrench, pliers, and saws are not. In fact, as you suspected, dear
children, brothers and sisters aren't but mommies and daddies are. What do you
think?"
Again, a relative silence reigned. Three minds had a thought that they did
not express. The older child was a delicious little girl and knew better than to say
anything about her father's sanity. "Are hot dogs?," she asked. "Oh, no," replied
the father, with a twinkle in his voice. The younger child, a boy often dazzled by
the three older members of his immediate family, asked, "What about pizza?" "Oh,
pizza is," said the father.
Many question and answers followed. The wife fell asleep, but the children
remained alert. Shortly before they pulled into a roadside restaurant for lunch, the
little girl said, "Home isn't, the library isn't, the supermarket isn't, and the flower
stand isn't either, but school is." It is sad to report that the father nearly lost control
of the car, so overpowering was his glee, which of course is.
8 THE ART OF PROBLEM SOLVING
Children can find patterns earlier in their lives than many people think.
Practice in finding those patterns helps to prepare them for more difficult explo¬
rations later in their lives. Read on.
It was late on a stormy Sunday night, and the five young men at a nameless
prestigious college were entertaining themselves as best they could in each other's
company. Puzzles and problems flew and became increasingly bizarre. Suddenly,
Igor's eyes flashed. "Have any of you heard about Jeanette and Giuseppe?"
Neither Bob, Chang, Mohammed, nor Ziggy had. Igor continued, "I am going
to utter a phrase, after which you may ask as many yes-or-no questions as you
wish, each of which I will answer. When you think you have solved the puzzle,
touch your right middle finger to your nose. I will then ask you five yes-or-no
questions, by which I will determine whether you have in verity solved the puzzle.
The phrase is: 'Jeanette and Giuseppe.' "
"Is that a phrase?" asked Mohammed.
"Yes," intoned Igor, in his best pedantic manner.
"He's playing the game already," said Bob.
"That is not a yes-or-no question," continued Igor.
The questions and answers followed quickly.
Question Answer
Is Jeanette French? No
Is Giuseppe Italian? No
Is Giuseppe a monk? No
Bob touched the middle finger of his right hand to his nose. Igor began the
fatal questioning.
Igor put the little finger of his left hand in his left ear. "This means that this
is a comment outside of the game. Your last answer transcended the game and was
right in a metaphysical sense." Igor removed his pinky finger from his ear. "You
have not solved the problem/' he intoned. "I will repeat the problem."
Igor repeated the entire spiel quoted before. The game proceeded, not for
hours but for days, interrupted only by the necessities of college life.
Several weeks later, Ziggy and Chang raced across the yard outside their
dorm and tackled Igor, throwing him on his back on the moist grass. He was not
at all surprised to see each of them touching the middle fingers of their respective
right hands to their respective noses.
No
Does 2 times 3 equal 6?
Yes
Does 1 plus 2 equal 9?
No
Are the uses of adversity sweet?
One thousand persons stood in a (large) circle. Each, starting with the
bald-headed woman, wore a sign on his or her back with a numeral from 1 to 1,000
NUMB
in a clockwise sequence. They began counting off. The bald-headed woman said,
"One, in," and remained in the circle. The 7'6" basketball player to her right (they
all faced out) said, "Two, out," and left the circle. The aesthetically repulsive Indie
Philology major (of undetermined gender) next in sequence said, "Three, in," and
remained in the circle. The cheerleader next in sequence (a lithe, attractive blond
male) said, "Four, out," and left the circle.
So it continued, with each person sporting an odd numeral stating the
numeral, saying "in," and remaining in the circle and with each person wearing
an even numeral stating the numeral, saying "out" and leaving the circle.
It is easy to visualize who was left when the count-off once again reached the
bald-headed woman. Since the person before her second turn had just said, "One
thousand, out," and left the circle, she now said, "One, in," and remained in the
circle. Continuing the alternating sequence, the Indie Philology major now said,
"Three, out," and left the circle. This time, only those persons whose number was
of the form An + 1, n a positive integer, remained in the circle, while those whose
number was of the form An + 3, n a positive integer, left the circle. As the count
approaches the bald-headed woman, we heard, "997, in," "999, out," followed by
her intonation, "One, in." The next statement, by our previously encountered
friend Igor (see section 2.2) was, "Five, out."
If this pattern of counting continues until exactly one person remains, you
are asked to predict the numeral on that person's back.
It is certainly possible to program a computer to give us the answer to this
problem, but we wouldn't gain any appreciation of the mathematics involved. You
are therefore forbidden to use electronic devices or, even worse, trial and error.
Any reader who writes the numerals from 1 to 1,000 in a large circle on a piece of
oaktag and attempts, perhaps with many colored pens, to simulate the count-off
process, is advised to close this book now and never reopen it.
How might we uncover the heart of this artichoke? In section four of this
chapter, the Josephus problem will be revisited. Let it remain temporarily a morsel
awaiting the lemon butter sauce of an ingenious solution. Read on.
An ancient prison was built in a long, long line, with numbered cells on one
side of a stone corridor. One night, the jailer imbibed too freely, and as each
prisoner slept, he unlocked every cell, walked back to cell 1, left it unlocked, but
locked cell 2 and each subsequent even-numbered cell. Back he walked to the
beginning, and proceeded as follows: He approached each third cell beginning
with cell 3, and changed its state, locking it if he found it unlocked (as he did with
cell 3) and unlocking it if he found it locked (as he did with cell 6). When he
Strategies for Problem Exploration 11
finished, he staggered back to the beginning and proceeded to change the state of
every fourth cell, beginning with cell 4. When he had completed that tour, he went
back to the beginning and repeated the process of changing the state of each fifth
cell, beginning with cell 5. Each time he reached the end of the corridor, whether
or not he was able to continue the pattern of that tour in which he was changing
the state of each nth cell, beginning with cell n, he returned to the beginning and
changed the state of each (n + l)sf cell, beginning with cell (n + 1). He concluded
this intoxicated operation by returning to cell 1, walking the entire corridor,
changing the state of the last cell, and collapsing.
Which prisoners were able to escape in the morning?
the most brilliant minds of each generation, as was the proof of Fermat's Last
Theorem. Developing skill at searching out patterns should be a major thrust of
education at home, in school, and on the job. The beauty of a game such as "Moon
Is, Sun Isn't" is that it models a form of pattern seeking that translates well into
classroom instruction.
Suppose, for example, a teacher wished to introduce the concept of a paral¬
lelogram. The teacher might well begin by providing students with a picture sheet
of perhaps 20 plane figures, some identified with the label is and the others
identified with the words is not. At the first level, students might be asked to
examine that page and create a page with 10 figures that they would label with the
word is and another page with 10 figures which they would label with the words
is not. On a third page, students would be asked to explain in their own words the
criteria they used to decide which of their figures "is" and which of their figures
"is not," and would then be asked to list, to the best of their ability, the clues from
the teacher's original page that led to their stated decision criteria. Students in
small groups would discuss the three pages and attempt to arrive at a simple and
clear statement of the decision criteria upon which they can agree. They might then
be asked to create an appropriate word for the figures listed under is. Many
students will come to a good definition for what we call a parallelogram and will
observe the need to prove or disprove that criteria established by different groups
are equivalent. „ . ,
Thus, for example, Group A might say that the figures listed as is had
opposite sides equal in length. Group B might say that the figures were those that
had opposite angles equal in measure. Improbably, but possibly. Group C might
say the "is" figures were those whose diagonals bisected each other. Probably one
or more groups would identify the "is" figures as those that had opposite sides
parallel. The teacher would facilitate the development of a sound definition by
12 THE ART OF PROBLEM SOLVING
introducing the need to replace the word figures by quadrilaterals and by suggesting
that a powerful line of investigation would be to see whether the conditions that
each group found could each be proved from others of the listed conditions. A
final outcome might be a powerful definition of parallelogram as a quadrilateral
that has any one of, and hence (as proved before the conclusion of the unit) all of,
the properties listed below:
In the game "Moon Is, Sun Isn't," a newly literate child might not write a
long composition but would do well to note, even if full spelling were not yet
within the child's capability, that each is word had a double letter (consecutive),
whereas, the isn’t words did not.
Jeanette and Giuseppe is a hidden-pattern game suitable for older children
and adults. It requires considerable sophistication to realize that the answers to
the questions have nothing to do with the content of these questions. The pattern
illustrated in Section B was simply that each question whose final word ended in
vowel was answered "yes"; each question whose final word ended in a consonant
was answered "no." The uncertainty related to the question in the section in which
the final word was given phonetically as "kat'l-og"' is that both catalog and
catalogue are acceptable spellings in English. This game develops flexibility in
pattern-seeking mindsets by giving students the confidence to act on seeming
contradictions. When the answers to their questions seem to be inconsistent, they
must move beyond the urge to give up to the urge to look at the problem from an
alternative point of view (a strategy we explore more fully in the next section).
The Josephus problem will not be discussed in this section, because it yields to
a combination of problem-exploration strategies. It will be analyzed in section 4.2.
The Long Cell Block problem is suitable for students cognizant of the possi¬
bility of factoring integers uniquely (except for order) into primes. By making
factorization tables after coming to the realization that the state of a cell is changed
once for every factor including one, they will see that for most positive integers,
factors come in distinct pairs. Thus, for 6, the pairs are respectively (1, 6) and (2, 3).
Cell 6 will be opened for 1, closed for 2, opened for 3, and closed for 6; in the
morning, it will be locked. Some students will come to the exciting discovery that
the only positive integers with an odd number of factors are those in which one of
the factors is paired with itself in the factorization, so that it leads to a single change
of state. Clearly, these special positive integers are the perfect squares. Thus, for
example, the pairings for 9 are (1,9) and (3,3). The pair (1,9) leads to the expected
two changes, but the pair (3, 3) leads to only one change. Thus, cells 1, 4, 9,16, 25,
and indeed all cells in the block of the form n2, n a positive integer, will be open in
the morning. Most people learn best by constructing their own knowledge. The
samples given here model that constructivist philosophy using words, number
patterns, and, in the case of the described activity with parallelograms, pictures.
There is no need to give students a formal definition of pattern or finding a pattern.
Strategies for Problem Exploration 13
With enough suitable practice, almost every student will come to preverbal
understanding, and once that preverbal understanding has been reached, the
student will grow in the ability to apply the problem-exploration strategy of
finding patterns.
Ann did what she had promised. To her practiced eyed, the two pans were
perfectly even and justice was guaranteed. Gluttonous Suzie began to whine in
her accustomed way, "Johnny has more than I do."
"We haven't said which dish is Johnny's and which is yours, dear," Ann
offered.
"It doesn't matter," screamed Suzie, "whichever one you give him, he'll have
more than I have."
Counselor Bob had been standing by silently. "Fairness," said Bob pedanti¬
cally, "is psychological, not scientific. You do everything too scientifically, Ann,
and I should know. Dear children, let me tell you what to do: Let's dump both
plates back in the bowl. Suzie, you divide the pudding in a way you think is fair.
Johnny, you select which plate you want after Suzie finishes. You see, Suzie, if you
divide the pudding fairly, whichever one Johnny picks will be OK. And Johnny,
you pick whichever you think is more."
"That's no good," Johnny shouted, "I want to divide and let Suzie pick."
The problem explorer, contemplating the ill-structured situation just de¬
scribed, recognizes a multitude of implicit problems. Perhaps the explorer decides
that the real problem is to divert the children from the impending war. King
Solomon might say, "Divvy it up any way you want, and when you're finished, if
I don't hear any screaming or fighting, you can each have a banana split with three
scoops of ice cream and ample hot fudge sauce."
A nearby mathematics professor, Carl Friedrich, might abstain from any
contact with the people involved and pursue the following thinking: "I should be
able to generalize this problem to n persons, using the scientific approach with no
difficulty. How can I generalize Counselor Bob's psychological approach to three
people, let alone to n?"
Fairness is a deep concept, whether it is viewed mathematically, psychologi¬
cally, socially, politically, or humorously. Wonderful and rich interdisciplinary
instruction in schools, from kindergarten through graduate school, could center
on fairness. Redefining fairness from a scientific to a psychological context is an
insightful shift that might elude many closed-minded, closed-ended problem
solvers.
ARITH 3.3. A Talmudic Tale: The Men, the Loaves, and the Gold
GEN
On a stetl farm in Poland many years ago, three men were working in the
field as the time for the midday meal drew near. One of the men, Ezekiel, was the
wealthy owner of the field who believed that by working one day each year with
the men who lived in poverty on his land, he would achieve the Lord's blessing
and the men's loyalty. He noted that he had 10 gold coins in his pocket, but that
the workers had brought loaves of bread. Abraham had three loaves and Isaac
had two.
Strategies for Problem Exploration 15
"I have a splendid idea," Ezekiel proudly declared. "Let us share the five
loaves you have between you equally among us, and in return I will give the 10
golden coins in my pocket to share between the two of you." Abraham and Isaac
were overjoyed, recognizing that a single gold coin was nearly a year's income.
The three men shared the loaves. After eating, Ezekiel announced that he had
grown fatigued and would cease work for the day, and that his men could also
quit work early at sundown.
As soon as Ezekiel left, Abraham and Isaac began to argue. Abraham said, "I
had three loaves and you had two. I should get six gold coins and you should get
four, right?"
"No, no," said Isaac. "We shared the bread equally, we should share the gold
equally, five coins apiece."
Long before sundown, the two men decided to refer their dispute to their
rabbi, who was noted not only for his talmudic wisdom but for his mathematical
acumen.
"So, Rabbi Jacob," Abraham concluded after recounting the events of the late
morning, "what should we do?"
A furtive tear appeared in the comer of Rabbi Jacob's eye. "Dear friends," he
said, "you present me with a terrible problem. If on the one hand I advise you to
share the coins as Abraham wants, you, Isaac, will be a trifle aggravated, but you
will soon concur and you both will leave me to my studies. If I counsel you to
divide the coins five and five, you, Abraham, will be annoyed, but soon you will
concur, and, again, both of you will leave me to my studies. On the third hand,
unfortunately, if I tell you the correct solution, both of you will argue with me until
dawn, I will not get back to my studies for a day, and neither of you will be fit for
work tomorrow."
In chorus, Abraham and Isaac shouted, "What do you mean. Rabbi Jacob?
Surely, there are no other possibilities."
Rabbi Jacob, who had foreseen section 3.2 about pudding apportionment
those many centuries before it was written, was faced with a quandary. He was
obligated to tell the truth by his own vision of rightness and rabbinical responsi¬
bility. He knew that this would not go well.
"My friends, I ask you to consider the reason that money is paid for goods.
The gold is intended to compensate the seller for what he has given to the buyer
in goods and in service. In this case, we need only consider goods. Abraham, you
and Isaac each ate one and two thirds loaves of bread, as did Reb Ezekiel.
Therefore, you, Abraham, gave away one and one third loaves of bread to your
honored landlord; whereas, you, Isaac, gave away only one third of a loaf of bread.
Since you Abraham, gave away four times as much bread as Isaac, you should get
four times the compensation. The fair distribution of the 10 gold coins is 8 coins to
Abraham and 2 coins to Isaac.
The screaming continued all through the night, and Rabbi Jacob did not get
back to his studies for several days.
16 THE ART OF PROBLEM SOLVING
"I get most of it, Mrs. Thom," Jacqueline said. "When you add two positive
numbers, I understand why you add the values and affix a positive sign. When
ARITH
you add two negative numbers, I understand why you add their values and affix
a negative sign. I begin to get confused when you talk about problems like 'positive
seven plus negative three' as adding because, to me, it's subtracting. In fact,
'positive three plus negative seven' is a little worse. But I can get all the answers
you want to all of those problems, even though I'm not sure what those answers
mean. I really get into a lot of trouble with subtraction.
"When you give us 'positive seven minus positive three,' I get positive four,
just like you want, but I have a lot of trouble when the second number is negative.
For example, 'positive three minus negative four' makes no sense, no matter
whether I give you your answer of positive seven or my answer of negative one.
Y'know, Mrs. Thom, I never even understood 'one minus three' back in grade
school. If you put a penny on the table and tell me to take three pennies away, I
can't owe the table two pennies. I really think that 'one minus three' has no
answer."
Mrs. Thom considered herself to be an exemplary teacher, cognizant of all
the most modem methodologies, doctrines, and insights related to her noble
calling. She took out a box of manipulatives and told Jacqueline to work with them
until she understood perfectly all the concepts involved. Jacqueline knew exactly
what to do. She puttered with the junk for 11 minutes and then announced joy¬
ously, "I get it, Mrs. Thom." Since Jacqueline was quite proficient at memorizing
rules, she knew that she could get high grades on all of Mrs. Thom's tests and that
her wonderful work with the manipulatives would earn her extra points on the
report card.
"Nobody really understands this stuff," Jacqueline thought, "but then, no¬
body cares about mathematics anyway after you get out of school."
would now mean, "What number measures the change needed to get from
the point marked negative four to the point marked positive three on the
number line?" I see you'd have to go seven units to the right, so the answer
must be positive seven.
The above sequence is complete and hence finite. The identity of the missing term
and the explanation for the sequence are wonderful to contemplate. Were I to give
an explanatory title to the puzzle, that title would be "Oh, So Sweet."
Two of the problems in this section dealt with fairness. The extension of the
pudding sharing in a psychologically fair way to three gluttonous children can be
done as follows. Let us call the children Dick, Jane, and Spot. The order in which
the action to be described is taken is determined by flipping a fair three-sided coin
or, if none is easily available, tossing a die, with Dick going first if ace or deuce
comes up, Jane going first if trey or four is on top, and Spot getting first lick if five
or six is uppermost.
We set a context in which Dick goes first. Dick creates a portion of pudding
that he thinks is fair—that is, a portion he would gladly accept.
Jane and Spot proceed sequentially. If Jane believes that Dick's portion is a
fair share or less, she passes, and it is Spot's turn to decide. If Jane believes that the
portion Dick created was too large, she stakes a claim by returning some pudding
to the big bowl and saying that she would take the reduced portion as her own.
Strategies for Problem Exploration 19
Ostensibly, this would be acceptable to Dick, since Jane is taking a portion smaller
than one which he considered fair. It is now Spot's turn. Spot can either agree that
the portion Jane is holding is to be hers or may stake a final claim to it by removing
some of the pudding and returning the removed pudding to the big bowl, making
the portion Spot's and Spot's alone. Once any one of the three gluttons has been
awarded a portion by this procedure, the remaining two partitioned the pudding
in the big bowl by the original two-person technique: "one divides, the other
chooses." (Preferable designation would be "one apportions, the other selects,"
because divide and apportion should not be used synonymously in mathematics
class. To see why, think about dividing six into two parts as contrasted to appor¬
tioning six marbles among two students.) The generalization of the psychologi¬
cally fair apportionment technique to four or more gluttons is left to the reader. I
hope the reader thinks this is fair.
The advantage of finding a pattern while exploring a problem situation is
that much broader predictions can be made about the structure in which a pattern
exits. In extreme conditions, the exact nature of an unfamiliar structure is revealed.
Consider the finite sequence mentioned above. If the listed numerals are viewed
conventionally, little is discernible. Suppose, on the other hand, that a practiced
problem explorer viewed the 15 numerals (actually, 14 numerals and a blank) as
each representing the same number. Perhaps it would be a deft leap to the vista
from which can be seen that the number being named is 16 (oh, how sweet) and
that the bases of numeration are, in order, 16,15,14,13,12,11,10, 9, 8, 7, 6,5,4,3,
2. It is undoubtedly unnecessary to inform the reader that the numeral to be
inserted in the blank is 31.
The description of a discovered pattern may take several forms. If the pattern
is a sequence or a linear geometric pattern, there are two simple ways to describe
the pattern: explicitly or recursively. An explicit description of a sequential pat¬
tern is a formula or complete description that gives the term or figure in n* position ALG
as a function of n. Thus, for example, an explicit description of the arithmetic
NUMB
sequence { an } with initial term al = 3 and with common difference d = 4 is an =
3 + (n _ i) x 4. This means that the seventh term in the sequence is3 + (7-l)x4 = GEN
if the circle were moved to the left or right, the circle would be tangent to the upper
and lower bases of the square. A recursive description of the same linear geometric
pattern would be that the pattern consists solely of one-by-one squares and
diameter-one circles oriented so that the sides of the square are parallel or perpen¬
dicular to the pattern and that the circle, if moved into an adjacent square, would
be tangent to the upper and lower bases and such that an is a circle if and only if
an_1 is a square. In both cases, the sequence and the linear pattern, the recursive
definitions just given are incomplete, because we were not told how to begin. We
must append to the recursive definition of the numerical sequence that a1 = 3 and
to the recursive definition of the geometric sequence that al is a square.
Computers can handle both recursive and explicit definitions of patterns
effectively. Mathematically, we can pose fascinating problems by giving an ex¬
plicit description of a sequence and asking for a recursive one, or giving a recursive
description of a sequence and asking for an explicit one. Thus, for example, the
classical Fibonacci sequence may be defined recursively by ax - a2 -1 and, for
n > 3, an = an _ x + an _ 2. Given that definition, it is an interesting challenge to deduce
an explicit description. It is interesting to speculate whether the recursive or the
explicit description of the Fibonacci sequence gives greater insight into its nature
and when each is preferable for further investigation of the sequence.
Both recursive and explicit descriptions are possible for higher-dimensional
numerical and geometric patterns. Consider whether one would be more likely to
obtain the designer's view of a large tesselated piazza if one were to present the
artisan with explicit instructions for which tile to place on each coordinate of a
design grid or by giving the artisan a rectangular block containing the complete
design with the instruction to place that block in the upper left-hand comer and
to duplicate exactly that design horizontally and vertically throughout the (pre¬
sumably) rectangular piazza, with no interstices and with parallel orientation of
all edges.
Let us return to the thrilling lines of yestersection 2.3. Polly Sue Seidlitz
had encountered the Josephus problem in her sixth-grade mathematics problem-
exploration club. She had been thinking about it for several days but could not
bring herself to write the numerals from 1 to 1,000 on a piece of paper. NUMB
"Teachers are so inconsiderate," she thought. "If Mr. Sperdel had made it 10
people instead of 1,000,1 could have done it."
Polly Sue hoped she was not deceiving herself. She wrote the numerals from
1 to 10 in an appropriate circle on the inviting blank sheet before her. She picked
up her red pencil and crossed out 2, 4, 6, 8, and 10. She put down the red pencil
and picked up the blue one, thinking, "1 in, 3 out, 5 in, 7 out, 9 in." She crossed out
3 and 7 in blue. It was time for the green pencil. "One out, 5 in, 9 out" resulted in
appropriate crossouts. One numeral remained. "I could have solved it if he had
said 10! The answer is 5." She stared at her work for a long time.
A spark within her spirit burst into flame. "I'll bet that if I solve the problem
for enough small numbers, I'll see a pattern."
Polly Sue took another blank sheet on which she began the following chart:
1 1
2 1
3 3
4 1
5 3
6 5
7 7
Sally stared at the chart. "I'm tempted to write '8,9' on the next line, but there
is no number 9 when there are only 8 people, so what sense would that make? I'll
bet the answer turns out to be 1 again." A little colored-pencil play verified her
conjecture.
22 THE ART OF PROBLEM SOLVING
Polly Sue contemplated the chart. "I'll bet the sequence of odd numbers starts
anew. After all, it would figure, knowing that the answer for 10 is 5, that the answer
for 9 would be 3." More colored-pencil play resulted in the following chart:
1 1 11 7
2 1 12 9
3 3 13 11
4 1 14 13
5 3 15 15
6 5 16 1
7 7 17 3
8 1 18 5
9 3 19 7
10 5 20 9
Polly Sue went to the bookcase and took down her father's copy of the book
you now hold in your hand. She opened to section 3.6 and read it with astounding
comprehension for one so chronologically young. "I can continue this chart recur¬
sively with no trouble. I wonder if I can get an explicit formulation that will let me
express the right-hand term on any row, knowing the left-hand term.
Polly Sue re-examined the rows in which the right hand term was 1. "What
would Daddy say ... I mean, what could I say is common to 1, 2, 4, 8, and 16?" It
was crystal clear. "Powers of two!" Polly yelled with the enthusiasm with which
many of the boys in her class yelled "Boom" when they threw water bags at her.
She filled in her chart up to 32, following the indicated recursive pattern and
ending with:
N.C. N.B.L.R.P.
30 29
31 31
32 1
Strategies for Problem Exploration 23
"I can solve this problem. Let me write the consecutive powers of 2: 1, 2, 4,
8, 16, 32, 64, 128, 256, 512.. . mustn't pass 1,000. Since 1,000 = 512 + 488, the
numeral on the last guy's back will be 2(488) + 1 or 977. Boy, I never would have
guessed that."
Mr. Sperling had had a very difficult morning. He had been issued a speeding
ticket on the way to school and was brooding about it, and his fifth-grade class
was treating him gingerly; they recognized storm signals.
"Today we are going to learn how to add fractions," barked Mr. Sperling.
"Try to add 3/17 and 6/13." There was dead silence. Even Jerry, one of the very
best math students Mr. Sperling had ever taught, stared at the blackboard with
puzzlement. Mr. Sperling already regretted having taken out his traffic ticket on
the class, but he knew no good way to get out of the bind.
Jerry wrote something on his paper, studied it for a few minutes, and then
raised his hand.
"Mr. Sperling, those numbers are too big for me. I always try to learn
something with smaller numbers because if get it with the small numbers, I usually
can get it with the big ones. I tried 'a half plus a half' and thought like this: One of
anything added to one of the same thing gives two of that thing. So a half plus a
half equals two halves and two halves is one whole. I'm not sure why, but I don't
feel good about saying it that way. You [Mr. Sperling] never say 'One half plus one
half equals one unit,' but somehow I see 'unit' as an appropriate abstraction and
I don't see 'whole' that way." Jerry fidgeted a little more. "I really don't have it yet
because thinking through a half plus a half just doesn t do it. It s a bit scary, but
I've got to think about a half plus a third. Can I go over into the math comer for a
while?"
A glimmer of a glimmer of a glimmer of a perception came to life m Mr.
Sperling's gray cell morass. There was a spiritual bond between Mr. Sperling and
24 THE ART OF PROBLEM SOLVING
a German teacher from a couple of centuries ago who respected the thinking of an
odd child in his class who found an unusual way of adding the numbers from 1 to
100 mentally. Unarticulated but present was the thought that by nurturing this
Jerry-thing he might one day be celebrated.
"Do you see what wonderful thing Jerry is doing?" said Mr. Sperling to a
class of students who were too stunned by what was happening to do anything
except stare with glazed eyes at a fixed point at infinity that they had never
previously contemplated. "I purposely gave you large numbers to add, hoping
that you would learn the wonderful strategy of simplifying. Many times you can
gain deep understanding of a problem situation you encounter by addressing a
simpler, analogous problem situation."
"Ann who?" asked Betty, Mr. Sperling's front row orchid.
Ignoring her honest inquiry in pursuit of knowledge, Mr. Sperling continued,
"Jerry saw that the tops of fraction are called numerators because they tell us how
much or how many (the number) we have of the unit named in the bottom, which
we call the denominator (the name). Jerry saw that thinking about one half plus one
half was an analogy (remember our lessons on analogy, Betty?) to a problem such
as adding one dog to one dog and getting two dogs. Jerry is off in the math comer
trying to figure out what one dog plus one cat equals."
Jerry started to laugh. He spoke without raising his hand or gaining the
traditional license to speak. "I think I understand what adding fractions means,
Mr. Sperling. You know, you have been wrong every time you told us that you
can't add apples and oranges. Of course we can add apples and oranges. One apple
plus one orange equals two pieces of fmit. One dog plus one cat equals two animals
or two mammals or two living things. All we need do is find a unit which takes in
both cats and dogs or, in the other case, both apples and oranges. So I said to myself,
'Hey Jerry, what takes in both halves and thirds?' So I said, 'Certainly not fourths
or fifths, but sixths, twelfths, or twenty-fourths would be just fine.' Well, Mr.
Sperling, it's really very easy. One half is three sixths and one third is two sixths,
and I have reduced the problem of adding halves and thirds to the simpler and
previously solved problem of adding two numerical expressions with identical
units. Two of anything plus three of the same thing is five of that thing."
There was a gentle knock on the door. Polly Sue Seidlitz entered the room
with a note from Mr. Grimmas, the principal, to Mr. Sperling about the traffic
officer waiting in the general office. As Mr. Sperling read the note and became
appropriately ashen, Polly Sue glanced at the blackboard and saw the original
problem of adding 3/17 and 6/13. "How can fifth grade be doing that when our
sixth-grade Bluebird class never does things this hard?"
"Take over the class, Polly Sue, I've got to go to the principal's office," whined
Mr. Sperling as he scooted out the door.
Polly Sue had waited all her life for the star to break his leg. "What's your
thinking about this, class?" she asked, without knowing what "this" was. "Are you
Polly Sue Seidlitz?" asked Jerry, with far more respect than he had ever accorded
Mr. Sperling. "I've heard about you. I can explain the problem on the board."
Strategies for Problem Exploration 25
She glanced at Mr. Sperling's oaktag with the kids' names on it. No one sat
in the comer of the room with the polygons hanging on the wall. "What's your
name, little boy?" she asked in as kindly a manner as such a child could manage.
"Jerry McBourbak," he answered.
"Well, Jerry, tell us about your thinking."
"In order to analyze this unfamiliar problem with big numbers, I first looked
at a simpler problem like it, namely a half plus a third. It was a question of getting
a common unit for halves and thirds, which could be sixths, twelfths, twenty-
fourths, or so on."
"Or eighteenths," said Polly Sue.
"Hey, you're right!" cried Jerry, with genuine enthusiasm. "I missed that."
"Actually, so did I the first time," admitted Polly Sue. "Go on with your
thinking."
"Well, with seventeenths and thirteenths the arithmetic was little harder, but
the idea was exactly the same. I multiplied 17 by 13 in my head, using the 15 times
15 minus 4 model—I'm sure you know what I'm talking about, Polly Sue—and
since I memorized all perfect squares from 1 x 1 to 31 x 31,1 knew that 15 x 15 is
225, so 17 x 13 is 221. So the right unit for the answer could be 221sts or other stuff
I'm not going to bother to figure out. So 3/17 is 39/221 and 6/13 is 102/221.
Clearly, 3/17 plus 6/13 is 141/221. Apples and oranges," Jerry concluded.
"One man, two automobile tires, and the four laws of Kepler," said Polly Sue,
smiling at a joke that was not a joke.
Many years later, as Jerry and Polly Sue lolled on the beach in Waikiki on
their honeymoon, Polly Sue said, "We really should resolve emerging problems
in our relationship before they fester. I never fully internalized the analogy
between adding halves and thirds by combining the new unit sixths, changing the
count of halves and thirds to conform to the new unit, to adding one apple to one
orange and getting two pieces of fruit without having to change the count."
Jerry held her tenderly and explained with no hint of patronization what he
had realized in a flashback in the math comer so long before. "Sometimes the count
will change and sometimes it will not when you change units. To mail a letter
requiring $1.20 postage, the current going rate for a one-ounce local letter" [after
all, a couple of years have gone by] "you can paste on one $1.20 foil stamp, three
40-cent disks, six 20-cent rectangular traditional, or 120 green antiques. Here, the
count changes. On the other hand, one dog is one animal is one mammal is millions
of chains of DNA bound in a complex pattern. Perhaps if we added one apple plus
one orange plus 17 cells of a pear, the analogy would be clearer."
"Again and again you show me why I love you," said Polly Sue, author of the
recent smash best-seller Infinite Dimensional Fractal Manifolds and Their Measures.
Polly Sue Seidlitz and Jerry had taken to doing their homework together. He
often let her carry his books, because they both agreed that she was physically
26 THE ART OF PROBLEM SOLVING
much stronger. They were going over to her house that afternoon because they had a
particularly delicious set of common assignments for their problem-exploration
club at Gallifrey High School. "I think the Doctor is one of the best teachers I've
ever had. I got a kick out of his asking me to find the center of a 6-7-8 triangle."
Polly Sue was exuberant.
"There are so many ways to think of center, even in a triangle," Jerry noted.
"When I was five years old. Mom bought me a compass."
"You've got to learn to be more precise," Polly Sue exhorted. "I'm not even
sure what you mean; something to find the magnetic north pole, or something to
draw circles?"
"Sorry. She bought me a pair of compasses. I tried drawing circles around
everything I could find and then I tried putting circles inside of them. I found it
hard to get a single circle to intersect the vertices of a given triangle and even
harder to get a single circle internally tangent to the sides or sides extended of a
triangle."
"So which did you decide was the center of the triangle? The center of the
inscribed circle, or the center of the circumscribed circle?"
"Neither, both, and more," Jerry responded. "After I found out how to
circumscribe a circle around a triangle, I cut cardboard triangles out of my father's
shirt cardboards and tried to balance them on a pin at the center of the circle. When
that usually didn't work, I tried the same gig with the center of the inscribed circle.
It took a while for me to find out about medians and centroids."
"Now that you know how to find the centroid of a triangle, do you think
that's the true center?"
"Center is as center does," said Jerry.
"Well," Polly Sue said, "even as the Doctor was asking us to think about what
we meant by the center of a triangle, I was way ahead of him for a change. I asked
myself how one would find the centroid of an n-sided convex polygon. Have you
ever thought about that?"
"Often, when I'm alone. I actually think that the Doctor gave the problem to
you in a weaker pedagogical sequence than he might have. Suppose he had asked
you how to find the centroid of an n-sided convex polygon: Consider what you
would have had to do!"
"I get it! We all would have had to think about solving the simplest case of
the problem, namely, finding the centroid of a triangle. We would all have learned
something about problem exploration in addition to learning about the problem
itself."
"Exactly. After you had simplified your general question to one about tri¬
angles and had learned all about the centroids of triangles, you would have pro¬
ceeded naturally to convex quadrilaterals. When you draw a diagonal in one of
those suckers, you get two triangles and can easily determine the centroid of each.
It follows that the centroid of the quadrilateral is on the line segment joining them."
"That's obvious, but where?"
Strategies for Problem Exploration 27
"Clearly, not the midpoint, at least not in general. The only time I'd expect it
to be the midpoint would be when the two triangles were equal in area. That
certainly doesn't happen most of the time."
"Sowhatchagonnado?"
"What we always do."
"What's that?"
"You know," Jerry said, with the innocence of a young mathematician.
"Draw the other diagonal."
"Oh, where the analogous segment meets the one you drew," Polly Sue ex¬
claimed. "That's great. Let's try to generalize that to convex pentagons. Looks easy:
Draw a diagonal and you have a quadrilateral and a triangle. We can find the
centroid of each and therefore have a line segment which must contain the centroid
of the pentagon."
They had reached the steps leading to her front porch and were both quite
excited. "Then use another diagonal and it works out just as with the quadrilateral.
That generalizes to any convex polygon by induction."
"Naturally," said Polly Sue, inserting her key into the waiting keyhole and
proceeding to unlock the door.
"Sometime let's look at other kinds of centers for polygons when they are
defined."
4.5. Discussion
This is not to suggest that proper teaching proceed from the simple to the
complex. Such teaching may result in temporarily faster integration of new ideas
by students but may well not be the best way for them to learn techniques for
profitable and imaginative exploration of ill-structured problem situations. Teach¬
ers should often pose structured problems in complex settings when they are
teaching students how to find simpler analogous problems. This technique is not
always apparent to students; it can and should be taught. An excellent paradigm
for instructing in this area is this: Proceed from the complex to the simple to the
complex. Ideally, students should suggest that this be done even when teachers
do not structure instruction in this way.
Of course, if real-world settings of problems are given precisely for the reason
of teaching students to proceed from the complex to the simple to the complex,
and methodologies for doing this have been thought through, a major objection to
using the real world to teach mathematics is removed, whereas the benefits
remain.
"What was the big deal about that?" asked Chang, looking up from the
cryptic puzzle at the back of the current issue of Harper's, which he had picked up
early in the conversation.
"I was only in seventh grade and I was having a bit of a problem thinking
about the general polygon, whether or not it was regular. I was standing by my
desk because I didn't want my father to see how comfortably I could sit. 'How do
you draw one?' I asked him."
"You'd have to draw something open, with dots, but is that the way you want
to go? How would you pick up a snake?"
" 'Quick and hard. I'd grab it below the mouth and at the tail,' I told him."
" 'But not in the middle,' he said, fixing me in his intense stare."
Each of the young men had become thoroughly engaged in the problem
Mohammed was describing. "He was telling you to look at the extremes!" shouted
Igor. "That's great."
Bob looked puzzled. "So, the equilateral triangle is just under the head, but
there's no tail."
"Sure there is," Chang said. "The infinite-sided regular polygon. Also known
as ..."
"A circle!" screeched Igor.
"Right," Mohammed resumed. "We'd learned that the area enclosed by a
circle was nr2 and that the area enclosed equilateral triangle is (s2V3)/4. Unfortu¬
nately, that didn't lead me anywhere. I fooled around with it for a while and
something hit me: Circumference was analogous to perimeter. 'Holy moley! I
shouted. The circumference of a circle is 2nr, and its radius is r. Suppose I multiply
them and then divide by 2."
Chang was grinning. "Perimeter times radius divided by two. I'll bet you
tried the same thing with the equilateral triangle. The perimeter is 3s, but what did
you take for the analogy to the radius?"
"Well," Igor mused, "there's no problem defining the center of an equilateral
triangle, so I guess you would want the length of the line segment joining the center
to an appropriate point on the boundary."
"A vertex?" asked Bob.
"Hardly!" said Mohammed, with just a tinge of patronization. "How does a
radius intersect a circle?"
"Of course," Chang explained, "it intersects the circle orthogonally, so you
need the line segment joining the center of the equilateral triangle to the side which
meets the side orthogonally."
"And the length of the segment of the altitude joining the center of an equi¬
lateral triangle to a side is one third the length of the altitude, or 1/3 x (sV3)/2. So
I multiplied the perimeter by the radius-analogy and divided by two!"
"Getting (s2V3)/4," said Bob, with dawning comprehension.
"So I figured, if it worked at the head and it worked at the tail, it might very
well work for the whole body. I conjectured that the area enclosed by a regular
polygon was always perimeter times radius-analogy divided by two."
30 THE ART OF PROBLEM SOLVING
"So that's why the apothem is important. It's the radius-analogy in any
regular polygon/' said Bob. "I think this is the first time I really understood that
formula."
Chang smiled. "There is another point that should be considered. Would this
procedure still work if we were talking about perimeter instead of about area?
Have any of you seen that old chestnut in calculus where a student who had just
learned integration attempted to approximate the length of the line segment from
the origin to the point (1,1) by constructing 'staircase approximations'? Each step
consisted of a vertical segment with one endpoint on the diagonal segment and a
connecting horizontal segment from the top of the vertical segment back to the
diagonal. The students then let the length of each vertical and horizontal segment
approach zero and claimed that the limit of the sums of vertical and horizontal
segments would be the length of the diagonal. Of course, the length of the diagonal
is V2~ and the sum of the lengths of the vertical and horizontal segments is 2, no
matter how small you make them."
"Wait a minute," interrupted Mohammed. "Length approximations must be
made with chords! Both ends of the chord must be on the curve or line segment
being approximated. That's not the case in your example."
"Exactly," said Chang. "And if we approximated the arc length—that is, the
circumference of a circle with chords—we wouldn't have a problem. Since the
length of a side of a regular n-sided polygon is 2r sin (n/n) where r is the length
of the segment joining its center to a vertex (the central angle formed by two 'radii'
and a side has measure 2k/n so that half that angle has measure n/n, and it follows
by right triangle trigonometry that sin n/n = (s/2)/r - s/2r). Thus, the perimeter
is n x 2r sin (n/n). Those students who have learned that lim [(sin x)/x] = 1 as
x —» 0 would let x = n/n and see that as n —» x —> 0. Thus as n —» », ns = 2rn
sin (n/n) - [2rn sin (n/n)]/(n/n) —> 2rn x 1 = 2rn, as we would wish. If you know
the ideas involved, it all works."
"I wish all of life were like that," said Mohammed wistfully.
Big Ed the bookie had had a profitable day. The favorites all had lost and he
had given short odds on the underdogs. He double-locked the metal doors and
GEOM bolted the reinforced shutters.
"I've done enough arithmetic; time to relax with mathematics." Big Ed
opened his well-worn copy of George Polya's Mathematics and Plausible Reasoning.
He was working on the problem of finding the angle of maximum measure formed
by the endpoints of a line segment and a variable point P moving on a noninter¬
secting line l in the same plane. His first conjecture had been that the point P
yielding the angle of maximum measure was the projection of the midpoint of
AB onto t. His flush of pride was utterly dispelled by a healthy dose of reality.
Strategies for Problem Exploration 31
"It can't be," he censured himself. "Winner and straggler." He sketched two
diagrams on his notepad. In one, AB was parallel to l; in the second, AB was
perpendicular to l without intersecting it. "It works in the parallel case, but it's as
bad as it can be in the perpendicular case. If it ain't true for winners and for
stragglers, it sure can't be true for all horses.
"Well, as a start, they all got four legs and a body holding them together.
How can I tie the line segment to the line in a solid way that's the same for the
parallel case and the perpendicular case?"
Ed stared at the page for a very long time. "I sure ain't getting to the winner's
circle on this one."
Big Ed's occupation once again served a scholarly purpose. "Circle," he
screamed, smashing his often-broken knuckle on the oak table he had inherited
from his father, who had met an early unnatural death. Ed drew circles (or slightly
distorted approximations thereto) containing the line segments as chords and
resting on the line. "I'll bet those are the points," Ed said. He took out his red pencil
and drew a second circle, symmetrically placed, in the case of the perpendicular
segment. "And not necessarily a oner. Winners don't have to be oners." He thought
that his black-sheep Uncle Ted, who had forsaken family tradition to become a
mathematics professor, wouldn't be satisfied with talking about a circle with its
feet on the ground. Big Ed attempted to mimic Ted s Boston accent. The angle of
maximum measure is formed with vertex at point P where a circle having the line
segment as a chord and tangent to the line intersects the line."
"Put can you prove it?" Ted chided from within Big Ed s psyche.
"Sure," replied Big Ed, yielding to schizophrenic tendencies. "When the
point of tangency is the vertex of the angle, it's measured by half its intercepted
arc. For any other point, the angle formed is measured by half the difference
between the measure of that same arc and some other arc of the circle, so it's
smaller."
"Good lad!" said Ted. "But how can you be sure such a circle always exists?"
Big Ed realized immediately that he should think about a way to construct a
circle having a given chord and tangent to a given line in the same plane which
did not intersect the chord.
"Easy in the parallel case, so let's look at the straggler." Ed figured that if he
could do the construction in the perpendicular case as well as the parallel case, he
would be able to work out a construction for the general case.
"I can see," Big Ed continued in his odd discourse, that if we examine the
line containing AB, it meets fat point Y. The circle^seek passes through A and B
and is tangent to tat point P to be determined. So YBA is a secant of that circle, and
yp ^ a tangent." Big Ed remembered his high school plane geometry class, which
he loved more than any other class he had taken in school but which had been
replaced in recent years by some atrocity containing algebra, geometry, and
several other uninteresting topics. "From a point external to a circle in the plane
of the circle, a tangent segment is the mean proportional between the entire secant
32 THE ART OF PROBLEM SOLVING
segment and its external subsegment. It's pretty easy to construct a mean propor¬
tional using a segment of length YB + YA as the hypotenuse of a right triangle with
the vertex of the right angle at a point at distance YB from the endpoint of the
hypotenuse. Once I have the length of the mean proportional, I can find the
location of P immediately, giving me the angle APB of greatest measure."
"Ed, your language does not conform to modem considerations of linguistic
precision," Ted chastised. "You should be careful to say 'the length of the tangent
segment...' "
"Shut up, Ted! You talk so much that nobody listening to you can conceptu¬
alize the content of what you say at length." Ed chortled at his own play on words.
"You talk your way; I'll talk mine."
"Your tirade is an ignoble attempt to cloud your own confusion. You are able
to construct point P in the extreme cases; you have learned little about what to do
in the general case."
"Then, you're a pretty bad teacher, Ted. Didn't you always say that we
should look at extreme cases to see if a pattern is established that lets us know
about the general case?"
"The key word, Ed, is if. In this case, examination of the extreme cases tells
you very little about the general case, except for the very beautiful insight that one
of the two extremes immediately generalizes. That's often the case, Ed. Even when
the extremes do not give you a complete insight into the proof or pattern you seek,
you have not wasted your time by looking at them. In this case, your insight about
the length of the tangent segment and the length of the whole secant and of its
external segment was in no way related to perpendicularity. Your proof general¬
ized immediately. The construction for the parallel situation is a special case."
"So, you're saying I should get in the habit of looking at the extreme cases
because they so frequently give insight; you're not saying exactly what insight they
will give. I guess it varies with the problem." Ed was realizing that his mathemati¬
cal investigations for the evening were drawing to a close. His long-dead Aunt
Agatha had just walked through the wall with a 5-year old version of his intoler¬
ably obnoxious cousin Lucy. "Time for some sleep," said Ed, resting his head on
his forearm, which itself was resting on his desk, and drifting off to a better world.
GEOM
Calvin threw his stuffed pet on his bed and began tormenting his father in
GEN earnest.
"So you got this bear. He walks a mile south of where he woke up, then walks
a mile due east, and finally walks a mile north. He finds himself at the exact spot
where he woke up. What color is the bear?"
"That's impossible, Calvin! If you walk a mile south and then a mile east and
then a mile north, you can't possibly get back to where you started."
"Ha, Ha! on you. Dad," mocked the charming little boy. "You just aren't
smart enough." A low guttural sound began to swell from the father's throat. He
Strategies for Problem Exploration 33
was rising from his chair with his hands locked in claw position when Calvin's
mother left her bizarre supper pot to intercede.
"You stay put!" she ordered her husband, who obeyed instantly. "Calvin,
what have I told you about teasing?"
"Oh! Don't tease dumb animals. Sorry, Dad."
"Calvin claims that it's possible for a bear to walk a mile south, a mile east,
and then a mile north and be back where it started," said the belabored father in
his victim monotone.
"The bear is white, Calvin," his mother said in her own quiet but triumphant
tone.
Calvin had become irrelevant. The father's eyeglasses tumbled to the floor.
He stepped on them and shattered them as he rose to face his wife.
"You see, dearest, the bear has to be a polar bear, because the bear had to
start at the extreme of northemness."
"Ha ha on you. Mom," Calvin interposed in an unlovely way. "That's not the
only solution."
"Of course it is, Calvin! Where else could a bear have started than at the North
Pole?"
"You forgot to look at the other extreme, Mom."
"Ha ha on you, Calvin," said the father, in a competitive and distinctly
nonparental tone. "You can't go south from the South Pole, so you're wrong."
"I didn't say to go south from the South Pole." Calvin's face exuded evil. "I
didn't say to start at the South Pole. Somewhere just above the South Pole there's
a little latitudinal circle of circumference one mile. Start anywhere on the larger
latitudinal circle one mile north of that and look what happens."
"Ha ha on you, Calvin," said his mother, getting into the spirit of this
common form of family discussion. "I can think of lots of other places to start.
Suppose you started anywhere on the latitudinal circle one mile north of the
latitudinal circle with circumference 1 /«, where n is a positive integer. Just look
what happens."
Calvin had a perfect rejoinder in a nascent Oedipal triangle. "Gee, Dad.
Mom's the smartest in the family and you're the dumbest."
A look of triumph was growing on the father's face. His tone was cool and
arrogant. "You are aware, both of you, that there are no bears down there. Your
theorizing was interesting, though of course obvious to me. I rejected it because
your results were in conflict with reality, and, after all, Calvin did say that a bear
was walking. The only valid answer was the one Calvin gave the first time when
I chose to let him feel good."
The concept of extreme is not always well defined. Even in law, a term such
as extreme cruelty requires some definition. Must extreme cruelty be lethal? If we
looked at the region consisting of the interior of a triangle and its boundary, are
the extremes the boundary or are they the vertices of the triangle?
34 THE ART OF PROBLEM SOLVING
6. Visual Representation
Only two students had signed up for Dr. Berg's geography class during the past
decade, but the school had a rare surplus of funds, and both of them had signed
up during the current semester. The course was conducted for the two of them.
"The exciting question is whether there has been a statistically significant
warming trend over the past century along the Atlantic Coast of the mideastem
states."
"Ah cum from Geohgiah," cooed Sweet Caroline.
"That's been something I've been thinking about every night for the past 10
years," said Hornet, adjusting his hom-rimmed glasses with the tape across the
bridge.
"Ah declah, ah nevah met a boy lak y'all." Even Caroline recognized that her
attraction to Hornet could be taken as certain evidence of insanity, but she had
learned long ago, in the wooded suburbs of Atlanta, never to resist an impulse.
"I would collect weather information from government reports, almanacs,
and newspapers going back as far as possible for a fixed site such as New York
Strategies for Problem Exploration 35
City, and I'm sure it would be fairly obvious whether or not there would be
climactic, I mean, climatic warming trends."
"Could we all do a joint project on this?" said Caroline, forgetting to drawl.
Dr. Berg was a kindly man and saw many advantages in a joint project, not
the most of which was there would be half as many papers to grade. "I think that's
a splendid idea. Get right on it."
A pattern of sporadic attendance by the class developed almost immediately.
On the day their first drafts were due, a 213-page report was deposited on Dr. Berg's
desk, collapsing the fragile right-front leg. Years of memoranda spilled to the floor.
"That's a beautiful blue sapphire ring you're wearing, Caroline," Dr. Berg
said to distract himself from the rage and frustration threatening to overcome him.
"We all are engaged," Caroline explained.
"We've put a lot of thought into this paper," Hornet said proudly. "I've been
thinking about it during daylight hours for 2 months."
That evening, as Dr. Berg began perusing the paper his prize (and only)
students had submitted, he noted with marked dismay that 207 pages of the 213
consisted of raw data organized by source. The sentence concluding the report
stated that no conclusions were obvious.
"They will be pleased to note that this requires a rewrite," Dr. Berg thought.
"I may even be able to teach them something important."
When Hornet and Caroline reached class that Thursday, they were puzzled
to see the chalkboard headed "Compare and Contrast." To the left were 3 of their
207 pages of data, held on the magnetic green chalkboard by cardioid magnets that
Dr. Berg had purchased at the Mathematical Notion Shoppe. To the right was an
artistically designed set of line graphs in each of the colors of the rainbow.
"That would look just beautiful over the fireplace in Daddy's mansion," said
Caroline. "What is it?"
Hornet knew. "It's a multiplex linear graphical representation of the data to
the left."
"So why do all the graphs go up?" asked Caroline, with the first genuinely
puzzled expression that had ever crossed her face.
"Upward tendencies," exploded Hornet, who had seen all the implications
of Dr. Berg's astute pictorial representation of his suggested revision of their
masterpiece.
"Hmmmm," said Caroline knowingly.
would be well over a million dollars if the series went the distance.
Sugar Pete was an eager apprentice. "The series is sure to go six," he asserted.
36 THE ART OF PROBLEM SOLVING
1/2
A
1/2
A
1/2
B
1/2
A
1/2
A
1/2
B
1/2
B
1/2
A
1/2
A
1/2
B
1/2
B
B
Strategies for Problem Exploration 37
1/8
A
1/4
1/16
A
1/8
B
1/16
B
1/2
1/16
A
1/8
1/16
B
1/4
B
1/16
A
1/8
1/16
B
1/16
A
1/8
1/16
B
1/4
1/16
A
1/8
B
1/16
B
1/2
B
1/16
A
1/8
A
1/16
B
1/4
1/16
A
1/8
1/16
B
38 THE ART OF PROBLEM SOLVING
ARITH Nick McBourbak loved second grade. The most wonderful concepts were
introduced almost every day. His mom had taught him about odd whole numbers
NUMB
a few years ago and his dad had impressed him with the joy of seeking patterns.
Strategies for Problem Exploration 39
Ms. Clearwhistle was having the collaborative groups explore sums of odd
numbers beginning with 1. The kids were moving their counters and charting
results.
"1, 4, 9,16, 25," said Jennifer. "What does it mean?"
"That's a great question," said Nicholas McBourbak with his winning Ar¬
chimedean smile. "Those are very interesting numbers." He looked down at their
chart.
1 1
1+3 4
1+3+5 9
1 +3+5+7 16
1+3 + 5 + 7 + 9 25
1+3 + 5 + 7+9 + 11
"I know the answer before you add them up," Nick stated joyfully.
"You can do that in your head?" asked Mike incredulously.
"That too," said Nick, in a tone which might have gotten him a black eye in
another setting. "Actually, I didn't add them at all." Jennifer was impressed.
"How'd you get the answer without adding?"
"The pattern, the pattern. The answers are all perfect squares, numbers
multiplied by themselves. The next number is 36 and the one after that will be 49."
"And then 64," said Jennifer. "Why is this happening?"
Ms. Clearwhistle could not suppress a smile of triumph. She had never been
very good at math, nor her mother before her, but some odd guy had spoken to
her methods class and fixed her with a steely gaze. "Never," had he said, a trifle
too loudly, "never tell the kids that you were not good in math. Give them good
things to think about and the miracles will flow." She was interested in the group's
discussion for a profound reason. She had read about this pattern in the national
arithmetic journal but had not seen an explanation of why this pattern emerged.
"Hot dog!" she thought. "I wonder what the kids can come up with."
Nick had put a yellow counter on the table and surrounded it in a square
open pattern with three blue counters. He placed five red counters in a parallel
column and a parallel row to form a three-by-three square. Now seven green
counters in a new parallel right column and a new parallel bottom row produced
a four-by-four square. "Match the column on the right and the row on the bottom
and add one in the comer; you get two times something plus one—the next odd
number and the next square."
Y B R G
B B R G
R R R G
G G G G
40 THE ART OF PROBLEM SOLVING
The three children glowed. "I feel as if we did something really good/' said
Jennifer. "Not just finding a pattern, but explaining it. I wonder if mathematicians
ever do this."
"Maybe late at night after they finish adding, multiplying, subtracting, and
dividing," said Nick sadly.
Diagrams, charts, graphs, and figures permit more intensive analysis of data
because they permit the purely rational processes involved in higher-order think¬
ing to be combined with direct sensory input. Often, something hidden in its
abstract form may blossom in a linear or quadratic curve familiar from previous
study. When data form a linear graph, the question. Why are we getting a linear
graph? becomes easier to pose and hence possible to investigate. At the highest
levels of mathematical research, purely abstract proofs must stand on their own,
edifices built from the bricks that are the axioms and held together with the cement
that is logic. Often, however, those edifices could not be built without the secret
picture relegated to the mathematician's desk drawer after publication of the pure
and pristine proof. Yet without that hidden picture or graph or figure or diagram,
many proofs would never be. It is the job of mathematics teachers to help students
learn how to construct knowledge, and using visual representations is a powerful
skill that should be mastered young.
"Dad, I'm thinking of a number between 1 and 10,000. Guess what it is,"
challenged Calvin.
"The square root of two," said his father, without looking up from The Wall
Street Journal.
"The who what of which?" Calvin asked.
Now the father looked up. "I take it back. My guess is 2,346."
"WRRONNNGGG!" screamed Calvin. "Guess again."
"One less than one more than the number you're thinking of."
"I'm not thinking," said Calvin, in an unintended admission that made the
Flame of Truth on Olympus flare.
"Raw guesses really are quite useless," said the father, kindly. "If I were a
stranger and asked you to guess what my favorite song was, what would be the
point?"
"I'd get to know you better," said Calvin, regaining his aplomb.
Strategies for Problem Exploration
41
Not only was Rip Larkin the best athlete at Franklin Pierce High School, but
he was also far and away the most popular boy, and, of course, he knew every¬
thing. After Pierce had been defeated in the first four games of their football
schedule, Larkin knew that their fortunes would have to improve, because of the
Law of Averages.
PROB "I beg your pardon, sir," said Bert Hill, the water boy for the football team.
"Things just don't catch up. Suppose you tossed a fair coin 10 times. What would
happen?"
"That's obvious!" screamed Rip, "you'd get five heads and five tails in some
order."
"Have you got a coin?" asked Bert.
Rip took a quarter from his pocket and handed it to Bert. Bert tossed it 10
times and got a total of 3 heads and 7 tails.
"I guess this wasn't really a fair coin," said Rip apologetically.
Bert had never heard Rip sound apologetic. Buoyed by this new experience,
he continued: "The coin may or may not be fair. In fact, a perfectly fair coin might
be thrown 10 times, and might come up all heads, although that's not very likely."
Rip was silent. That, too, was a new experience for Bert. "You see, when you
toss a fair coin 10 times, there is a non-zero probability for every possible combi¬
nation of heads and tails. Let me make a chart."
Bert drew a two-column chart diagram and labeled the left-hand column
"number of heads in 10 tosses of a fair coin" and labeled the right-hand column
"probability of occurrence." He filled in the top row and the bottom (eleventh) row
of the chart immediately. Next to 0 he wrote "1/210 = 1/1024," and next to 10 he
wrote the same equation.
"I'll fill in the 1 row and the 9 row next." He wrote x (1/2)10 = 10/210 =
5/512" on both appropriate lines.
"What does 10C1 mean?" asked Rip.
"It's the count of the number of length-10 head-tail strings which have a total
of 1 head and 9 tails, or one head somewhere in the string of 10. Since there are 10
places you can write the H, 10Cj equals 10."
"I guess you're going to write the 2-row and 8-row next. I suppose you'll
write 10C2 x (1/2)10. But it's not clear what 10C2 is numerically."
"Well, there are 10 places you can write one of the Hs, and for each of them
there are 9 places where you can write the other, but you'll end up counting
everything twice, because there's no difference between the Hs. Therefore 10C2 =
(10 x 9)/2, or 45.
"Let's see. I guess the 3 row and the 7 row will be 10C3 x (1 /2)10 and 10C3 equals
10 times 9 times 8, all divided by 2."
"Why are you dividing by 2?"
"You said everything is counted twice, so I'm dividing by 2."
Strategies for Problem Exploration 43
"That was when we had two heads. Now we have three. Think of them as a
red head, a white head, and a blue head. There are six different ways to write them,
so now we must divide not by 2 but by 6."
"I think I'm getting it. So 10C3 is 10 times 9 times 8, all divided by 6, or 120.
Let me try the 4 row and the 6 row myself. You should write '10C4 x (1/2)10,' and
itA would be 10 times 9 times 8 times 7 divided by the number of ways you could
arrange a red head, a white head, a blue head, and a blonde." Rip reflected on this.
I should say yellow head. Anyway, I can put any of the four colors first, and any
of the remaining three second, and any of the remaining two third, so I guess there
are 4 times 3 times 2 times 1, or 24 ways to do it, so 10C4 must be 10 times 9 times
8 times 7 over 24, or 210."
"Which brings us to the middle."
"Wow!," cried Rip. "I see the pass all the way. On the 5 row, we must write
\0C5 x (1/2)10 = (10 x 9 x 8 x 7 x 6)/(5 x4x3x2xl)x (1/2)10 = 252 x (1/2)10.' "
"Shouldn't we not have simplified?" asked Rip. "It would be simpler to check our
calculation if all of the denominators were 1,024."
"How would you check?" asked Bert.
"If you left all the denominators as 1,024, I'd add up all the numerators and
see if the sum was also 1,024, because something had to happen, and we've taken
care of all the cases."
"No wonder you're all-state left end," said Bert, grinning from ear to ear.
"Now that we've got the chart, we can really find out what's going on."
Number of Heads in
10 Tosses of a Fair Coin Probability of Occurrence
0 1/1,024
1 10/1,024
2 45/1,024
3 120/1,024
4 210/1,024
5 252/1,024
6 210/1,024
7 120/1,024
8 45/1,024
9 10/1,024
10 1/1,024
44 THE ART OF PROBLEM SOLVING
"First, let's check the sum of the numerators. 1 + 10 + 45 + 120 + 210 + 252 +
210 + 120 + 45 + 10 + 1 = 2(1 + 10 + 45 + 120 + 210) + 252 = 2(386) + 252 = 772 +
252 = 1,024. Bingo."
"Not much to choose from among 4, 5, or 6 heads," said Rip, with some
amazement. "If someone asked me how many heads they'd get if they tossed a fair
coin 10 times, my educated guess would be 4, 5, or 6, but it's no certainty even
then. By the way, something's been bothering me since you started the chart. Why
didn't you write 10C6 in the 6-row instead of 10C4? Were you counting the 10-chains
with 4 tails instead of the 10-chains with 6 heads, and simply recognizing that you
were counting the same thing?"
"It's really a shame that you dropped math after your sophomore year," said
Bert, with great sincerity. "You've got quite a flair."
"I don't have a flare," said Rip. "I'm a left end, not a quarterback."
Ace was the tennis pro at West Shore Tennis Club. He had eight classes of
NUMB
students, each with a different number of persons. Strangely, the sizes of the classes
formed a consecutive sequence of natural numbers starting with two and ending
with nine. He was paid by each student according to the number of students in
that student's class, so that he got the same amount of money for each class, but
that's another problem. He wanted to order eight boxes of tennis balls, one for each
class, and in each case to divide up the number of balls in the box equally among
the students in the class.
"I wonder how many balls I should order in a box so that I can divide them
equally in each case." He multiplied out 2x3x4x5x6x7x8x9. "That's a lot
of balls," he said. "I wonder if I could do with fewer." He decided to ask one of his
students, a Mr. Sperling, who was some sort of a math teacher.
"You really only have to worry about the primes," said Mr. Sperling. "If the
number of balls in the box is evenly divisible by 2,3,5, and 7, you only need worry
about 4, 8, and 9, and putting in two extra twos will take care of 4 and 8, and one
extra three will take care of 9."
"Huh?" said Ace.
"So I guess you want to order 2x2x2x3x3x5x7 balls in each box," said
Sperling.
"How many is that?" asked Ace.
"How should I know?" said Mr. Sperling. "I'm a math teacher, not an ac¬
countant. But," he added, "it's over 2,500 balls, and I don't think it's practical. Why
don't you order a smaller number, give the same number to all the students in each
class, and keep the extras for your special students who take private lessons? If
you ordered boxes of 50, it would work out quite well. You give 25 to each of the
students in your class of two, 16 to each of the students in your class of three, 12
Strategies for Problem Exploration 45
to each of the students in your class of four, 10 to each of the students in your class
of five, 8 to each of the students in your class of six, 7 to each of the students in
your class of seven, 6 to each of the students in your class of eight, and 5 to each
of the students in your class of nine, and you'll have 14 balls left over for your
private lessons."
"I hate having different numbers of balls left over from each class. Couldn't
we even it out? Say, what if we had exactly one to be left over from each box after
the partition?"
"Are you asking, 'What is the smallest number that yields a remainder of one
when divided by each of the natural numbers from two through nine?' "
"I guess so."
"I'm travelling to Beijing next month; I'll find out for you."
"I don't get it," said Ace.
Mr. Sperling laughed at his own wonderful joke. "We mathematicians call
that kind of question an application of the Chinese Remainder Theorem."
Mr. Sperling had, in fact, never understood the Chinese Remainder Theorem
very well, but he loved the name. One of his former students, who was always
very good in math, was coming up for open school night, and he thought he could
get the answer with very little trouble.
"Good to see you again, Polly Sue. You know, Nick is so good that he's being
moved to a specialized school next year. He developed some result about eigen¬
vectors that nobody in the school understands, and he's always right about
everything, and I liked you a lot better than I like him. Oh, by the way, what's the
smallest natural number that yields a remainder of one when divided by each
natural number from two through nine?"
"I'm probably the first parent in recorded history to be asked about the
Chinese Remainder Theorem on Open School Night," said Polly Sue, smoothing
her hair. She went up to the blackboard and wrote: "n = 1 (mod 2), n = 1 (mod 3),
n = 1 (mod 4), n = 1 (mod 5), n = 1 (mod 6), n = 1 (mod 7), n = 1 (mod 8), n = 1
(mod 9)." She then opened her handbag and removed a carefully wrapped stick
of yellow chalk. "You still don't use colored chalk properly, and it's been many
years since I first told you about it, way back when Jerry was in your class."
Polly Sue had not really been Mr. Sperling's student, but he liked to take
credit for both her and Jerry, and no one had been around long enough to challenge
the inaccuracy, except, of course for Polly Sue and Jerry themselves. He watched
as Polly Sue neatly crossed out the irrelevant equations on the blackboard with
the colored chalk. What was left was: "n = 1 (mod 5), n = 1 (mod 7), n = 1 (mod 8),
n = 1 (mod 9)."
"Why are you crossing the others out?"
"Because they're irrelevant," replied Polly Sue, finally cognizant that she was
answering his questions as well as he had answered Jerry s so many years before.
"Think about it for a minute and it'll be obvious."
"I see," lied Mr. Sperling. "But how are you going to solve even those four
simultaneous congruences?"
46 THE ART OF PROBLEM SOLVING
I'm presenting you with this explanation just so that you can check
your own investigations against it.
Since 5, 7, and 11 are pairwise relatively prime, there is no danger of
contradictory information. There are four key numbers in the analysis
of the problem I posed to you on Open School Night. They are 5x7,
5 x 11, 7 x 11, and 5x7x11.
5 x 7 is important because it is congruent to 0 (mod 5) and congruent
to 0 (mod 7), but not congruent to 0 (mod 11). It actually happens to be
congruent to 2 (mod 11). We want n to be congruent to 4 (mod 11), so
multiplying 7 x 5 by two will give us a number that is congruent to 4
(mod 11) and congruent to 0 (mod 5 and mod 7).
Similarly, we will be concerned with the congruence of 5 x 11 (mod 7).
This product is congruent to 0 (mod 5 and mod 11) and is congruent to
6 (mod 7), and we are looking for an n that is congruent to 3 (mod 7).
One way to put the appropriate question is to ask. What multiple of 6
is congruent to 3 (mod 7)? You would obtain the correct answer by trial
and error and find that 4 times 6 fills the bill. A better way to think about
it would have been to note that 5 x 11 is congruent to -1 (mod 7), and
hence, since -4 is congruent to 3 (mod 7), 4 times -1 is congruent to 3 in
this case.
4x5x11 will be congruent to 3 (mod 7) and congruent to 0 (mod 5
and mod 11).
Finally, consider 7x11.7x11 is congruent to 0 (mod 7 and mod 11) and
is congruent to 2 (mod 5). Since we want a number that is congruent to
1 (mod 5), we ask what multiple of two is congruent to 1 (mod 5), and
clearly 3 times 2 fills the bill. Thus, 3 x 7 x 11 is congruent to 1 (mod 5)
and congruent to 0 (mod 7 and mod 11).
The number n = 2x5x7 plus 4x5x11 plus 3x7x11 clearly satisfies
the three congruences n = 1 (mod 5), n = 3 (mod 7), n = 4 (mod 11),
because each term establishes one of those three congruences and the
remaining two terms have no impact with respect to the modulus
considered in the original term.
The only worry remaining is whether n is the least such positive
integer. In this case, n = 70 + 220 + 231 or 521 satisfies the three con¬
gruences but is not the least positive integer that does so.
The final key number comes into play. Multiples of 5 x 7 x 11 are all
congruent to 0 (moduli 5, 7, and 11). Thus, adding or subtracting any
multiple of 385 will not affect the three congruences. We subtract 385
from 521 and get 136, which is the least positive integer satisfying the
three congruences.
LOGIC
"What day of the week is it, Sweetcardioid?" Jerry yelled out in his sleepiest
GEN voice.
"Yesterday was Saturday, dear Fiber Bundle," replied Polly Sue from the
dining room, where Nick was serving her breakfast.
"But if tomorrow is not Monday," shouted Nick, "what can we deduce?"
"That's easy," said Jerry, hugging the best mathematician in the house."
When your mother said that yesterday was Saturday, it must have been a few
seconds before midnight on Sunday, so that when you spoke, tomorrow had
become Tuesday."
"Aha!" said Polly Sue. "When I said, 'Yesterday was Saturday,' how did you
know whether I was making a statement of fact or creating a premise for an
argument?"
"Right!" Nick chimed in. "It is important to recognize that when informa¬
tion is given in a mathematical argument, as you so carefully explained to me, there
is no presumption that the given information has any relevance to truth. Mathe¬
matical arguments must hold whether or not the premises are true. The sentence,
Strategies for Problem Exploration 49
If triangle ABC is isosceles, then it has at least two congruent angles' is true for
right triangle ABC with sides measuring 3, 4, 5."
"True," said Jerry, "but remember, Nick: Not only do your mother and I
never lie to each other, but we try to the best of our abilities not to be deceptive
while telling the truth."
"How did you guys ever stay together long enough to have me?" asked Nick,
with genuine admiration for two parents who probably would have been saints
in another era. Truth telling puts an awful strain even on my friendships in school.
When I get old enough, we'll really have to talk about this."
"Not until you're at least 25," said Polly Sue, very close to laughter.
"Or just before your wedding," said Jerry, in his most paternal tone.
"The real facts of life," said Nick.
"Anyway, if your mother had meant 'Yesterday was Saturday' as a premise,
I would have expected her to say, 'Suppose yesterday were Saturday.' When she
actually said, 'Yesterday was Saturday,' I took that as a statement."
"But Fiber dear, yesterday is a pronoun. No sentence including a pronoun
can be a statement in written exposition, and we always swore that we would
speak to each other as if our words were being written for posterity. Had I said
that July 25,1997, would fall on a Saturday, that would have been a statement."
Nick was puzzled. "What if my teacher asks me, 'What is your name?' Should
I say that 'you' is a pronoun?"
"Of course not," cautioned Jerry. "What teacher speaks with words that
should be written for posterity? Furthermore, no question can be a statement.
Finally, neither your mother nor I want to go to school again because of your literal
interpretations of instructions and questions."
"In common speech, 'Yesterday was Saturday' would be viewed as a state¬
ment. Formal mathematics is a written discipline, but it is quite proper to apply
logic in spoken arguments."
"Even at my tender age," said Nick, "I have learned the lesson that it may be
quite proper to apply logic to spoken arguments, but it does place one at physical
risk."
"One thing is certain," said Jerry, attempting to terminate the now-out-of-
hand discussion. "If we accept the cyclical sequence of days of the week in their
traditional order as axioms, then it follows that if yesterday were Saturday, then
today is Sunday."
"Amen," said Polly Sue.
8.3. I Am a Parallelogram
Quadrilaterals with
"Suppose we prove that, for quadrilaterals, each of the given conditions is suffi¬
cient to prove each of the other conditions. We then say, 'A parallelogram is a
quadrilateral with one of, and hence all of, the above conditions.'
"Consider the pedagogical implications. For you math phobic postadoles¬
cents, there would be no hidden information. You would have a gestalt of the
quintessence of parallelogramism. The mathematics would be in the extensive
proof prior to the definition, which used our illustrious discipline to establish the
necessity and sufficiency of each listed condition.
"Contrast this with the traditional approach through which each of you has
long suffered. Minimal definition hid the full meaning of the word being defined
rather than elucidating it.
"Even the commitment to a minimal definition is hypocritical. What is the
minimal definition of a regular polygon?"
"That's easy," said the indescribable one (Oscar). "A regular polygon is a
polygon all of whose sides and all of whose angles are equal in measure.
"I thought a neater definition was the one a substitute teacher gave us late
one Friday. He said, 'A regular polygon is a polygon with a well-defined center
such that for every line through the center there exists a one-to-one correspon¬
dence between the points on the polygon on one side of the line and those on the
other, such that the perpendiculars to the line from each pair of corresponding
points are equal in length.' "
"That doesn't work for polygons with an odd number of sides, but it is
elegant for those polygons with an even number of sides."
"In any event," said the Master, "the first definition was not minimal and the
second was not general. A minimal definition for a regular polygon might be as
follows: a regular n-gon is an n-sided polygon in which all of the angles are equal
in measure and at least n -1 sides are equal in length. I challenge you to prove that
such a polygon is regular, but I assure you that it is. Alternatively, consider the
following: a regular n-gon is an n-sided polygon with n congruent sides and at
least n - 1 congruent angles.
"Neither of the last two definitions I gave to you is psychologically sound.
Fewer of you would be math phobic if definitions were formulated not through
minimalist criteria but through their psychological soundness. Be children again
in high school and think which of the following definitions of a square would be
less likely to produce mathphobes?
"The point that I make is this: If two definitions are each mathematically
sound, why not select the one that gives a greater instantaneous gestalt to the
greatest number of people? Redundancy is superior to obscurity."
Strategies for Problem Exploration 53
In section 7.2, we raised the point that the standard technique for solving
equations consists of making deductions from the equation to be solved that yield
necessary conditions for the solution—that is, that yields a set which includes all
numbers that are candidates for solutions of the equation. The check is the
sufficiency proof, which weeds out the elements of the candidate-set that are
unqualified for the office of solution.
Each time a student engages in a standard equation-solving procedure, the
student in his deduction and check is giving a proof of necessity and sufficiency.
"The locus of points satisfying a condition is nothing more than the set of I GEOM
points in a prespecified space that satisfy that condition. Consider acute angle
ABC. Find the locus of points equidistant from the sides of that angle."
Nick had been working on this problem for 15 minutes when Ms. Gale
walked over to his collaborating foursome. "What's going on?" Dorothy Gale in¬
quired. "Everyone else has finished."
Nick looked up. His teacher had come from the heartland of America, but
she seemed to lack heart. Nick felt feisty and decided that he wouldn't mind
another visit with the principal.
"They may have finished the problem in Kansas, but they're still working in
Massachusetts."
On the drive home from the latest in the series of principal's disciplinary
conferences, Polly Sue was acutely aware of her son's impending adolescence and
was uncharacteristically taciturn.
"Mom, isn't there anywhere where correct mathematics is more important
than political sensitivity?"
Polly Sue decided to take a purely academic approach to a delicate question
she preferred not to answer. "What problem were you working on?"
"You know Miss Gale?"
"She wasn't in the school when your father and I attended. She had some
special history that we talked about some years back, but it escapes me."
"Well, she asked us to find the locus of points that were equidistant from the
sides of an angle. I sharpened her question by limiting my inquiry to the plane of
the angle. Nevertheless, she thought that the answer was the line containing the
angle bisector of the angle, and most of the kids in the class were astonished to
consider that the half-ray exterior to the angle was part of the locus. It didn't take
me long to see that the problem was considerably more complex than Ms. Gale
had intended, and I actually became quite engrossed in my investigation. I guess
I got angry when she interrupted me, and I made a disparaging remark about
Kansas. In retrospect, and in toto, I was off-base."
54 THE ART OF PROBLEM SOLVING
Polly Sue had not been fully attentive, but for some reason, something in her
son's answer struck a chord. "What did you say about Toto?"
"I didn't say anything about Toto. Who's Toto?"
Polly Sue was in that state of proximal development when she could not put
the fingers of her mind on the substance of her recollection.
"Anyway, Mom. I realized that to find the necessary and sufficient condition
for a point in the plane of angle ABC to be equidistant from the sides of the angle,
it would be necessary to define the distance from a point to a ray. Most of the kids
assumed that the distance from a point to a ray was the distance from the point to
the line containing the ray, but I recognized immediately that that was not right.
If we drew a line perpendicular to the ray from its endpoint, and took any point P
in the half-plane not containing the ray, the distance from P to the ray would be
the length of the line segment joining P to the endpoint of the ray. I drew rays
perpendicular to AB at B and CB at B, each of which formed an obtuse angle with
the ray of angle ABC to which it was not perpendicular. Let's call the two rays BM
and bU. I realized about n seconds before Miss Gale interrupted me that not only
was the ray bisecting angle ABC part of the locus, but so was every point P in the
interior or on the boundary of angle MBN, and I can prove the necessity and suf¬
ficiency of being an element of the union of that angle, its interior and the bisecting
ray for assuring that a point is equidistant from the sides of angle ABC. I deserve
a principal's citation for this insight."
"You got one," said Polly Sue ruefully.
"So tell me. Mom, why does Miss Gale so often wear red shoes?"
9. Sequencing
Four events are given, and the children are asked to put them in proper time
sequence:
"As always," barked the strange boy with the long ears who sat in the back
of the class, "your sequencing problems are ambiguous."
Nobody got angry, because nobody understood him. As always.
Big Ed had made a great deal of money from the game of 31 over the many
years when he played it with the unwary. The rules were simple enough so that
his victims came to believe that in just one more play of the game they would win.
They rarely did.
Thirty-one is a two-player game with alternating turns. The initial layout
ARITH
consists of cards marked with the numerals 1 to 6 with exactly four of each; all the
GAME cards are sorted and placed face up on the table. The victim has the choice of
picking first or second. Each player in turn picks a card, places it face down on the
table, and adds the number indicated by the card to the current total of previously
selected numbers (which is zero at the start of each new game) with the stipula¬
tions that when a particular domination is exhausted, that number is not available
and that under no conditions may a player turn a card making the total greater
than 31.
Ed liked to use the aces through sixes from a standard deck of cards. "Do you
want to go first or second?" he asked.
"You go first," said The Victim.
Ed never used the correct strategy at the start. He turned a 6. The Victim
thought (about nothing in particular) for a few seconds and turned another 6,
saying, "That makes 12."
"So it does," confirmed Big Ed. He now turned a 5 and said "17."
Again The Victim thought. He turned the third 6 and said, "23."
Ed turned an ace in a bored gesture. "24."
For the first time The Victim actually thought about the situation. He realized
to his mortification that the $10 wager had been lost. Big Ed would be able to reach
31 exactly to win the game. "I guess 24 wins."
Ed began to become really interested in the situation. The best suckers were
always the ones with a little bit of knowledge.
"So you get to 24," challenged the bookie-tumed-shark.
The victim knew a little mathematics. He began to analyze the significance
of 7 in the game: 7 was the least number unmakeable in a single turn. Thirty-one
minus 7 was 24. "That is the start of a backward sequence of winning moves," he
thought triumphantly. "The sequence is (31, 24,17,10, 3)," he continued silently.
"I've got him."
"Let's up the wager a little," said The Victim ripely. "Can we make it a
hundred?"
"Go slow," said Ed in a feigned cautionary tone guaranteed to reel in his fish.
"Make it a hundred. I want to go first."
"Don't say I didn't warn you," said Ed, thinking about tomorrow.
Strategies for Problem Exploration 57
The Victim turned the first of the four 3s and said, "3."
Ed turned a 4. “7
The Victim turned the second of the four 3s. "10."
Ed turned a 4. "14."
The Victim began to feel uncomfortable. He turned the third of the 3s, say¬
ing "17."
Ed turned the next-to-last 4 and said, "21."
The Victim turned the very last 3 and said, "24."
Ed turned the last 4. "28."
No 3s remained. If The Victim turned a 1, Ed would win by turning a 2; if
The Victim turned a 2 Ed would win by turning a 1. The hundred dollars had been
lost.
"Tell you what," said Ed. "Til give you a week to think about the game. We'll
meet here next week and play one game for $250. What do you say?"
"You're on!"
The Victim had a plan. He was going to see a client of his named Jerry Mc-
Bourbak that very evening to discuss a life insurance policy, and that McBourbak
was a math whiz. He would ask McBourbak for help.
After he had wrapped up a nice commission. The Victim broached the game
to not one but three very interested pairs of ears. He recounted the events of the
day with great accuracy.
"He was able to force you to deplete the 3s," summarized Nick.
"Right," said The Victim, still addressing Jerry rather than the two better
mathematicians in the room.
"It is apparent that going first and playing 3,4 or 6 loses," said Jerry. "We $ee
the depletion strategy if you take a 3; the other two moves permit Ed to use his 10,
17, 24, 31 sequence."
"The only time 24 is a winning move is if there are one or more of each
denomination left," said Polly Sue.
Nick smiled. "Seventeen wins if there are two or more of each denomination
left, 10 wins if there are three or more of each denomination, and starting with 3
would win if there were four or more of each denomination left, which is impos¬
sible in Ed's little game."
"So should I ask to go second?"
"Not at all," said Nick as Polly Sue and Jerry beamed with pride. "Turn the
depletion strategy against Ed. Start with 5. He'll have to take the second 5 to make
10 or let you get into winning sequence with enough of each denomination left."
"I get it," shouted The Victim erroneously. "Then I take the third 5 and
make 15."
"No, no!" said Nick. "Ed will take a 4 making 19, forcing you to take the last
5 to make 24. He will then win by turning a 2 to make 26 and you'll have no 5s
left."
"So starting with 5 doesn't work?"
58 THE ART OF PROBLEM SOLVING
"Sure it works. After Ed takes the second 5 to make 10, you turn a 2 to make
12. That forces Ed to take the third 5 to make 17. Now you turn a 2 to make 19
yourself. He has to take the last 5 to make 24 and you now turn a 2 to make 26
defeating him."
"A word of caution," said Polly Sue. "I know Big Ed. Don't let him change
anything before you play and don't play him again."
"Come on. I'm no sucker. What could Big Ed do?"
The week passed slowly for The Victim, but he thought it would give too
much away to request an earlier date. "I'm ready," he announced to Big Ed.
"So am I," said Ed. "Do you want to go first or second?"
"First," said The Victim gleefully.
"And I'll double the bet if you protect me from my unlucky number, 11."
"What does 11 have to do with anything?"
"My granddad was shot in Abilene after rolling an 11 in a game of craps and
it's been a family curse. What do you say? A $500 bet if you protect me from 11."
"OK, I won't make 11."
"OK. You won't make 11 and at no time will the sum of the cards you alone
have turned be 11."
"I agree."
Ed scribbled on a piece of paper: "I agree not to play so that I make the
number 11 at any time in the game either with my own cards alone or as a sum of
all cards played. Ed agrees to pay me $500 if I win; I will pay him $500 if I lose our
next game of 31 (rules attached)."
"Read the rules and initial them and then sign the agreement. See, I already
signed it at the bottom 'Countersigned by Big Ed.' "
The Victim signed. "Now can we play?"
"Sure thing," said Ed, grinning from lobe to lobe.
The play proceeded: Victim 5, Ed 5 (making 10), Victim 2 (making 12), Ed 5
(making 17), Victim 2 (making 19), Ed 5 (making 24). The Victim reached for the 2
to make his planned total of 26 with no 5s left. "No can do," said Ed.
"Why not?"
"Eleven protection," said Big Ed. "If you take a 2 it makes the sum of all the
cards you took so far be 11. No can do."
Another fish had been reeled in.
ALG
In section 3.6 we discussed the beautiful mathematics that can result from
ARITH giving a recursive definition of a sequence and seeking an explicit definition or
from giving an explicit definition and seeking a recursive one.
Mr. Sperling wrote on the blackboard: "An arithmetic sequence is a sequence
of real numbers such that there exists a real number d such that for all integers
n > l,an=an_1 + d/'
Strategies for Problem Exploration 59
Children can begin to learn about sequencing long before they learn to read.
When a child learns that socks are put on before shoes, he or she has learned a
sequence with two terms. Puzzle picture sequences that must be arranged in
chronological order appear in picture books. We learn about sequences in a
recursive format. In practical terms, we learn what to do next; in a long process we
are often able to say what has come prior to or subsequent to something else.
Talking about sequences explicitly requires more formal instruction.
Sequences lend themselves to accessible but important investigations. It is
easy to discuss the limit of an infinite sequence intuitively and to pose questions
such as whether or not the sequence formed by applying simple arithmetic
operations to two convergent sequences is itself convergent and whether the limit
60 THE ART OF PROBLEM SOLVING
of the new sequence can be found by applying the same simple arithmetic opera¬
tions to the limits of the individual sequences. The concept of divergent sequences
is easier to talk about and the divergence of specific sequences easier to prove than
corresponding results about continuous functions. Most of all, the recognition of
the recursive-to-explicit redefinition problem (and its reverse) opens intriguing
new possibilities to students. "Predict the next term and explain your thinking"
problems help students to see that alternative answers may arise when least
expected and help to make students more appreciative of the difference between
a convincing argument and a proof.
The possibilities are extensive. Students could be told the Gauss anecdote
recounted in section 9.4 and then asked to generalize it to obtain a formula for the
nth partial sum of an arithmetic sequence. They might be introduced to difference
equations as one means of solving recursive-to-explicit redefinition problems and
might be asked at the university level to explore how some of the techniques for
solving difference equations relate to certain techniques used to solve differential
equations. Some questions about sequences have intuitively evident answers with
elusive proofs; others have counterintuitive answers.
At the present time, not enough is being done early in formal education to
mine the richness of the sequential ore. Many students would be helped to develop
a love of mathematics by addressing this lack in imaginative ways.
"Three days before yesterday it was the day after Sunday," said Ms. Othmer.
"What day is it today?"
Lucy looked at the calendar. "Monday."
"No, no, Lucy," Violet whined, "Three days before yesterday it was Monday."
"No it wasn't."
(The reader is urged to reread section 8.1 rather than taking sides prema¬
turely.)
"So let's count ahead. With Monday as zero, count up to three and get to
Thursday, so yesterday was Thursday. Today is Friday."
"No it isn't," insisted Lucy. "We came back to school this morning after a
weekend off. Today is Monday."
"Let's go along with Lucy," suggested Ms. Othmer. "Lucy, what was yesterday?"
"Sunday."
"And what day was it 3 days before yesterday?"
"Thursday."
"Therefore, if I said that 3 days before yesterday it was the day after Wednes¬
day, not only would it be Monday on the calendar, today would be Monday in the
problem."
Strategies for Problem Exploration 61
That's 2 days before the day you would have said in the problem you gave
to Lucy just now. Wouldn't that mean that in the problem you gave us,the answer
had to be 2 days before the day you mentioned originally? Isn't that an interesting
way to get the answer to your problem!" said Franklin with great joy. "I'll bet there
are a lot of problems that can be solved by saying anything you want and working
backward and then noticing the relationship between the answer you get and the
thing you said and then mirroring that relationship in the actual problem to get
the actual answer."
"You need psychiatric counseling, Franklin," said Lucy. "I have office hours
this afternoon."
Ms. Othmer was thinking long and hard about what Franklin had said. After
a very long pause she said, "That's excellent thinking, Franklin. Try to write it out
in your journal as simply as you can. I don't want you to lose it."
"Why didn't you give us the problem with Wednesday instead of Sunday?"
asked Lucy critically.
"Because Ms. Othmer wants to teach us to be better thinkers and problem
explorers," said Franklin, returning his attention to the comic book he had hidden
in his mathematics notebook.
"The sum of any two consecutive integers is odd," Binky said to Nick, "but
the sum of any three consecutive integers is divisible by 3. The sum of the sets of
NUMB
four consecutive integers that I have tried are never divisible by 4, but all the sums
of five consecutive integers I've tried are divisible by 5. Does this always work for
odd numbers and never for even?"
Nick was part of the new mathematics mentoring program in his school.
Binky was a precocious first grader who everyone had been happy to get out of
their hair and into Nick's. Nick loved it.
"What made you say that it always works for three consecutive integers?"
"Because in any set of three consecutive integers, exactly one of them is
divisible by 3, one is one more than a multiple of 3, and one is one less than a
multiple of 3. They must add up to a multiple of 3."
Nick found it easier to think in terms of congruence. In any set of an odd
number of consecutive integers, one would be congruent to zero modulo that odd
number and the others would be nicely paired around zero, one of them congruent
to 1, one to -1; one congruent to 2, another to -2; they would continue to fall in
pairs adding up to zero modulo the odd number. If you added an even number of
consecutive integers, the lack of a middle would ensure a non-zero sum modulo
that even number. Binky's guess was right on target. "Why don't you try working
out an argument for seven consecutive integers and then see why that argument
would fail for four or six consecutive integers?"
"I did that last night. I took the septuple (5, 6, 7, 8, 9,10,11) and crossed out
the 7 in red. Then I crossed out the 6 and the 8 in green, the 5 and the 9 in blue.
62 THE ART OF PROBLEM SOLVING
and got stuck. They were on opposite sides of 7 and the less-than part canceled
out the more-than part for each pair. Ten and 11 are both more than."
"Why not think 7 and 14?"
Binky looked puzzled for a second or two. "Oh I get it. Ten is three more than
7, and 11 is three less than 14. Or I could do it the other way, with 10 four less than
14 and 11 four more than 7. The multiples of 7 are all the same in this kind of
problem. That's great!"
Binky went right on. "For 6 I took the 6-tuple (5, 6, 7, 8, 9,10)."
Nick knew exactly why Binky had not used parallel verbal construction on
his tuples.
"I crossed out 6 in red and then crossed out 5 and 7 in green. With your 'next
multiple' trick I could now cross out 8 and 10 in blue. Nine would be left alone,
three more than 6. There's no middle with even numbers; there always is a middle
with odd numbers."
Nick couldn't wait for Binky to get a little older. What a great kid!
maintained its careful modulation. "Think of it this way: You have the general
parabola neatly enveloped on this piece of waxed paper. Are you not free to draw
coordinate axes anywhere you want on that paper?"
A new voice, wholly unfamiliar to Veronica, responded, "We are nothing if
we are not free."
"Who was that?" asked Veronica.
"Exactly right," said Leela, getting up lightly and leaping through the open
window to the ground 20 feet below with perfect grace along a parabolic arc.
Veronica was left to ponder her original question with powerful new in¬
sights.
GEN
Some students and their parents pay a lot of money to learn about specifica¬
tion without loss of generality. In some preparation courses for college entrance
examinations, students are taught to posit specific answers to problems as a basis
for analysis. For example, a student is asked to "find the single discount equivalent
to giving an 8% discount after a 10% discount." The student is told to imagine that
the item costs $100, though that information is not given. The 10% discount would
lead to a price of $90, and the 8% discount on the $90 would reduce the price by
another $7.20. The final selling price would be $82.80. They are now taught to say
that the composite discount is 17.2%. This is a simple application of the strategy
under discussion. Sometimes this strategy permits people to develop counterex¬
amples to false generalizations or to gain insights that lead to proofs. Once again,
schools have been remiss in developing skill in this important strategy in the
majority of students. Opportunities for deeper insights into the nature of mathe¬
matics and the thinking needed for creative productivity in the discipline are lost.
Students are missing out on a lot of fun.
It is traditional in American football for the team captains to toss a coin at the
start of a game to determine which team kicks off and which team receives. Rip
watched as the coin spun up and Lorenzo called tails.
Down spun the coin, lodging inextricably in a circular metal grate, but
showing the United States Treasury Building.
"I didn't get a head," shouted Lorenzo, triumphantly. "We elect to receive."
"Hold it, Bozo," said Ox, the quarterback for the Downtown Questionable
Trades High School Football team, the Convicts. "I've seen heads and I've seen
tails," roared Ox, "but this is a building, and that ain't no head, and it ain't no tail."
Strategies for Problem Exploration 65
"I find absolute value inequalities daunting," Binky admitted. Nick reflected
that most second graders would.
"For example. I've been working on this one: 14x - 3 I + I 7 - 2x I < 3. My
general strategy is to break the analysis into cases, allowing for the four key
possibilities:
IA' x > 3/4 and 7/2 > x or 7/2 > x > 3/4
IB' x > 3/4 and 7/2 < x or no solution
IIA' x < 3/4 and 7/2 > x or x < 3/4
IIB' x < 3/4 and 7/2 < x or no solution
"I originally found this very confusing, because the two conditions with
non-null solution sets seemed to combine into the condition 7/2 > x, but that is not
so. They must be kept separate, because each of them leads to a different inequality
when the absolute value signs are removed.
"(IA) leads to 4x - 3 + 7 - 2x < 3, or 2x + 4 < 3, or 2x < -1, or x < - 1/2. Since
the set of points {x I x < - 1/2} has null intersection with the set of points deter¬
mined by (IA), namely, those points x with 7/2 > x > 3/4, (IA) leads to no solutions.
66 THE ART OF PROBLEM SOLVING
"The only remaining possibility is that (IIA) leads to solutions. (IIA) results in
the given inequality becoming 3 - 4x + 7 - 2x < 3, or 10 - 6x < 3, or 7 < 6x, or 7/6 <
x. Since the set of x with 7/6 < x has null intersection with the set of x with x < 3/4,
I couldn't find any solutions whatsoever to the inequality."
"You said that you found absolute value inequalities daunting," marvelled
Nick, "but where does the daunt come in?"
"I expected to have an intuitive grasp of why the inequality has no solutions,
but, except for the analysis I shared with you, I see no reason to have guessed that
there is not a single real number in its solution set. I so rarely have found anything
counterintuitive that this daunts me."
Nick was pleased to have something to say to Binky. "What you say is not
precisely so . The result is not counterintuitive, because your intuition really gave
you no clue as to whether this inequality had any solutions. Perhaps a better term,
although I admit it is a term of my own invention, would be that this result is
paraintuitive, much as your ability to move objects with your brain is paranormal."
"I can't move objects with my brain, Nick."
"Yet."
"Now just this morning, I started working on 18x2 + 2x - 3 I + 17-2x1 <3.
Can we talk about that one?"
"Next time," said Nick, with a curious mixture of inexplicable emotions.
"I have 12 coins, 11 of which are identical and true, and one of which is an
insidiously clever counterfeit," said Professor Moriarty to his freshman mathemat¬
ALG
ics class at Cambridge University. "It is your job to find the counterfeit in the fewest
possible weighings."
"Oh, no," said young Van Dusen. "The old balance scale problem. How
boring."
"You arrogant pup!" said Moriarty, accurately reading Van Dusen's mind.
"You're no Thinking Machine yet. Actually, you don't have a balance scale. One
of my colleagues, a Frenchman named Dupin, has invented the most accurate
scale, which yields weights correct to a ten-millionth of a gram, and I assure you
that the difference in weight between a true coin and the counterfeit exceeds the
tolerance of the scale. Since you find this so simple, would you please explain to
the class how many weighings you require and exactly how you will proceed?"
Two students in the class actually snickered. One was a Chinese young man
who had enrolled under the name of Ming, and the other was a promising young
Swiss science major named Sivana. "An interesting conundrum," said the Chinese
gentleman.
"Van Dusen would need the wisdom of Solomon to solve this one," thought
Sivana.
Strategies for Problem Exploration 67
"A truly curious problem," said the one known as Ming. "If we had but one
coin and were ensured a counterfeit with the set, we could identify the counterfeit
with no weighings; yet if we had two coins, the problem would be unsolvable."
Sivana picked up the chain of reasoning with some delight. "The simplest
productive case to contemplate is the one with three coins. It is obvious even to
Van Dusen that three weighings would suffice. But perhaps it is an open question
to him whether it can be done in fewer."
"It is eminently clear that it cannot," said the one called Ming. "One cannot
gain knowledge of the weights of all three in two weighings, and knowing the
weights of two would leave the question of which is counterfeit unsolved, were
the two of different weights."
"How far could one go with three weighings?" asked Sivana, rhetorically. "I
conjecture that we may be able to do as many as five."
"Seven," said the one called Ming. "Recognize that in two weighings, know¬
ing the weight of a true coin, one could determine the one that is counterfeit among
four coins with certainty. With one weighing, knowing the weight of a true coin,
one could do the same thing with at most two coins. In general, in n weighings,
after we know the weight of a true coin, we can find the counterfeit among a set
2” coins by dividing successively by 2, always limiting our attention to the set
known to contain the counterfeit.
"With seven coins, we might productively begin by weighing four coins on
the first trial, and then two of them with one of the three remaining coins on the
second trial. Let us say that the weights are, in order, p and q. If q/3-p/A then all
five coins which have been weighed are true, and we can compute the weight of
a true coin, so that on the third weighing, by weighing the sixth coin alone, we can
determine whether it is the sixth coin or the seventh coin that is counterfeit."
"But what if q/3 * p/4? How could we possible determine the counterfeit
from among the five candidates in one weighing?"
"Let us call the first four coins A, B, C, D, the fifth coin E, and the other two
coins F and G. The combined weight of A, B, C, and D we have called p, and the
combined weight of A, B, and E is q. We have assumed that q/3 * p/4, so that we
know that one of the five coins A, B, C, D, E is the counterfeit. On the third and
final weighing, we can weigh A and C together obtaining the weight r. We proceed
by breaking into cases, systematically accounting for all possibilities.
"If r/2 = q/3, we deduce that all of the coins A, B, C, E are true and, therefore,
D is the counterfeit. If r/2 = p/4, we deduce that E is the counterfeit. If all three
weighings are incompatible (i.e., if r/2 * q/3, r/2 * p/4, and q/3 ± p/4), we must
resort to algebra.
"Let w be the weight function. We solve simultaneously the equations
»
"Under each of four cases, using t to represent the weight of a true coin and x the
weight of a counterfeit:
Case I: A is counterfeit
x + 3t = p
x + 2t = q
x+t=r
This would lead to t = r/2 and x = q-r. Again, we see whether both of these
quantities are positive and whether 3t + x does equal p. If so, B is indeed counter¬
feit."
Sivana interrupted. "What if cases I and II both lead to consistent results?"
The one called Ming smiled. "But, dear Sivana, consider the value of p in each
case. In case \,p-2r-q + 3{q- r), or 2q - r. In case II, p = q - r + 3(r/2) or q + r/2.
Thus 2 q-r-q + r/2, or q/3 = r/2. But this result violates our assumption that the
first two weighings were not consistent!"
"I'll be a mad scientist!" cried Sivana, "And I'll wager as we carry this
through, the same type of thing will continue to occur."
"Let us test your wise wager.
"We must now rule out the possibility of simultaneous consistency of this case
with each of the previous three cases:
Cases IV and I
Cases IV and II
"One thing remains," said Sivana. "How do we know that it would not be
the case that none of the four leads a consistent solution?"
"That violates either the given information that there is exactly one counter¬
feit coin or our previous deduction that the counterfeit had to be either A, B, C, or
E, because of the inconsistency of each pair of three weighings."
Professor Moriarty had approached them. "Have you decided how many
weighings are need for 12 coins?"
The one called Ming looked up with the expression that a teacher would have
in contemplating a brilliant student rather than what seemed to be the actual
situation. "Dear Professor Moriarty, that number 12 was a snare. You could do as
many as 15 in four weighings. In fact, in n weighings you could find the counterfeit
from among as many as 2n - 1 coins. I'll mail you a general proof of that next
month."
"Why will you have to mail it?" Moriarity asked.
"A slip of the tongue," said the one called Ming, dismissing Moriarity.
Sivana contemplated his classmate closely. "That was no slip of the tongue.
Would you share the truth with me?"
"I am simultaneously enrolled in three doctoral programs. As you know, I
am enrolled in a Doctor of Philosophy program at this university. I am also a
candidate for a Doctor of Medicine at the Sorbonne and for a Doctor of Laws at
Harvard University. Soon you may call me Doctor ..."
70 THE ART OF PROBLEM SOLVING
"Doctor Ming?"
"No. There will be another name."
ALG
"Other than 3-4-5, 5-12-13, and 7-24-25, are there any other primitive Py-
thagorean triples?" Mary asked Binky.
Although Binky was only 9, he had developed a crush on the little red-headed
second grader who her teacher was touting as a roaring genius, based on 3 months
of teaching experience.
"Do you know how to program a computer yet?" Binky asked.
"Sure," Mary said. "My mom taught me last year."
"Write a program generating triples (p, q, r) for all ordered pairs of integers
(m, n), 0 < m < n, by the algorithm p = n2-m2, q = 2mn, and r = m2 + n2, and print
all such triples for which p and q are relatively prime."
"What does 'relatively prime' mean?"
Binky felt a warm glow through his whole body. "Two integers are relatively
prime if they have no common prime factor."
Mary thought for a moment. "So 8 and 9 are relatively prime, even though
neither of them is prime."
"You've got it," said Binky.
The next day, Mary, with a little bit of help from her mom, came to school
with four pages of densely packed printout of primitive Pythagorean triples. In
addition to standards such as 8-15-17 and 9-40-41, her list included such unfamiliar
triples as 177-1736-1745.
"My mom says that all triples of the form you gave are Pythagorean, though
not necessarily primitive. I really like this list. I'm gonna have a lot of fun with this
in school."
"Not for a few years," said Binky. "They don't teach algebra in second grade.
You'll have to wait until fourth grade in this school. Then you'll have fun."
"I wonder if I would have had as much fun with this list if I knew the proof."
72 THE ART OF PROBLEM SOLVING
"Sure you would. The computer helps me to see that even the most abstract
proofs have real meaning. It helps to make mathematics real."
Mary wanted to learn chess as quickly as possible so that she could have the
pleasure of playing Binky on fairly equal terms. She asked her father for a chess
program with many levels of skill. The program had 15 levels, ranging from "We
Just Move" to "If You Beat Us Now, You Ought to Be Ranked." Within 1 week,
Mary had gotten to Level 6, which was called "You Can Probably Beat Your
Father." She decided that the program had been written by a woman, because
Level 7 was called "You Can Probably Beat Your Mother."
Binky was genuinely surprised that it took him 16 moves to mate Mary. Most
opponents he played fell in under 12 moves, and Mary had just learned the game
a couple of weeks ago.
"I love learning chess on a computer," said Mary. "I turn it on when I feel
like it, I can take back moves, it never laughs or gloats, I can turn it off anytime I
want and decide whether or not to save the games so far. I can choose the level on
which to play, and I can take the things I learned and write about them during or
after the play. I think I could learn anything if I didn't have to deal with people
saying, 'Very good, Mary, but I suggest you talk it over with some of the other kids
in your group before entering it in your journal.' "
"Someday," Binky predicted, "computers using games, simulations, and
multilevel instructional density will be commonplace."
"I don't understand what you mean yet, but when I do, I'm sure I'll agree
with you."
Binky reflected that Mary wasn't a roaring genius in math but, in his own
words, she was plenty good.
NUMB 1 Mary had reached fourth grade without getting to use her sheaf of primitive
Pythagorean triples, but a wonderful new opportunity presented itself when Mrs.
Leeds said, "Between any two fractions, there are infinitely many fractions."
"She meant rational numbers, Mary. Teachers often confuse numbers and
numerals, but nowhere is this more confusing to students than when they're
studying rational numbers. A fraction names a rational number. It is a symbol,
specifically a numeral. I mean, how many letters does Mary have?"
"Four," Mary replied, in a perplexed tone.
"Let me see them," said Binky.
"Can I borrow a piece of paper?"
"Why?"
"To write my name."
Strategies for Problem Exploration 73
I didn t ask how many letters there were in your name. I asked how many
letters Mary had."
Mary was still perplexed. Binky took out a piece of paper and wrote a
minuscule numeral 5 in the lower left-hand comer and a monstrous numeral 2
spanning the rest of the sheet. "Which represents the larger number?"
It all hit home at once for Mary. Of course 5 was a bigger number than 2, and
of course she had mailed the letter her mother had given her this morning before
she had got to school.
But if Mary has no letters, how would you ask me about the number of
letters in my name?"
"I'd use your name's name. Mary is a girl (oh yes), but 'Mary' is a girl's name,
and has four letters. Mary had one letter when she left her house, Mary has no
letters now, but 'Mary' always has four letters."
Mary went up to the blackboard and wrote, "One half is a rational number,
as is 1/2, but '1/2' is a fraction. Also, '0.5' is neither a rational number nor a fraction,
but a decimal, yet 0.5 = 1/2 is a rational number. Finally, '0.5 = 1/2' is an equation."
"Brava!" said Binky.
"So," continued Mary, "if two rational numbers are named by fractions, there
are infinitely many nonequivalent fractions naming rational numbers between
them."
"You ought to be on videotapes," said Binky.
"So I was wondering, Binky, how can you predict the fraction with least
denominator that names a rational number between two given rational numbers?"
"Let's go to the computer," said Binky.
After about 20 minutes, they were running a very nice program that permit¬
ted them to enter an ordered pair of ordered pairs, each component of which
consisted of relatively prime natural numbers, and to obtain the ordered pair
which, when used to name a fraction by writing the simple numeral for its first
component before a slash line followed by the simple numeral for its second
component, was the fraction with least denominator naming a rational number
between the rational numbers named by the fractions generated by the two input
ordered pairs, when the process described above was applied to each of them.
(Note: The previous accurate but nearly unreadable sentence is not recom¬
mended for classroom instruction. It is far better to illustrate the meaning with a
couple of samples, such as:
Input Output
13.1. Syllogisms
LOGIC
"If you let me have a bite of your peanut butter and jelly sandwich. I'll let
GEN you have a bite of my croissant," said Mary to her friend Michelle.
"If you let me have a bite of your croissant. I'll let you have a bite of my peanut
butter and jelly sandwich." Michelle paused for a minute. "So give me a bite of
your croissant."
Strategies for Problem Exploration 75
Why? said Mary. "You haven't given me a bite of your peanut butter and
jelly sandwich yet. Somebody has to do something before any promises need be
kept."
I m not going to give you a bite of my peanut butter and jelly sandwich,"
said Michelle.
"I'll let you have a bite of my croissant anyway," said Mary.
Michelle took a bite. "Now you're a liar," she said.
"No, I'm not," said Mary, "I haven't broken any promises. By the way, are
you going to let me have a bite of your peanut butter and jelly sandwich now?"
"No," said Michelle, "I told you I wasn't going to."
"Then you're a liar," said Mary. "You broke a promise."
"How could I be a liar if you're not?" asked Michelle.
"Because I didn't do the 'if' part of anything that required me to do a 'then'
part. But I did the 'if' part of what you said, so you're a liar for not doing the 'then.'
'If-then' only obligates us to do something when the 'if' happens."
"I never understand you, Mary," said Michelle. "You always make me feel
angry."
"So tell me, Binky, what is this all about? If I know 'if p then q' and I also
know 'if q then r/ then the Law of Syllogism lets me know 'if p then r,' but that
doesn't really tell me anything unless I somehow know 'p.' You could also know
'not r/ which would let you deduce 'not q' and 'not p.' But mathematics always
seems to be about deducing things from if-then sentences. How will we ever know
anything?"
"Mathematics can be done in a black box, without reference to reality, but in
order to know something about reality, we must have a peephole in that box.
Without real-world input, we can have the most elaborate system, and yet have,
no real-world output."
"So a syllogism without anything independently known or given is just a
silly gism."
"You have a way with words, Mary."
"Doctor Suss was talking about deductive and inductive reasoning today in
our debate seminar. Some of the students who are also taking fourth-year mathe¬
matics said that they knew about inductive reasoning from their study of mathe¬
matical induction. 1 said that mathematical induction is deductive, and they all
started laughing. I'm right, am I not, Dad?"
Jerry beamed at his son. "Of course you are. I think it a misnomer to call
inductive reasoning 'reasoning.' You are not really reasoning when you do induc¬
tive reasoning. You're making educated guesses."
"Which doesn't exactly tie into something that I've been meaning to ask you
about. We have a new guy in the class who transferred from some foreign school.
76 THE ART OF PROBLEM SOLVING
"In that old billiard ball problem, we are attempting to prove that any set of
solid-colored billiard balls must consist of balls all of the same color. The fallacious
attempt at natural induction, which I prefer to say to the common 'mathematical
induction,' goes as follows:
• Anytime I have a set of one billiard ball that is solid color, every ball in the
set has the same color (establishing condition 1)
• By the inductive hypothesis, we may assume that any set of m solid-
colored billiard balls are all of the same color. You hand me a set of m + 1
billiard balls that you assure me are each solid colored. I keep my eyes
closed and put one of them in my large left pocket. Since I now count m
balls on the table, I know they are all of the same color, which, for purposes
of clarity we may call 'carmine.' I now, still with my eyes closed, take one
of the carmine balls from the table and place it in my large right pocket. I
remove the solid-colored ball of unknown color from my left pocket and
place it on the table with the other balls. I again have a set of m solid-
colored balls, so I know they are all of the same color. Since those remain¬
ing on the table were carmine, they must all be carmine. Thus, when I
remove the known carmine ball from my right pocket and place it on the
table, I am guaranteed that all m + 1 balls are carmine (completing the
proof of condition 2)
• By natural induction, we may now deduce that, for all natural numbers n,
any set of n solid-colored billiard balls are all of the same color.
"Which is a bit troubling, since just yesterday I saw a set of three billiard balls,
two of which were solid black and one of which was solid white. Of course, as I
said earlier, this 'proof' is fallacious, because the proof of condition 2 fails when
m + 1 equals two. Clearly, if there were a set of two balls on the table, and I put
one in my left pocket and the other in my right pocket, there would be none left
on the table to ensure the commonness of color. This little problem taught me how
careful we have to be in the proof of condition 2, to see that there is no value of m
for which our argument breaks down."
Strategies for Problem Exploration 77
Your discussion was astute, but I think it would have been clearer had you
specified that you were taking p(j) as the sentence 'Any set of j solid-colored
billiard balls all must have the same color/ When you're being so careful, you
should not assume that your listener is willing to supply the information you have
omitted."
"I always expect something good from my listener, especially when it's you
or Mom."
"So tell me, Nick, what did Allard puzzle you with?"
He said: 'Certain descriptors define natural numbers. For example, the least
natural number defines one. Suppose we specify some existing dictionary as the
standard dictionary. I ask you to consider the set S of all natural numbers which
can be defined by descriptors of 25 words or less, each of which appears in the
standard dictionary.
" 'Any dictionary, and specifically our standard dictionary, must contain a
finite number of words. There will therefore be only a finite possible number of
descriptors of 25 words or less, each of which appears in the standard dictionary.
Thus S is a finite set.
" 'We know that the natural numbers are well-ordered; that is, we know that
any nonempty set of natural numbers has a least element. In fact, the well-ordering
principle is formally equivalent to the principle of natural induction. Because of
the finiteness of S, its complement S' is non-null. Therefore, by well-ordering,
S' has a least element, which we may call z.
" 'Thus z is the least element of the natural numbers that has no descriptor of 25
words or less, each of which appears in the standard dictionary.
" 'Assuming that we have taken a standard dictionary of reasonable size, it
will contain each of the words in the descriptor above, yielding a gruesome
paradox. We have found a descriptor of 25 words or less, each of which appears
in the standard dictionary, for a natural number that, by definition, can have no
such descriptor.'
"This guy Allard is weird. After he presented this paradoxical situation, he
stood there laughing a weird hollow laugh that gave every one of us the chills."
"Do you see how to resolve Allard's paradox?"
"No. Do you know anyone who knows?"
Polly Sue's voice drifted in from the kitchen, where she was preparing a
delightful snack of fresh morels sauteed in butter. "Give a careful definition of both
'descriptor' and 'define,' and there will be no paradox."
"Do you understand what Mom means?" asked Nick.
"Let's see. If you allow descriptors to be random sequences of words and
mean by 'define' to establish a one-to-one correspondence between these random
sequences and natural numbers, there will be no paradox, because the descriptor
that was put forward as creating the paradox will have already been used. If, on
the other hand, you insist that descriptors have meaning, then definitions must
uniquely specify natural numbers within the language of our mathematical sys¬
tem. In this case, the descriptor that was put forward as creating the paradox does
78 THE ART OF PROBLEM SOLVING
not define a natural number, because it uses the word 'descriptor/ which is neither
a prespecified undefined term nor a defined term in our mathematical system.
Since it is not a definition, there is no paradox.
"Allard's discussion gets at critical issues in mathematical logic. You may
want to bone up on that discipline for your own edification. I'd like to meet Allard
sometime."
"If you guys don't hurry," warned Polly Sue, "there won't be any left."
Old King Cole's grandfather had been a slick cookie. He admired mathemati¬
cal acumen and had decided that his beautiful daughter should marry the best
logician in the country. He had invited the three best logicians he could identify
in his kingdom to a timed contest that would determine who would be given his
daughter's hand in marriage. He had told the three men that he had a box
containing exactly three red hats and two white hats and that he would blindfold
them and place a hat on each man's head; he would then direct his chamberlains
to remove the three blindfolds simultaneously, so that each man could see the color
of the hat on each other man's head but not the color of the hat on his own. The
first man who could correctly identify the color of the hat on his own head through
logic would win the hand of the princess if he could explain that logic flawlessly.
An incorrect guess or faulty logic would be penalized by 30 years of servitude
in the stables.
Old King Cole's grandfather had a commitment to absolute fairness, so he
decided to symmetrize the problem by placing a red hat on the head of each of the
three logicians. Thus it was that when the chamberlains simultaneously removed
the three blindfolds, each of the three men saw two red hats.
Cole's grandfather said, "I am ready to hear a cogent explanation but be¬
ware—the shovels are ready."
Jason approached the throne and spoke as follows: "I said to myself, 'Suppose
I were wearing a white hat. Each of the other two men would be seeing one red
hat and one white hat and would know that their own hat could not also be white,
because the other contestant would be seeing two white hats and would instanta¬
neously know that his own hat must be red.' Since neither of my esteemed
opponents spoke within a few seconds, I realized that my hat must be red."
Old King Cole knew that this problem was old hat to even the lowliest
peasant in his domain, let alone to the logicians and scholars. Thus it was that his
own problem seemed to have no solution. He had a handsome, young, brainless
son, for whom he wished to find the most intelligent maiden in the kingdom by a
process as much like his grandfather's as possible, yet in some important way new.
His scouts, knights, and fiscal agents reported to him that in fact his kingdom
included 17 incredibly beautiful and amazingly intelligent young women who
were stars in the world of logic. He invited all 17 to his court, but he was not at all
sure what he was going to do with them.
Strategies for Problem Exploration 79
Can't I keep them all?" his son asked. Old King Cole realized that he had
better lock his son in the east wing of the palace until the selection process was
completed.
Cole consulted with Professor Silbert, who had been his father's teacher and
then his own. "What should I do?" he asked.
"Do as your father would have suggested, if he had lived," Silbert replied.
"Tell the young ladies that you have a large quantity of red hats, but only 16 white
ones and that you will do much the same thing that your grandfather did, by
putting a hat on each of their heads while they are blindfolded and then having
the blindfolds simultaneously removed, with the prize that the first young lady
correctly to identify the hat color she wears will be the bride of the prince if she
can cogently justify her reasoning. The penalty for error or noncogency will be 17
years servitude in the scullery."
Old King Cole wasn't sure exactly how Silbert's suggestion would work, but
it was far better than anything he could think of, so he carried it out exactly as
suggested. He watched impatiently as the young ladies were gathered, blind¬
folded, hatted, and simultaneously unblindfolded.
Minerva of the Double Helix was by far the wisest lass to have graced Cole's
kingdom in many generations. She knew that Cole was going to put red hats on
everybody before she opened her eyes to observe 16 red hats. "Telling him, 'Wom¬
an's intuition' will earn me hard time in scullery," she thought. "If only this had
been a simpler case such as the traditional three, or even four. I could do it with four
contestants and three white hats by saying, 'If I had a white hat, the other three
smarties would see that the problem reduced to the traditional one and solve it in
a flash, because they can leave me and my hat out and just look at each other.' "
"Holy moley," shouted Minerva, raising her hand just before Athena of the
Triple Catenary.
"You may approach the throne," said Old King Cole merrily.
"Will a mathematician be present to judge the cogency of my solution?"
"I know everything," said a voice behind the throne. "I am Silbert."
"I have an elegant inductive proof that leads to the inescapable conclusion
that I am wearing a red hat," said Minerva saucily. "It is an induction on the
following proposition: if you are in any group of n maidens, all hatted from a set
containing at least n red hats but exactly n -1 white hats and in which each maiden
is a brilliant logician, it is deducible after the unmasking that if you perceive only
red hats and no other maiden raises her dainty hand within a few seconds, then
you are also wearing a red hat.
"The induction begins not with n = 1, but with n = 3. Will you allow 'classical
case' as a proof for n = 3?"
"Of course," said Silbert, already in love with Minerva of the Double Helix.
"Let us then proceed. Let k > 3 and let us assume (by the inductive assump¬
tion) that in any group of k maidens, the proposition holds. We now consider a set
of k + 1 maidens. Each reasons that if she were wearing a white hat, the other
maidens would eliminate her and her hat from consideration, reducing the prob-
80 THE ART OF PROBLEM SOLVING
lem to a set of k maidens. Each would raise their hands as fast as their nubile young
reflexes would allow. Hence, since no hands shot upward, each can now deduce
that she is wearing a red hat, and would attempt to raise her hand as fast as Athena
of the Triple Catenary tried. My nubile young reflexes are stronger, faster, and
better."
"And you shall have my son. Heaven protect you. You are one sharp cookie.
Send a courier to tell my son that we have found him a bride."
"And how would you modify your proof for a group of n young men?" asked
Silbert.
"Mutatis mutandis," said Minerva of the Double Helix, blushing slightly.
"And, dear future father-in-law," she added, "wait until you see what I make of
your son within a year."
It is evening in the great hall of the most brilliant and most feared man in all
of China. A helpless victim knelt before him, begging for his life.
"I'll give you a sporting chance," said the ornately clad gentleman on the
golden throne. "You need only solve a trifling problem in mathematical reasoning
and your wretched life will be spared. Of course, if you fail, your death shall be
slow and hideous. Do you agree?"
The wretched man stared up. "I have little choice, oh great noble Doctor."
"Bring in the table," commanded the regal figure, as five slaves leaped to
obey his command. Within moments, a gold-encrusted ivory table was set before
the groveling man. On each comer of the table there was a tumbler.
"Blindfold him!" commanded the man in the golden robes. "Now invert one
or more of the tumblers, as your hearts dictate." His slaves obeyed.
He addressed the terrified man on the floor. "Your challenge is as follows:
While blindfolded, you may touch any two of the tumblers at the comers of this
ivory table to determine their orientation; you may then change the orientation of
none of them, one of them, or both of them; if all four tumblers, after your move,
are oriented in the same way, all top up or all bottom up, the great gong of Genghis
Khan will sound immediately.
"If no gong sounds, you may be assured that the four tumblers are not
oriented all up or all down. The table will now be spun a random number of
90-degree rotations without disturbing the orientation of the tumblers. Again, you
will be permitted to touch any two tumblers, determine their orientation, and
change the orientation of none, one, or both. What is the minimum number of turns
that you will need to guarantee that the gong will sound?"
The men felt the hand of Death upon his heart. Clearly, the randomness of
the spin made it impossible to guarantee that he would ever be able to touch the
same pair of glasses on any two given turns. Perhaps that was the insidious plan.
Perhaps "the minimum number" did not exist.
Strategies for Problem Exploration 81
13.5. Proof
Proof is the connecting tissue holding the body of mathematics together, but
there are many kinds of proof. Some proofs establish algorithms such as the proof
which yielded the technique for sounding the gong in section 13.4. Some proofs
are based on suppositions, others on powerful techniques such as natural induc¬
tion. Some proofs establish the existence of a mathematical entity, while others
show uniqueness. Some proofs construct an elusive mathematical object; others
82 THE ART OF PROBLEM SOLVING
show that no such object may exist or that something is impossible. Some proofs
work from given information to a desired conclusion, and some begin by denying
that conclusion and deducing a contradiction of a result that has already been
established or of a posited hypothesis. Some proofs are achieved by working
backward, and some, by working both ways from the middle. Some proofs are
done in one mighty stroke, and some are done by cases. Some proofs are estab¬
lished within months of the time that a proposition is put forward; others elude
the finest mathematical minds for centuries. Some proofs are painstaking and
messy, others are crystal clear and elegant.
Mathematics is a tapestry woven with many varieties of proof. Skill in all of
them is among the most important mathematical exploration strategies there are.
From the day that a child first realizes that the hidden face of a fair coin is a head
when the face he sees is a tail to the day that he basks in the glory of a valid proof
of Fermat's Last Theorem, exposure to proof and experience with proof should be
a continual part of his mathematical experience.
Proof and disproof are ends and beginnings. They should be looked upon
with love but not with awe.
Afterword
ALFRED S. POSAMENTIER
Troblem solving has been the cornerstone of mathematics instruction from the
earliest times. The National Council of Teachers of Mathematics has included it as
a key component of its Agenda for Action (1980) and its Curriculum and Evaluation
Standards for School Mathematics (1989). It is, as a matter of fact, listed as the first
standard! These latest efforts seem to have placed problem solving as a topic
deserving special attention. Fitting this new role is the categorization of problem¬
solving techniques or strategies. What most mathematicians have for years done
intuitively is now being brought to the fore in a "cookbook" type of way. This will
certainly be of use to the average student, since one important tenet of problem
solving is to recall similar circumstances from the past and use these experiences
to solve the present problem.
For purposes of this presentation, we shall consider a few interesting strate¬
gies (See Introduction, p. viii). In each case the method was selected since it is
generally a bit "off the beaten path" and not naturally expected to be part of a
typical student's problem-solving approach. Students have a tendency to try
straightforward methods and rely heavily on the methods shown in class. Thus
more traditional problem-solving techniques frequently form the basis for their
83
84 THE ART OF PROBLEM SOLVING
Find the area of the shaded region, when ABCD is a square of side
length 1, and the two arcs are quarter circles of radius 1, with the center
at opposite vertices of the square.
A B
A B
Figure 2.2
Unconventional Problem-Solving Strategies 85
Experience shows that this is by far the most common method but not the
most elegant!
A somewhat more elegant procedure, and one far more infrequently shown
to be used by students, involves a "complementary approach." Here the student
seeks the area of half of the nonshaded region, say the region ABD, by subtracting
the area of the quarter circle BDC = n/4 from the area of the square = 1. Then
doubling this area = 2 (1 - rc/4) = 2 - tc/2, and subtracting it from the area of the
square to get: 1 - (2 - tc/2) = k/2 - 1.
A B
Figure 2.3
The gifted student has shown an even more streamlined version of this
method. He or she would find the area of quarter circle ABD and quarter circle CDB
and deduct the area of the square from the sum of the areas of the two quarter cir¬
cles. This requires "seeing" that the regions I and III are each considered once, while
the region II is used twice. This taking the square from this sum leaves region II.
That is.
A B
86 THE ART OF PROBLEM SOLVING
These three methods for solving a simple problem afford an opportunity for
observing various thinking styles. They are shown here to sensitize the reader to
various levels of thinking on the part of the student. This must be kept clearly in
mind when considering the problem-solving methods discussed in the rest of this
chapter. Naturally, modification of the presented material must be made by the
teacher in discussing these methods with a secondary school class. Teachers play
a key role in interpreting these methods in the most appropriate fashion to their
students. There is no substitute for a fine teacher, especially not in the all-important
topic of problem solving.
Working Backwards
If the sum of two numbers is 12 and their product is 4, find the sum of
their reciprocals.
Typically, a student will solve this in the traditional way, arriving at the
answer 3 after some tedious calculations. Consider the following rather straight¬
forward and somewhat inelegant solution:
x(12-x) = 4
x2 - 12x + 4 = 0.
x = 6±4^2.
X 6 + 4V2 6-4V2
y 6 - 4^2" 6 + 4V2
1 _ 1 _ 3-2a/2
x ~ 6 + 4V2 ~ 2
1 _ 1 _ 3 + 2^2
y ~ 6 - 4V2 _ 2 '
so that
1 1 3 - 2V2 3 + 2V2 6
- + - = - +--— = ^ = 3 (Answer)
Using a backwards strategy, one asks. What are we looking for and how can
we get back to the starting point from there?
We seek to find
1 1
- + —
x y-
This could have come from the sum of these fractions:
x+y
xy
1 1 x+y 12 „
x y xy 4
Clearly, this backward strategy was much more elegant and efficient than the first
method. A reverse strategy is certainly not new. It was considered by Pappus of
Alexandria about a.d. 320. In Book VII of Pappus's Collection, there is a rather
complete description of the methods of "analysis" and "synthesis." T. L. Heath, in
his Manual of Greek Mathematics (Oxford University Press, 1931, pp. 452-453),
provides a translation of Pappus's definitions of these terms:
Consider the problem alteration where x + y = 12 and xy = 40. Here the values
of x and y are imaginary, which is only discovered when a straightforward method
is used to first find the values of x and y. The backwards method enables a solution
to the problem without first finding x and y. This may be a disadvantage in
understanding the mathematics in the problem situation but certainly solves the
given problem more efficiently.
A second example of this backwards or reverse problem-solving strategy
shows the wide variety of application possibilities.
Using only one 5-liter and one 11 -liter bucket, how can 7 liters of water
be obtained in the larger bucket?
The solution requires considering the desired result and then working back¬
wards (analysis) to the starting point. The reversal of this analysis (synthesis) will
then provide a correct procedure to solve the problem.
Analyzing Extremes
affect the problem s solution but make the problem considerably simpler. Con¬
sider some examples:
The distance between two concentric circles is 10. What is the differ¬
ence between their circumferences?
GEOM
Since the size of the circles is not given, let us assume that the smaller circle
is extremely small, so small that it becomes a point. Then the distance between the
circles, 10, now becomes the radius of the larger (or remaining) circle. The circum¬
ference of this larger circle is then 2071, which differs from the "zero circumference"
circle by 20tc
Once again this exhibits how changing variables to suit a simple solution is
helpful in problem solving. The statement of the problem implied an indepen¬
dence from the size of the circles. This can be shown as follows.
The differences of the circumferences is
Thus the differences between the circumferences is 2rc times the difference
between the radii. This independence from the size of the circles leads to a very
humorous problem.
Unconventional Problem-Solving Strategies 91
It is known (and can easily be proved) that the sum of the vertex angles
of any pentagram is the same. What is that sum? (See Figure 2.9a.)
Since the type of pentagram was not specified, we can either assume the
pentagram is regular or that it is one which is inscriptable in a circle (i.e., all the
vertices lie on a circle, as in Figure 2.9b). In the former case, where the pentagram
is regular, all angles are congruent and a solution is readily at hand—a number of
methods can be used.
In the latter case, we notice that each of the angles is now an inscribed angle
of the circle and so has half the measure of the intercepted arc. Consequently, we
get the following:
mZA = l/2 mCD; mZB = l/i mED; mZC = Vi mAE; mZD = Vi mAB; mZE = Vi mBC
[CD is arc CD]
mZA + mZB + mZC + mZD + mZE = Vz( CD + ED + AE + AB + BC )
92 THE ART OF PROBLEM SOLVING
That is, the sum of the vertices is one half the circumference of the circle, or 180°.
Again there was no loss of generality by allowing the nonspecified penta¬
gram to assume a more useful configuration. Yet this change made the solution
much more manageable (and solvable).
An enthusiastic student, having difficulty finding a solution to the following
problem, recently asked to have his attempted solution reviewed.
The raindrops in AB are the last he will hit when reaching P. The parallelogram
ABCD represents (in two dimensions) the raindrops he will come into contact with
throughout his journey from C to P. Were he to go infinitely fast, he would sweep
out an area (again, in two dimensions) of the rectangle QPCD. Part b of the problem
is then reduced to comparing the area of ABCD with the area of QPCD. Since they
share the same base CD and have the same altitude PC, their areas are equal.
Thus, for the case where the rain is falling uniformly (at the speed of the run¬
ner) and without wind, the runner's speed does not affect the front surface, only
the top. Therefore, traveling faster in the rain makes for an overall drier person!
The analysis of extremes, used with success in this problem, is a very
powerful source for asking a key question. Many everyday (nonmathematical)
problems sometimes can be more easily analyzed by posing a question about the
extreme situations. For example, when a question in economics is posed, a consid¬
eration of the behavior of the variables under the extreme conditions (both high
and low) is useful in understanding and ultimately answering the question.
(x + y)8 = V x8 + V x y
v°y V
7
1
,
+
©
2
V V
*V +
© x5y3
3
v y
^
+
© x4y4
4
v y
L
+
5
v
3 5 ,
xy +
V x1y6 +,
v
6 7
v y
7 ,
xy +
8
V ©
The sum of the coefficients is
f8 +
f8l +
V+ + V+ V + + V + V = 1 + 8 + 28 + 56 + 70 + 56 + 28 + 8 + 1 = 256
0 i 2 3
^ y
4
k y
5
^ y ^ y
7
^ y
6 8
y
v y ^ y ^ y
For the second solution, let x = y = 1 in the first equation of the above solution.
Then
/0\
V 8
(x + y)8 = (l + l)8 =
0 2 3 4 5 7 8,
v y v y v y v y v y © v y v y
The way a problem is physically presented and then rearranged (or modi¬
fied) for solution is another way of changing variables (without disturbing the
generality implied in the problem statement) to facilitate solution.
Consider the following:
GEN A bug perched on corner B of a solid box chooses to reach the di¬
agonally opposite corner H by the shortest route along the faces of the
GEOM
box (AB * AE.) Which path does it take?
The solution to this problem demonstrates the role that imagination as well
as insight plays in problem solving. Open the box along several edges to flatten
it out. The illustration below is one possibility. BH is the shortest path between
B and H.
F G
B c
\X
S
V
X
X
xX
X
X
A X
X D
X
'x X
X
X
X
X
X
X
X
X
X
X
X
X
X
E H
F G
Figure 2.12
Unconventional Problem-Solving Strategies 95
The solution to some problems hinges on the solver asking the right kind of
question, or "key question," to open the door to a clever solution to the given
problem. Not only "crazy people" talk to themselves—so do people trying to solve
a mathematics problem. Consider the following problem:
Suppose you have 7 pairs of blue socks all alike, and 7 pairs of red
socks all alike, scrambled in a drawer. How many socks must be taken COMB
out at the same time (in the dark) to be certain of getting a matching
pair?
As you begin to tackle this problem, you must ask the right type of key
question. To be certain of getting a matching pair, we must assume the worst
possible luck. We therefore ask. What is the greatest number of socks, each of a
different color, which can be taken from the drawer? Since there are only two
colors, the answer is two. Thus the third sock must match one of the first two, so
we would have to take three socks from the drawer to be sure of having a matched
pair (regardless of color). (See Chapter 4, p. 123)
We shall consider a second problem which also requires asking the right key
question.
To solve this problem we search for the right question. We ask. How many
letters would be needed to give each letterbox 2 letters? Although these 40 letters
could be distributed so that one box has more than 3 letters, we must assume the
worst possible luck, namely, that each letterbox has exactly 2 letters. Then the 41st
letter would increase the contents of one of these letterboxes to 3 letters. Therefore,
41 letters are needed to be certain that at least 3 letters are put in one box.
To further reinforce the notion of asking the key question, we shall consider
the following problem:
96 THE ART OF PROBLEM SOLVING
Success with this problem once again depends on your asking the right
questions. Assuming the worst possible luck, what is the greatest number of
ribbons which can be taken from the drawer to have only 3 different colors? We
would take all the ribbons of the most plentiful colors: 6 red ribbons, 7 green
ribbons, 9 yellow ribbons.
We can have 22 ribbons and still only have 3 of the 4 colors represented.
However, the 23rd ribbon must be of a different color! Thus, after 23 picks, we
must have at least one of each color. (This is not to say that this could not have
been achieved on the first four picks, but it was not to be assumed.)
A geometric problem also can have a solution which hinges on asking the
right key question.
O O
O O O
O O O O
Figure 2.13
Here, the key question is, "How many checkers would remain unmoved if the
direction were to reverse?" or "Which checkers are positioned independent of the
triangle's direction?"
The following 7 dark checkers can remain unmoved, while the 3 light ones
would move as indicated.
Unconventional Problem-Solving Strategies 97
O o
(8) o
Figure 2.14
The key to the solution to some problems lies in the question one asks
(oneself) in the process of looking for a solution. Asking the right question alone
may not necessarily yield the solution immediately. Other techniques could also
be required. However, stepping back from the problem, reconsidering it from a
different perspective, and then asking the right (often unusual) key question, may
prove to be the key to a solution. Only practice will ensure success with this notion!
Seeking Complements
Notice that sometimes the "key question" is one that asks the "opposite" or
complement of what we seek. This is a helpful approach but does not always lead
to the right question. The next example will demonstrate how the key question
again asks for the opposite of what is being sought.
It is very easy to get bogged down with a problem like this. Instead of trying
to find the water content of each day's mixture, simply look to ask the key question:
"How much water has been added to the mixture each day?" This key question
asks for the "complementary substance" to the one required.
In a similar vein, the following problem's solution becomes simple when the key
question is asked.
(^3^) Instead of counting the number of games the winners must play, ask the right
question (looking for the complementary component): How many losers are there
in the tournament? Since there is only one winner, there must be 35 losers,
requiring 35 games for the tournament.
How would the solution change if the problem required two losses to
eliminate a player?
The key question to open the door to a problem's solution might be the
indicator of the key variable to be found. Such is the case in the following problem.
Two trains, 200 miles apart, are traveling toward each other on the
same track. One travels at 60 mph, while the other travels at 40 mph.
GEN
A superfly perched on the front of the slower train flies toward the faster
I ALG train at 240 mph, and when it reaches the train, it instantly turns around
and flies back toward the slower train (again at 240 mph). When it
reaches the slower train it again turns around and instantly flies
towards the faster train. This continues until the two trains crash and
smash the fly. What was the total distance that the fly traveled?
To try to find the distance, directly, that the superfly traveled altogether would be
complicated, at best. The key question here is, What is the time required for the
fly's travel? Since we already know the speed of the fly, the distance traveled
would simply be the product of the time and the speed. The fly's time of travel is
actually the time required for the two trains to collide.
Unconventional Problem Solving Strategies 99
The last two problems show solutions that made use of asking key questions,
which (at first glance) do not directly aim at the solution sought. The questions are
directed at finding some other information within the problem situation. With this
additional information, we illuminate the problem further so as to render a
solution to the original problem quite simply. We see from these examples that the
solution to a problem can sometimes be best found by a somewhat circuitous route!
In this chapter we have considered problems that have lent themselves to
some rather unusual solution methods. The problems were specially selected to
show these problem-solving strategies in a dramatic fashion. Imitation is often
seen to be a useful tool of problem solving. We sometimes recall problem-solving
procedures without actually remembering the context in which they were used.
Students who are shown these examples (and others like them) should get a "feel"
for these methods and thereby begin to develop a fine reserve of problem-solving
strategies to use in their study of mathematics as well as in everyday life. Often,
problem-solving strategies in the mathematics context can be interchangeable with
their use in everyday life. Students should begin to see problem solving not just a
means to an end, but rather an end in itself.
3
Interest Grabbers
Exciting Motivational Problems
With Punch and Personality
STEVEN R. CONRAD
Introduction
101
102 THE ART OF PROBLEM SOLVING
asking ourselves questions not directly asked of us, and we must reflect on the
significance of these activities. In learning a foreign language, "immersion" is the
most successful approach. In much the same way, the route to success in mathe¬
matical problem solving is immersion—in problem-solving activities.
George Polya taught us that certain techniques are in common use by
successful problem solvers. The very first section of Polya's classic How to Solve It
(Princeton, NJ: Princeton University Press, 1945) is his list of four steps for solving
problems. Polya's second step, "Devising a Plan," is the longest step. It concen¬
trates on finding a "connection between the data and the unknown," and it begins
with the questions "Have you seen it before?" and "Do you know a related
problem?" Indeed, the more experienced we are at problem solving, the better we
are able to solve new and more difficult problems. Successful problem solvers
carry an arsenal of weapons; they have already seen or solved a wide variety of
problems and have thereby learned a wide variety of techniques.
If you want to become a successful problem solver, the most important step
is to immerse yourself in solving problems. Athletes train for many years—and
continue to train throughout their careers. In much the same way, the success of
students who participate in math contests (mathletes) improves when time is spent
solving problems.11 am reminded of a story told to me some years ago by a friend.
When my friend, who had prided himself on his mathematical ability, was given
his first homework assignment in math class at college, he whizzed through all the
problems save one. Not wanting to hand in an incomplete assignment the very
first day, he was determined to solve the remaining problem. Day turned to night,
and then back to day. He'd spent all night on the problem and had not solved it.
The next day, in class, the professor asked if anyone had any question they wished
to ask about any homework problem. My friend raised his hand and asked, "How
do you begin problem 6 on page 3?" The professor opened his book, turned to the
page, and read the problem to himself. "Problem 6 on page 3?" he inquired.
"Cannot be solved," intoned the professor. My friend was visibly shaken. "What
do you mean, 'Cannot be solved'? I spent all night working on that problem." The
professor looked at him kindly and said, "Well then, you must have learned a lot!"
What are all ordered pairs of unequal positive integers (a,b), with a > b,
which satisfy — = — + —?
9 a b
Solution 1
What does the equation say about bounds on the possible values of a and b?
Since a > 0 and b > 0, neither can equal 9. Let's prove this. Suppose b <9. Then l/b>
1/9 and 1/a < 0. Since a > 0, this is impossible. Since a > b, the values of a and b
must satisfy a > b > 9. Let a = 9 + x, and let b = 9 + y, with x > y > 0. We get 1/9 =
1/(9 + x) + 1/(9 + y). Clearing fractions and simplifying the result, we get xy = 81.
Since x > y, (x,y) = (81, 1) or (27, 3). Finally, {a,b) = (9 + x, 9 + y) = (90, 10) or
(36,12).
The example is of a general form, as is its solution. We can summarize by
saying:
1 1
All solutions of — = - satisfy xy = n2.
n n+x n+y
Solution 2
Let's concentrate on the possible values of b, the smaller of the two variables.
As in Solution 1, b > 9. If both a and b were bigger than 18, the sum of their
reciprocals would be less than 1/9. If both a and b were smaller than 18, the sum
of their reciprocals would be greater than 1/9. Therefore, one is smaller than 18
and the other is bigger than 18. Since a > b, a > 18 and b < 18. Concentrating on b and
combining requirements, 9 < b < 18. The possible values for b are 10, 11,..., 17.
Let's try all 8 possibilities and take note when a is integral. We see that if b - 10,
then a = 90; and if b = 12, then a = 36.
Another way to concentrate on the value of one variable—and let the value
of the other follow—is to solve for that variable. Let's use this approach for the
previous problem. Clearing fractions, ab-9a + 9b. Solving for b (or a, if you prefer),
you get b = 9a/{a - 9). Continue dividing to get b = 9 + SI/{a - 9). In order for b to
be an integer, a-9 must be a divisor of 81. Since a > b, a - 9 = 81 or a - 9 = 27 (using
any other divisor of 81 would make a < b). The two solutions readily follow.
At first glance, the next problem looks nothing like the previous one. The
problem statements appear very different—but we have another opportunity to
concentrate on one variable. An equation whose solutions must be integers is
called a Diophantine equation. Try to solve this Diophantine equation:
104 THE ART OF PROBLEM SOLVING
Solution
Before trying to focus on one variable, use any information readily available.
The roots are A and B and their sum is -B/A, so we have an equation in A and B:
A + B = -B/A.
-A2
Solving for B, we get B = ———
1
Dividing, B = -A + 1 - —-
A+1
1-1 +
1
9 “ a b ’
- = 1 + l;and
9 ~ c
_2_ _1
+
1
37 ~ e f '
n- 12
What is the least positive integral value of n for which is a
5n + 23 NUMB
non-zero reducible fraction?
Solution
5n + 23 _ 83
— 5 + .
n - 12 n - 12
ALG
Factor completely: (x + 1 )(x + 2)(x + 3)(x + 4) - 3.
Solution
Be careful not to let the previous example mislead you. Often, a good
substitution makes a problem simpler but not yet easily accessible. Sometimes, the
difficulty can be resolved by considering the general case instead, because in some
situations, the general case is easier to handle than a specific case!
Solution
The thought of clearing fractions and solving the resulting equation should
be abandoned with all due speed! Instead, let's substitute where obvious and take
a look at the resulting equation. This does introduce new variables, so things may
not improve (but it's better to have tried and failed than never to have tried at all).
Let a - 2x3 - 3x2, b = x + 1, c = 3x3 - x2, and d = 5x- 13. Substituting^ + ^ - SctA
a-b c-d
Wouldn't those numerators look simpler if instead we could add the denomi¬
nator to the numerator in each fraction, or subtract the denominator from the
Motivational Problems 107
If asked how to drive from New York, in the northeastern United States, to
a small town in the southwest, one would not begin with a map of New York:
There are just too many routes that leave New York. Where would you begin?
Begin at your destination and try to find the best way back to New York!
In the Unlucky Lottery, all the prizes are powers of $13 ($1, $13, $169,
etc.), and the total prize money to be given away is $1 million. What is ARITH
Solution 1
The prize amounts are $1, $13, $169, $2,197, $28,561, and $371,293. To give
the least number of prizes, start with the largest prize: award as many of these as
possible before giving out smaller prizes. There can be at most 2 prizes of $371,293;
and then $257,414 would remain to be awarded. We could then give away 9 prizes
of $28,561. This would leave $365, which could be most efficiently given away as
2 prizes of $169, 2 prizes of $13, and 1 prize of $1—for a total of 16 prizes.
Is there a way to do all this more simply—or at least to describe the procedure
more easily? Yes! Use thirteen as the base of numeration, as in Solution 2:
Solution 2
In base thirteen, the number 1 million is 290221, meaning there are 2 prizes
of $371,293, 9 prizes of $28,561, 2 prizes of $169, 2 prizes of $13, and 1 prize of $1.
The least possible number of prizes is 16.
ARITH How can $1 million in prizes be given away in the Lucky Lottery, if all
prizes are powers of $11, and if no more than 7 people can receive the
same prize?
Though there are several ways to sum the series in the next problem, none
of them is obvious. Give yourself time to puzzle over the following problem for
quite a while before looking at the solution. The best approach to solving this
problem is to first investigate a very similar series that you can sum. You'll enjoy
this problem.
Solution
Let's sum n terms of 9 + 99 + 999 + ..., instead. We can rewrite this series as
(10 - 1) + (100 - 1) + (1,000 -1) + ... The sum of n addends will be 10 + 100 + 1,000 +
Motivational Problems 109
10,000 + ... -n. The powers of 10 form an n-term geometric series whose sum is
10 -10 , mn + 1 _ in
9 so the sum of n terms of 9 + 99 + 999 + ... is--- - n. But what
about the original problem?
Since 7 is 7/9 of 9, the sum of n terms of the original series is
7ricr+1-io \
n -9 n -10).
9 9
?
IT'10'
It is interesting to note that, for n < 9, the sum is 7(123456 ...), where the last
digit used inside the parentheses is the value of n. (When n > 10, this method does
not work.)
1. What is the sum of the infinite series 1/3 + 2/32 + 3/33 + .. ? ALG
1 1 1 _L J_
2. What is the value of 22 x 44 x 88 x 1616 x ... (2n)2" ? ALG
If you had 2 coins made of gold, and you had 2 coins over 20 years old, you
could have 2,3, or 4 coins altogether—the total depends on how many of your gold
coins are over 20 years old. Any time you count, remember to account for the
possibility of duplicates when objects can be categorized in more than one way.
There's an analogous result in set theory: For two sets, if n(S) is the number
of elements in set S, then n(A u B) = n(A) + n(B) - n(A n B). The next problem arose
from the observation that, although V9~ = Vi" + Vl", and VlfT = V9 x 2 = 3^2 = 2V2" + ©
V2" = VfT + V2, the number VTcT cannot be written as a sum of square roots of integers.
Solution
NUMB 1. How many positive integers less than 1,000 are divisible by neither 5
nor 7?
NUMB 2. For how many integers x between 20,000 and 60,000 is xthe square of
an integer or the cube of an integer?
NUMB 3. How many integers from 1 to 100 inclusive have no repeated prime
factor?
In solving difficult algebraic equations of the types given in contests, the key
step often involves combining the equations in some nifty way that makes the
solution apparent. Recall that an algorithmic way to solve a two-variable linear
system is to multiply each equation by its own multiplier, then add the results.
This linear combination of the two equations is often all we need in order to solve
the system. Though a calculator trivializes the next problem (since the problem's
only difficulty is the size of its coefficients), the original wording has not been
Motivational Problems 111
revised. It is hoped that when those who love mathematics see the symmetry in
the equations, they'll eventually seek an "artistic" approach.
With each major curriculum change, we gain something big and lose some¬
thing small. There's a beauty in some parts of mathematics that is best witnessed
with technology. And there's a beauty in other parts that will be lost. Perhaps the
next problem will remind us of an era in which we could find much cleverness in
parts of arithmetic.
Solution
r2 n
(*,y) =
3' 2
1. Solve for (x, y): 713x + 637y= 4,164 and 637x+ 713y= 3,926. ALG
2. If each of four numbers is added to the average of the other three, the
ALG
respective sums are 28, 32, 40, and 44. What are the four numbers?
nor an inequality, the solution involves both. There is no other problem quite like
this one—it's a one-problem course in trigonometry!
Solution
If there is an arc sec y, then I y I >1. Next, since tan arc sec y = ±Vy2 - 1, we
know that cos arc sin (±Vy2 - 1) = ±V2 - y2, so I y I < fl. We know that the value of
esc arc cot (±V2-y2) is ±V3 - y2. But the cosine of an angle cannot exceed 1 in
absolute value, so in arc cos (±V3 - y2), I y I must equal V2; and arc cos (±1) = 0 or
n. Then, cot arc cos (±1) is undefined, and sec arc csc(undefined) is ±1. Finally, the
value of sin arc tan (±1) is ±v2/2.
TRIG
1. Solve the equation 2sinx = 5X2 + 2x + 3.
TRIG 2. Solve the equation cos7x+ sin4x= 1.
ALG
3. Solve the equation x = Vx2".
Years of experience solving math problems has two effects: It ages you (the
years, not the experience), and it gets you started with the right strategy more
quickly (the experience, not the years). When solving a circle problem, a strategy
that often works well is to draw one or more strategic radii. An additional strategy
useful in more difficult circle problems is to draw one or more lines through the
center, perpendicular to another line. The actual technique varies with the problem
at hand, so try your hand at the three problems below:
Solution
Remember that a sector with radius r and central angle a (in radians) has area
a. To evaluate r, draw the lines shown and use the Pythagorean theorem.
In all three problems above, the ratio sought turns out to be 8/571. Is it always
8/57t? Not at all! Let's consider the case where exactly two of the vertices of the
square lie on the arc, as in the first two drawings. By drawing a radius and a center
line perpendicular to the far side of the square, show that if the central angle is 20
(in radians), then the ratio sought is always 4/[(5 + 4cot0 + cot2©)©].
(Note: In the first two diagrams, 0 = 7t/2 and 0 = 7i/4, respectively.)
Construction problems are not as popular as they once were. But construct¬
ing figures that satisfy a given condition is really an exercise in correctly remem¬
bering and applying the theorems learned in geometry, so constructions are
among the best (and also among the most difficult) problems in geometry.
Draw two circles intersecting at two points so that the degree measure of
the arc cut off on one circle is twice that of the arc cut off on the other.
Figure 3.3
114 THE ART OF PROBLEM SOLVING
Solution
Construct a triangle given the lengths of all the medians. (Note: As before,
begin by drawing the completed triangle. Then, draw in the medians. To complete
the construction, you'll have to construct a parallelogram, given two sides and a
diagonal. Which parallelogram? That's the difficult part, but the secret lies in the
figure you've already drawn—so study it carefully!)
What are three numbers whose sum is 13, whose product is -165, and
the sum of whose squares is 155?
Solution
roots?
3. If a, b, and c are different numbers, and if a3 + 3a + 14 = 0, b3 + 3b + 14
= 0, and c3+ 3c+ 14 = 0, what is the value of 1/a+ Mb+ 1/c?(Hint: Since ALG
A fuel tank receives a continuous, steady flow of 2,000 liters per hour.
The tank experiences a steady rate of fuel usage within each of the 6
ARTTH
consecutive 4-hour periods every day. Every day, usage during these
periods is respectively 6,000 13,500, 7,300, 10,000, 8,000, and 3,200 GEN
liters. What is the capacity, in liters, of the smallest tank that could
ensure there would always be at least 200 liters of fuel in the tank?
Solution
Suppose there are x liters in the tank at the start. Then, during each of the 6
consecutive 4-hour periods, the tank would get 4(2,000) = 8,000 liters. Thus, at the
end of the six periods, the tank would contain, respectively, x + 2,000, x - 3,500,
x - 2,800, x - 4,800, x - 4,800, and x liters. In order for the smallest of these, x - 4,800,
116 THE ART OF PROBLEM SOLVING
to be at least 200, x must be at least 5,000. But the tank must be able to hold x + 2,000
liters at the end of the first period. Consequently, the minimum liter capacity of
the tank must be 7,000 liters.
ALG What are all values of x for which (x2 - 5x+5)x2_9x+2° =1?
Solution
The left side can equal 1 in only 3 cases: first, the base = 1; second, the
© exponent = 0 and the base * 0; and third, the base = -1 and the exponent is even.
Let's consider each case separately.
Case I. [base = 1]
In this case, x2 - 5x + 5 = 1, so x2 - 5x + 4 = (x - 4)(x - 1) = 0 and x = 1 or x = 4.
Solution
If such a line existed, then each time it crosses a side of the polygon, one vertex
of the polygon would fall on one side of the line, and one would fall on the other
side of the line. But, for a polygon of 999 sides, it is not possible to have an equal
number of vertices on both sides of the line.
Six circular regions in the plane have the property that none contains
GEOM
the center of any other. Prove that these regions have no point in
common.
Solution
Assume, to the contrary, that the circular regions do have at least one point
in common. Now draw six segments so that each joins this common point to the
center of a different one of the six circles. There are only six segments, and the sum
of the measures of the six angles about the common point is 360°, so there must be
118 THE ART OF PROBLEM SOLVING
at least two segments for which it is true that the angle between them is less than
or equal to 60°. Let the lengths of these two segments be a and b, with a < b. Now
draw a circle, centered at the common point, whose radius-length is b. This circle
must contain the center of the circle associated with segment a. But this center must
also be included within the circle associated with segment b, which contradicts our
original assumption.
Miscellaneous Examples
for Further Consideration
GEN Find the maximum and the minimum number of Friday-the-13th’s that can
occur in a calendar year. Consider both leap and nonleap years.
GEOM A square has a side of length 2. The midpoint of each side is joined to
the two opposite vertices. What is the area of the octagon at the center
of the square? Show that, if “square” is replaced by “parallelogram,” the
area of the octagon is 1/6 the area of the parallelogram.
ALG Ifa + b + c=1,a2 + b2 + c2 = 2, and a3 + b3 + c3 = 3, show that a4 + b4 +
c4 = 25/6.
GEOM The opposite vertices of an xx y rectangle are brought into coincidence,
and the rectangle is then flattened out to form a crease. Show that the
length of the crease is (^ yVx2 + y2 .
GEOM Which fits better: a round peg in a square hole or a square peg in a round
hole?
GEOM Prove that no regular polygon with more sides than a square can be
inscribed in an ellipse.
ALG What are all values of x for which (16X2- 9)3 + (9X2 - 16)3 = (25X2- 25)3?
If x is a real number, what is the minimum value of
ALG
(x+8)(x-8)(x+6) (x-6)?
Motivational Problems 119
Reprise
Is it worth it? If you're "doing quite well, thank you," is it worth investing
the time needed to introduce techniques and technologies that may be unfamiliar
to you but that may improve problem investigation and problem solving? If you
are comfortable in the classroom, why try anything new?
First and foremost, it helps us to teach better: There are always some topics
we've "got to get going" a little better. It's a matter of the old maxim "If you always
do what you've always done, you'll always get what you've always gotten."
Flard work at teaching pays off in the end: When you're alert to new ideas,
your students will be alert to you.
Note
MARIO SALVADORI
jAlII of us who teach mathematics and use it in our daily activities often solve a
problem by the intuition engendered in us by familiarity with our subject. Yet,
once in a while, upon deeper reflection, we begin to suspect that our intuition may
be misleading us. We then check the solution with pencil and paper, a calculator,
or a computer and realize, to our surprise, that we had been wrong. (If this happens
to us professionals, how much more often is it bound to happen to our students?)
This experience is common not just to mathematicians but to all scientists
and technologists. In my professional activity as an engineering designer, I myself
have experienced such doubts and the need to "check the answer." As a forensic
engineer, I have encountered cases where not "checking the answer" had led the
inexperienced or careless practitioner to tragic consequences. A famous case comes
to mind where an apparently "minor suggestion" by the contractor, unchecked by
the designing engineers, led to the collapse of two walkways, or pedestrian
bridges, connecting two sections of a hotel building (in Kansas City, Missouri) at
AUTHOR'S NOTE: All of the engineering-design failures mentioned in this chapter are
taken from Why Buildings Fall Down, by Matthys Levy and Mario Salvadori (New York:
Norton, 1992).
121
122 THE ART OF PROBLEM SOLVING
the second- and fourth-floor levels. The collapse of the walkways within its main
hall killed 114 dancers and maimed almost 200 more people.
To exemplify this kind of occurrence, I will outline the solution of a few
problems in mathematical probability (that purest and most useful chapter of
mathematics), which on the face of it have "obvious" but incorrect solutions,
together with a few additional examples of engineering-design failures due to the
lack of careful checking of design drawings or physical hypotheses.
I suggest that some, if not most, problems in probability are approached
PROB incorrectly by our students because of a lack of clear understanding of the so-called
fundamental theorem of probability theory, a hypothesis that can be easily dem¬
onstrated by its application to a classical problem.
We all know that if we flip an ideally balanced penny, it has the same chance
of showing heads as tails. Since, moreover, it can only show either one or the other,
the probabilities of heads and of tails must both be l/z, because their sum must be
1, the probability of certainty.
It should also be obvious that, since no preceding throw can influence a
subsequent throw (i.e., since the throws, if honest, are independent of each other),
the probability of getting one more head after having thrown 20 heads is still Vz.
Why, then, do most people feel that it is not so and bet tails after getting 20 heads?
It is because they confuse the answers to two different questions. First, we ask what
the probability is of heads on one particular throw; since the penny is unaware of
what has happened before, the chance of heads on the 21st throw is Vz, as it was
for every preceding throw. In the second question, we involve the previous throws
in the count by asking what the probability is of getting 21 heads in a row; of course,
we obtain a totally different answer.
For example, if we ask what the probability is of getting two heads in a row,
we may get the following combinations of heads (H) and tails (T):
The probability of HH is one out of four, or V4—that is, Vz x l/z, the product of the
probability of Vz for heads in each throw. Similarly, taking into account the
preceding throws, the chance of 21 heads in a row have the (extremely small)
probability of (V^)21, or 1 chance in 2,097,152. Thus, the layman's feeling that the
21st throw of heads has a minimal chance when it is preceded by 20 heads is
correct, even if, in fact, the 21st throw itself has the same Vz probability as any
other.
We can now state the fundamental theorem of probability: The probability that
the frequency of a sequence of occurrences will approach the theoretical probability of that
sequence increases with the number of occurrences.
In this statement, the frequency of an occurrence is how often it actually
happens in a sequence of independent occurrences, and what increases is only the
probability that the frequency should approach the theoretical value of the prob¬
ability. The theorem explains that when we throw the penny the 21st time after
Check the Answer, Please! 123
getting 20 heads, there is no guarantee that we will not get one more head, only
guarantees that the probability of getting increasingly many heads will become
less and less. With this preliminary (and much needed) clarification, we can now
take chances knowing what their probabilities are.
In Chapter 2 (p. 95) of this book a problem is ingenuously solved that is a
probability problem but leads to the wrong solution if its statement is not carefully
read (as all statements should be).
Suppose you have 7 pairs of blue socks all alike and 7 pairs of red
socks all alike, scrambled in a drawer. How many socks must be taken GEN
out at the same time (in the dark) to be certain of getting a matching
pair [of any one of the two colors]? (Emphasis added)
All of these combinations satisfy our demand for the red or the blue socks so that the
probability of getting a matching pair is a certainty (i.e., the probability equals 1).
But if we misread the statement and solve the slightly different problem that
asks how many pairs of socks must we take out of the drawer to have the certainty
of getting two socks of the same color? (say blue), we may be unlucky enough to
take out first as many as 14 red socks in a row before getting 2 blue socks—the
worst case, with the tiny probability of j^j14, or 1 in 16,384. Fortunately, the
fundamental theorem reassures us that the probability of being that unlucky is
very small indeed.
Let me mention here another engineering catastrophe due to a lack of care¬
ful reading of their own drawings by the designing engineers of an important
structure.
The 300 ft. x 360 ft. roof of the Hartford (Connecticut) Civic Center arena
consisted of a "space frame" of almost 5,000 interconnected steel bars supported
on four pylons of concrete—a structure famous worldwide because it was entirely
built on the ground and then lifted hydraulically to the top of the pylons 100 feet
above it. The designing engineers had made sure that the highly compressed steel
bars of the roofs interior sections would not "buckle" (collapse in bending), a most
dangerous phenomenon that occurs suddenly without any warning. They did not
realize, however, that the bars along the perimeter of the roof were not prevented
from buckling outward—they did not make sure that these 50 bars could support
the loads on them. Five years after its erection the roof collapsed on a winter night
(at 5 a.m.), dropping to the ground 15,000 tons of steel.
124 THE ART OF PROBLEM SOLVING
There are 30 students in your class. What is the probability B that two
of them have the same birthday? (Assume a year of 365 days.)
Since the probability B that two students have the same birthday added to the
probability p that two students do not have the same birthday is a certainty, they
must add up to 1, and
B = 1 - p.
DEFDBL A - Z
P = 1
FOR i = 2 TO 30
P = p * (365 - i + 1) / 365
NEXT i
B = 1 - p
PRINT ; B
The B probability for 30 students given above was evaluated by this program,
which shows that for a party of 55 people B = 0.986—that is, almost a certainty—
and for a party of 100 people, B = 0.9999997, which to all practical purposes is a
certainty.
When the Comets, the first commercial jet planes built in Great Britain in
1952, first began to fly internationally in 1954, seven of them crashed inexplicably
within 2 years, killing a large number of passengers. It took the Royal Aircraft
Establishment a long time and a lot of demanding research to find out the cause
of these disasters. Eventually, experiments proved that the fuselage of the Comets
had exploded due to minute cracks at the comer of one or more windows. These
cracks were the consequence of two physical phenomena: (a) the stresses at the
comers of the windows, which were much higher than elsewhere in the fuselage
(as they are at the reentrant comers of any opening), and (b) the weakening of a
metal element when the stress in it changes frequently from tension to compression,
from pushing to pulling and back again (you can easily check this phenomenon
by bending back and forth the wire of a paper clip: It will break, usually, after 10
to 20 bendings). The designers had not checked to find that both phenomena did
occur simultaneously during the long, bumpy flights of the Comets, eventually
breaking up the plane after a few thousand miles of flight.
In the following problem, a student shopping at the point of "reasonable¬
ness" has no chance of reaching the correct answer. As teachers, we must insist on
checks beyond just ensuring reasonableness, though its importance is not to be
minimized.
A man has 3 records. The first has songs on both sides (S,S2y the
second instrumental music on both sides (lf2y and the third songs on
one side and instrumental music on the other (S3/3). He comes into his
living room, puts on one of the records, and hears a song. What is the
probability that the other side of the same record also has songs?
126 THE ART OF PROBLEM SOLVING
GEN Since the 3 records have altogether 3 song sides and 3 instrumental sides, all
equally likely, your students may think, at first, that the required probability is
Vl. Some of them may even believe, by analogy with two consecutive heads or tails
in the penny-throws problem, that the probability of two "consecutive" songs is
VixVi = V4. But the two sides of the records are not independent, and the (equally
probable) ways of putting on one of the two records with a song on either one or
two sides are 3:
of which the first 2 satisfy the question. Hence, the probability that the man will
hear a song on the other side of the record now playing is (Neat, but tricky at
first.)
Another example of a design failure due to an unusual physical phenomenon
requiring careful checking is the collapse in 1979 of the 324 x 360 ft. roof of the
magnificent Kemper Arena in Kansas City, Missouri, 6 years after its erection. On
an exceptionally rainy night, water started to accumulate on the roof of the arena,
bending it slightly into a dished shape. As more and more rain fell, accumulating
more and more water on the roof for lack of sufficient spouts, the roof dished more
and more until this self-increasing phenomenon, called "ponding," allowed so
much water to accumulate on the roof that its weight collapsed it. Fortunately, the
arena was not in use that day. (Perhaps because the evaluation of ponding of this
complex roof demanded such unusual calculation, or perhaps because it was
considered wise not to waste too much money on lawyers, the case was settled in
court in 2 days, without involving the engineers' responsibility, despite their
obvious failure in designing the roof's water spouts.)
In order to pass time when sitting in a car next to a silent driver, I often play
the "car-plate game." The game gives me a feeling for the probability of the
particular plate of the car in front of ours, measured against all the possible car
plates, a question I feel compelled to answer because, for some mysterious reason,
those plates look almost always "peculiar." (To enjoy the game, your students
don't have to be mathematical geniuses like the Indian mathematician Ramanujan,
who was once told by one of his English mentors, "The license plate of the taxi you
took to come here exhibits a 'most indifferent' number." "Not so," exclaimed
Ramanujan, after just a glance at the plates, "It is the only integer that can be
expressed in two different ways as the sum of the squares of two integers.")
Ask your students.
Ni N2 N3 - L\ L2 L3,
©
and you should assume that both numbers and letters are assigned to plates at
random and, hence, all have the same probability of being chosen. Assume also COMB
that plates starting with zero or having all zeros are acceptable (although in the
United States they are not). Do not consider vanity plates (e.g., ILOVEYOU2).
Assume that you have 3 separate sets of the 10 numbers 0 to 9, and choose
the first number for the plate out of the numbers of the first set, the second out of
the second set, and the third out of the third set. You then have
ways of combining the numbers of the 3 sets to get the 3 numbers of your plate.
Hence, all the possible combinations of 3 numbers in a plate are 1,000. (In this case,
you can simply count 1,000 numbers from 000 to 999, yet the general counting
technique is important as follows.) Similarly, if you have 3 sets of the 26 letters of
the English alphabet, you can combine them into groups of three in
263 = 17,576
ways. Suggest to your students they check that, therefore, one can make 17,576,000
different car plates of the type we are considering.
To build up a plate with 3 different numbers from the same set, you can take
any one number from the first set in 10 different ways, but then you are left with
only 9 numbers in the second set unequal to the first number and only 8 numbers
in the third set unequal to the first two numbers you have chosen. Thus, you have
10 x 9 x 8 = 720
possible combinations of three different numbers, and the probability of one such
plate among the 1,000 possible plates is 0.72.
Since the sum of the first and second probabilities must add up to 1, the
probability of a plate not having 3 different numbers is only
1 - 0.72 = 0.28.
Of course, you may check this result by computing first the probability of a plate
having 2 equal numbers and then of one having 3 equal numbers. The first can be
obtained in 3(10 x 9 x 1) = 270 ways, because the unequal number can be in 3 posi¬
tions, with a probability of 270/1,000 = 0.27; and the second in 10 x 1 x 1 = 10 ways,
with a probability of 10/1,000 = 0.01. Their sum is 0.28, as more easily evaluated
above.
128 THE ART OF PROBLEM SOLVING
COMB How many different plates can you get with 3 different numbers and 3
different letters?
How many different plates can you get that have 3 equal numbers and
3 equal letters? What is the probability of one of them?
Ni N2 Ni — Li L2 L\?
Once you choose Ni, something you can do in 10 different ways, there are 9 ways
for N2 to be different from Ni but only 1 to be equal to Ni. Hence, there are
10 x 9 x 1 = 90 ways of obtaining a "symmetrical" number with 3 digits. Similarly
there are 26 x 25 x 1 = 650 ways of getting "symmetrical" 3-letter groups. Thus, the
number of "symmetrical" plates is 90 x 650 = 58,500, and each such plate has a
probability of 58,500/17,576,000 = 0.00333, or 1 in 300.
You can use probability theory with your students for a better purpose than
we have so far, so ask them this:
PROB
Is it a good investment to play the New York State Lotto in the hope of
hitting the big jackpot and making millions of dollars?
In this betting game, 50 balls, numbered 1 to 50, are put into an urn. Then 6
balls are taken out of it, one at a time. To hit the big jackpot, the 6 extracted numbers
must check the 6 numbers you have chosen to bet on.
Check the Ansiver, Please! 129
The probability of the first ball checking one of your 6 numbers is 6/50; of
the second checking one of your remaining 5 numbers, 5/49; of the third, 4/48;
and so on. Hence, their combined probability is
or 1 in 15,893,198! (Even if you lived to be 100 and played the Lotto 52 weeks a year
from birth, your chance would be only 1 in 3,056. I would tell my students that
they should find a better investment for their money than playing Lotto in New
York State.
I hope that with this kind of problems you will be able to convince your
students that probability can be both fun and useful and that it always pays to
know what our chances in life may be, even if we will never know for sure whether
they will match our probability forecasts.
I would like to end this chapter with the well-known story of the tragic
explosion of the space shuttle Challenger, which killed seven astronauts in 1986.
The explosion occurred because the NASA designers had not checked the loss of
strength due to lowering temperatures in the O-rings of the shuttle's fuel tanks.
Nobel physicist Richard Feynman proved the influence of low temperatures on
the strength of the rings with a very simple experiment: He dropped an O-ring in
a glass of iced water, making it so fragile that he could break it with his own hands.
No wonder the rings broke up as soon as the shuttle reached the higher layers of
the atmosphere!
Mentioning to your students these and similar catastrophes due to forgotten
and incomplete checks in engineering design may impress them with the basic
importance of checking the answers to their own problems, thus helping them to
acquire a "checking habit" to their own advantage and, perhaps later on, to that
of all of us.
The Logic of Error
ETHAN AKIN
Sometimes you hear something surprising, then you suddenly notice you're not
surprised. In the early 1970s a counselor at the City College of New York organized
a discussion group for students who were afraid of mathematics. To several of us
in the Math Department he described his students' view of our subject: They saw
mathematics as a disorganized mess of unreasonable rules. Each rule appeared
unrelated to its predecessors, and all were rendered unreliable by mysterious
exceptions. In short, they thought of mathematics patterns the way I think of
irregular verbs in a foreign language. Shocked at first, our impulse was to cry out:
"No, math is not like that. Math is beautiful, neat, and logical." However, we
became uneasily aware that their descriptions were merely making explicit atti¬
tudes that were already familiar to us. This view of mathematics as muddle was a
mistake with which we did daily combat.
It took much longer before I noticed that I myself was making an analogous
mistake about the students' patterns of thinking. I began to realize that their errors
were not entirely chaotic. Underneath the designs of correct logic, their errors were
organized in layers of attractive, wrong rules and natural, if incorrect, patterns of
thinking. The perceived chaos resulted, in part, from the discordance between the
patterns—the clash between the false rules and the true.
131
132 THE ART OF PROBLEM SOLVING
Looking for patterns in the students' thinking made me much more self-
conscious about my own. It became apparent that the high-level thinking used in
research mathematics, and which we are trying to teach as "problem solving," uses
other, almost unnoticed, skills. Especially the symbolism and notation in algebra
is packed with subtlety to which familiarity blinds us. For example, I cannot
remember when I learned the routine for long division of numbers. I do remember
that some teacher told us that this was a very advanced topic in the Middle Ages.
Only when I began teaching did I try to imagine a long-division problem in Roman
numerals and appreciate the power of the Arabic number system and the associ¬
ated arithmetic routines.
I don't really understand much about this buried psychology of logic, al¬
though I believe it is well worth excavating. What I present below is a collection
of fragments, samples of mistaken rules, or examples where symbolism exerts a
force supplementary to the mathematics. With these as samples, I recommend that
you look out for and collect similar mental tangles, for these can often be the simple
blockades to successful problem solving. These often overlooked deterrents need
to be addressed before the student can embark on the path to successful problem
solving.
Everything Is Linear
I commence with what I think of as the most popular wrong rule, which we
have all seen in operation and which appears in the following examples:
ALG
V(a + b) = Vo” + ViT
TRIG sin (a + b) = sin a + sin b
Of course, students don't usually quote such a rule; they just apply it as in
9x2 + 1=4,
ALG
3x+ 1 = 2
(* + !) + (2x - 5) = 9.
The Logic of Error 133
My students tend to recognize and accept this rule as we go over the problem:
You square both sides, right? So we get the square of this plus the square of that
equals the square of 3, right?" Right.
This pattern is so seductive that you have to confront it. I tend to talk almost
as much about this false law as about the true rules for manipulating radicals. A
more advanced, but similar, example comes up in matrix algebra.
The most important rule in dealing with matrices is not the associative or
even the distributive law but the commutative law for multiplication: AB = BA. It
is so important because it is not true and because students use it anyway.
Of course, it is not always clear that the benefit of describing a wrong rule
outweighs the risk of implanting the rule without the accompanying "Thou shalt
not." However, in my opinion, the wrong rule is sometimes a good introduction
to the right one, as demonstrated in the following section.
I recall a book in which the author sneers at a student who adds fractions
according to the rule
a+c
~b + d'
different words are used for cardinal and ordinal numbers: "three and four," as
distinct from "third and fourth." This distinction fits together with the familiar
pie-cutting illustration of fractions. The numerator of the fraction is the count, the
number of pieces of pie. The denominator is what you are counting, the size of
each piece. When you think of the denominator as the unit in which you are
making a measurement then the strange computation
2 5 _ 7
3 + 3 “ 3
134 THE ART OF PROBLEM SOLVING
Order-of-Operations Nonrules
3 + 2x5.
You can arrive at the correct 13 and avoid the error of 25 by consciously applying
the rule that multiplication takes precedence over addition. But the instinct that
leads to the 25 is really the right one. The way the problem is printed looks strange
to anyone who is accustomed to algebra or even arithmetic, because once you have
digested the order-of-operations rules, you no longer think about them. Instead,
you see them. For instance, you see the expression
3x2 + 7
3x2 +7.
In other words, you see a (+) or (x) sign as forming a linkage between symbols like
a chemical bond, and the (x) linkage is much tighter than the (+) link.
The analogy I use is playing a musical instrument. When the playing is
correct and smooth, you do not consciously direct your muscles. The knowledge
The Logic of Error 135
is "in the fingers." Only in correcting an error do you slow up and take conscious
control, practicing and repeating to retrain the fingers to get the passage right. As
the error is corrected, you gradually relax the conscious control until finally the
corrected portion is played fluidly. With order of operations the situation is
similar. You don't practice to use the rules; you practice to train the eye so that you
no longer need to use them.
An even better example, where the symbolism is intended to invite an
unconscious blurring of meaning, occurs in the following section.
For many years, I taught that the symbol (-) had two different meanings:
ARITH
subtraction and, when attached to a number, the negative sign. I don't remember
how long it was before I noticed the third meaning of the symbol. I think I had to
discover it on my calculator, where it is designated "chs" for "change sign." The
confusion between this unary operation and the sign of a negative number ac¬
counts for the common mistake of thinking that -x is negative. (This causes
incredible confusion when you attempt to define the absolute value function
formally by using a branching formula, or "-x when x is negative.") Especially
because a calculator must use different symbols to distinguish them, the questions
arise: Why use the same symbol to refer to these three different things? Is this not
proof at last that mathematicians are engaged in a sadistic conspiracy to make their
subject harder to learn? Actually no. The intention is to allow you to switch
meanings without thinking about it.
While it is hidden by the symbolism, subtraction of signed numbers works
like division of fractions. Just as you "invert and multiply" to divide, you "change
sign and add" to subtract. That is, to perform
(-3) - (-5),
what you really do is change the sign of minus 5 and replace the subtraction by
addition—that is, replace the original problem by this equivalent:
But the whole point is to avoid for subtraction of signed numbers what is unavoid¬
able for fraction division. You should think of the only operation as addition. Each
minus is to be regarded as multiplied onto the term that follows it. Thus, the
problem above should look (unconsciously) as though it were written
(-3) + - (-5).
Again the question arises whether it is causes more confusion to point out
the three meanings than it does to shift from one use to another without comment.
For an algebra class, I favor the former if for no other reason than to avoid the
“-x is always negative." mistake mentioned above.
There are also cases where the teacher operates unconsciously in ways which
cause some students to lose the thread of the argument.
ALG 2y = 3x + 4,
(3x + 4)
or
+ 2
or
y 2
We choose which one to use based upon our subsequent intentions. Thus, if we
are about to substitute values of x, we are likely to use the first or second form. On
the other hand, if we are computing the slope we use the third form, which displays
the coefficient of x. These equivalences are not obvious. For example, second-term
calculus students often fail to recognize that
r dx
J x+1
and
1
I x+1
dx
are the same. I have found it hard to teach this facility for choosing the appropriate
form. The best I can do is to point out the various alternatives as they happen and
indicate the advantages of each form. However, this leads to the classic question
posed in the next section.
The Lbgic of Error 137
Moving left to right is "reducing fractions" by canceling a common factor from the
numerator and denominator. Moving right to left is "building fractions," the
central step in fraction addition.
It is precisely the fact that these two processes seem to many students
unrelated that accounts for a characteristic error in fraction addition that would
be funny if only it did not happen so often. The student recognizes that a common
denominator is necessary, computes the least common denominator correctly, and
furthermore converts each summand to the common denominator. Then, on the
brink of success, having completed all of the hard work, the student suddenly
notices he or she can reduce the fractions. This the student proceeds to do without
realizing that he or she is exactly undoing the work of the previous conversion
step. Having lost the common denominator, the student now wanders further into
the darkness, adding the numerators and adding or multiplying the denominators.
To See or to Do?
Monomial Versus Polynomial Factoring
Factoring is factoring, after all. But monomial and polynomial factoring not
only feel completely different, but the skills developed for the first work against
success in the second. For a monomial like
30x3y2,
the student should see the number 30 as containing its prime factors. Furthermore,
r3 is three xs multiplied together, with y2 similar. Students learn this skill easily,
and it is just what is needed for problems like reducing monomial fractions.
However, they are then confronted with
6^-y
and asked, "What are the factors?" They reply, quite reasonably, "2,3, x, x, and y."
You can point out the difference: In a monomial the only operation is
multiplication, while a polynomial contains a mixture of multiplications on one
hand with additions and subtractions on the other. It is this mixed state that causes
the difficulties in the polynomial case. So the mixed state should be used as a
danger signal. While for monomials the factors are just what they appear to be,
polynomials are to be factored only by one of the short list of procedures—for
example, dividing out the common monomial, the difference of two squares
method, or trinomial factoring—leaving other, subtler polynomial factoring skills
to be developed at a later time.
However you justify the difference, I fear that the students perceive this as
one of the irregular verbs I mentioned in the introduction.
The Logic of Error
139
x7x5 = x7 + 5 = X12
just says that with all the xs multiplied, you only have to keep track of the count. So
if a bag with 7 xs is poured together with a bag containing 5 xs the result is 12 xs
all multiplied together. Similarly,
says that 3 bags, each containing 5 xs, contain all together 15 xs. Trouble arises
when you want to define negative or fractional exponents. How should you
multiply together minus 3 xs to get x~3? On the other hand, multiplying half an x
seems perfectly simple and so x1/2 - x/2.
It is worth pausing here to talk a bit abstractly about the process of abstrac¬
tion. The problems the students have, illustrated by the above questions, arise from
trying to extend the original definition of exponent by extending the original
physical sense of what exponents are. However, we are at a fork in the road, and
we should pause before we decide which way to go. From our original physical
picture of exponents we observed some useful rules, the Laws of Exponents. In
fact, they are so useful that when we extend our notion of exponents to negative
and rational numbers, we want to make sure that we retain the rules even at the
cost of our original physical intuition. That is where the peculiar new definitions
come from. Think of x"3 as an unknown quantity. If it satisfies the Law of Expo¬
nents, then we must have
x~3x4 = x“3 + 4 = X,
Similarly,
(x*)2 = x’^*2 = x
direction, in contrast with a scalar, which is just a magnitude. Vectors are pictured
as little arrows. The physical motivations for the idea are force and velocity. We
discover notions of addition and scalar multiplication for vectors that relate
naturally to phenomena concerning forces and velocities. We then observe that
these operations satisfy a list of useful rules analogous to those of ordinary algebra.
In linear algebra, these ideas are abstracted to define a "vector space." Again, the
physical motivation is thrown away. What is retained is the list of algebraic rules
as the axioms for a vector space.
5x = 3x + 6
easy to solve. However, even calculus students are often unable to correctly solve
5x = 3x
I am still not sure why. Sometimes I think it is symmetry that prevents an "obvi¬
ous" choice of side on which to isolate x. It is, for example, true that the equation
5x + 8 - 3x + 6
is harder than the first equation. But it is not nearly as hard as the second. Another
possibility: In the first you have the 6 from which to take away the 3x. You have
something against which to brace your foot, as it were. Whereas in the second there
is nothing there.
FRED PAUL
141
142 THE ART OF PROBLEM SOLVING
A fly and a jogger are 12 km apart. The jogger runs toward the fly at a
rate of 4 km/hr. The fly travels toward the jogger at the rate of 6 km/hr.
When the fly meets the jogger, it turns around and flies back to its
starting point, then heads again toward the jogger, then flies back
again. The process continues. What is the total distance the fly travels?
With certain assumptions, this is a very simple problem but one that stumps
many. Often, when there is no intimidation and there is a willingness to make
guesses, a variety of answers is offered. Eventually (with proper guidance, if
necessary), but not always quickly, the ambiguous aspect will be discovered: There
is no mention in the problem, as stated, of when the process stops. Most will
assume that it stops when the jogger arrives at the fly's starting point, but this must
be stated in the problem in order to avoid the ambiguity. Even with this done, the
problem is not always solved quickly and can be the source of a good exchange
among students. It is not just in geometric problems that students must be
reminded to consider all the information given. Because of the constant activity of
the fly, going back and forth, and because it is the fly's distance traveled that is
being sought, students usually concentrate all their attention on it. Actually, by
ignoring the fly and working with the jogger, the time of travel is easily arrived at.
By using nothing more than the formula rate x time = distance, it can be determined
that the fly traveled 18 km.
Many mathematicians will object to the purposeful use of ambiguity in a
problem, but, while best avoided in a testing situation, it is a useful tactic to
encourage a diversity of questions and to make students aware that problems do
not as a rule appear in neatly wrapped and predictable packages. It might be added
that many of the same people who would object to the ambiguities would consider
mathematics in many ways to be a guessing game.
Another important aspect of a problem like the jogger-fly example is that it
demonstrates that difficulty is not a proper measure for a useful and valid problem.
Whether easy or difficult, routine or nonroutine, the good problems are those that
will nurture questions, diverse thoughts, differences of opinions, guesses, and,
above all, a willingness to pursue various alternatives and keep going if unsuc-
144 THE ART OF PROBLEM SOLVING
cessful. This is the true trial-and-success pattern, the joy of the problem-solving
process. A solution is merely one other part of this process and could and should
be a catalyst for further questions, discussion, and analysis. In this way, mathe¬
matics becomes the vital and interesting subject that truly helps students learn to
think.
We have in the following example a routine problem with some nonroutine
features:
Three circles are mutually tangent, and each is tangent to the same
GEOM
line. Two of the circles are congruent and the third, incongruent circle
has a radius of 3. What is the radius of the congruent circles?
Trial and Success 145
Jim asks a classmate how old she is. The classmate answers, “I was
14 the day before yesterday, but I’ll be 17 next year.” What are the GEN
Student reactions to questions of this type can range from puzzled concern
to outright denial of the possibility. However, questions and guesses will point out
the importance of the students' ability to communicate and work together and
demonstrate the idea that this is what problem solving is all about. There is nothing
ambiguous about this problem, with a December 31 birthday being talked about
on January 1. The problem solver may make many guesses or "errors," however,
before arriving at the correct response or "success."
Good problems can also be used to expand and illustrate major mathematical
concepts:
The ensuing discussion will most likely involve fractions (part of the crate,
even though the problem avoids this) and possibly the need for more information.
Ultimately, however, there should be agreement that, no matter what, half of what
is in the truck must remain there after each drop-off point. There is no "last" crate
to drop off. The "trial" in this situation will be the various answers offered, and
the "success" will be the realization that there is no answer. Problems like this are
valuable since they lead to a different kind of answer and allow the instructor and
the class to delve into other topics such as limits and infinity.
The questioning and discussion resulting from problems like these often
should be considered in a full-class situation. Many teachers will claim that there
isn't enough time for much of this kind of activity because of curricular and testing
requirements—but actually there is not enough time not to do it. The ability to
think through and understand the problem-solving process is arguably the pri¬
mary goal of mathematics instruction. It is necessary for students to hear their
questions, answers, and comments clarified, interpreted, and challenged by other
students. This exchange of ideas will become valuable for later tasks. It is equally
important for the teacher to monitor the class discussion. Many who are not
necessarily "good" mathematics students in the traditional sense—that is, those
who don't score well on tests and may have little interest in the mechanical or rote
aspects of the subject—may show original thinking and insight in the nonroutine
problem-solving situation when given the opportunity in an open-ended discus¬
sion. In this way, a teacher can discover abilities not otherwise apparent, and the
student can develop a more positive attitude toward mathematics. These students
may then view the subject as more than a set of rigid rules where the idea of any
kind of error denotes failure. This prospect further emphasizes trial and success.
A full-class approach to problem solving also allows for examples containing
topics and terms not familiar to everyone. For example, there are those who will
frown on the use of a problem involving sports, since there will be some in the
class who will not be familiar with the given sport and its terminology. In adult
life, we often face problems that include unfamiliar aspects; an important part of
the problem-solving process is the clarification of language and definition of terms.
The full-class situation is ideal for pursuing these matters and stressing the
importance of a clear understanding of what is given and what is sought. An
example:
preassigned problems, the discussion of which will be about work and thinking
that students have already done. It is also important that problems be tackled by
full class participation from the beginning. In this way, all students get the benefit
of being part of the entire process, contributing to and hearing different ways of
interpreting a problem and determining how to reach a solution. Discussions of
this nature become valuable tools for the instructor's assessment not only of
student performance but also of the mathematics program itself. Even for the
teachers and for the instructional validity of the program, it is a matter of trial and
success.
7
STEPHEN KRULIK
JESSE A. RUDNICK
Rarely in problem solving does the problem solver resort to a single strategy. In
most cases, two or more strategies are combined to resolve the situation. In fact,
reduction and expansion almost always depends on discovering a pattern. Ironi¬
cally, many years ago, the mathematician W. W. Sawyer described mathematics
as a search for patterns.
In order to fully appreciate the strategy of reduction and expansion, we must
first differentiate between simple reduction and reduction and expansion. In many
textbooks, the strategy of reduction is referred to with the direction solve a
simpler problem." When using this strategy, the numbers in the problem are
reduced to simpler quantities, the simpler problem is solved, and the solution
procedure is then applied to the original version of the problem. On the other hand,
reduction and expansion is an appropriate strategy to use when the problem
contains arbitrarily large numbers and the solution is virtually independent of
these numbers. When using this strategy, one reduces the arbitrary number to the
simplest possible case (usually 0 or 1), notes the results in a table or chart, and then
permits the number to expand. Again, the results are noted in the table, until an
underlying pattern emerges for which a generalization can be made.
149
150 THE ART OF PROBLEM SOLVING
Teachers should not treat these tables and charts in a casual manner. Children
need experience in handling data. The reduction-and-expansion strategy almost
always involves the use of a table to keep track of the changes that occur as the
variable expands. These tables should be designed and completed by the chil¬
dren. A careful analysis of this table permits the student to observe the develop¬
ing pattern. This analysis will serve students well in all of their mathematical
experiences.
To illustrate this strategy, we will present a series of problems. Note that
problems are merely vehicles for the teaching of reasoning and problem solving.
They need not be from the real world nor applicable to any particular occupation.
It has been our experience that children can easily be motivated by science fiction,
fantasy, or recreational topics.
Example 1
Laura is training her pet white rabbit, Ghost, to climb a flight of 10 steps.
Ghost can hop up 1 or 2 steps each time he hops. He never hops down,
only up. How many different ways can Ghost hop up the flight of 10
steps?
Table 7.1
1 1 1
2 2 1-1, 2
4 5 1- l-l-l 2-1-1
2- 2 1-2-1
1-1-2
5 8 1- l-l-l-l 2-2-1
2— 1—1—1 2-1-2
1—2—1—1 1-2-2
1—1—2—1
1—1—1—2
6 13
7 21
8 34
9 55
10 89
the conjecture. The fact that the rabbit takes only 1- or 2-step hops implies that
Ghost can only reach the 5th step from either the 3rd or 4th step. Thus the total of
these possibilities (step 3 and step 4) is the entry in the box for the 5th step. This
exhibits the notion seen in the Fibonacci sequence, where each term after the first
two is the sum of the previous two terms.
The main purpose for this problem was to involve the students in analysis,
reasoning, and the reduction-and-expansion strategy, but more should be done.
The problem should be used to stimulate further discussion. Creative teachers
would ask "what if" questions: What if the rabbit could hop up 1, 2, or 3 steps at
a time? How would this affect the answer? Would the number of ways increase or
decrease? Using the same reduction-and-expansion technique reveals interesting
patterns, including a Fibonacci-type sequence with three consecutive terms being
summed.
A top priority for all mathematics teachers at every level is to reveal the
power of mathematics to their students—that is, to use mathematics to completely
describe a situation or a phenomenon and to predict with certainty the outcome.
This goal can often be achieved by generalizing after reduction and expansion have
revealed the fundamental pattern.
152 THE ART OF PROBLEM SOLVING
Table 7.2
0-0 1 0-0
• •
8-8 81
Example 2
At the end of the 7th inning of last night’s baseball game, the score
was 8-8. How many scores were possible at the end of the 6th inning?
Once again the arbitrary numbers 8-8 complicate the problem. Thus we
reduce the score to 0-0, expand it to 1-1, 2-2, 3-3, and so on, keeping track of the
results as we go. Again we make a table (see Table 7.2).
We can generalize the set of perfect squares revealed in the table to (n + 1)
(n + 1), or (n + l)2, where the tie score was n-n.
This reveals the unique power of mathematics to describe a situation. Re¬
gardless of the size of the score, a tie produces the expression (n + l)2.
Again, the creative teacher can ask. What if the score was a-b, not a tie? This
leads to the expression (a + 1 )(b + 1). Thus the tie score n-n is now a special case of
the general expression, where a = b = n.
Some students may recognize that this is an illustration of the counting
principle. Since the first team can have any of nine scores (0-8) and the second team
can also have any of nine scores (0-8), there will be 9 x 9 or 81 possible scores.
Reduce, Expand, and Look for a Pattern 153
Table 7.3
1 0 •
2 1
.
3 3
(
4 6
5 10
•
#
•
26 325
Example 3
There are 26 teams participating in the annual football draft. Each team
office has a direct line to each of the other team offices. How many
telephone lines are there?
We have placed the column with the number of telephone lines directly
adjacent to the column with the number of offices. Although this is different from
the order suggested by the problem, it has been our experience that close proximity
of the numbers makes discovery of the relationship easier. Notice that a problem
that begins with telephone lines results in a geometry problem involving the sides
and diagonals of a polygon of n sides.
Some students may require carrying the table further, in order to observe the
pattern (i.e., the difference between the terms in the "Number of Lines" column is
the counting numbers 1, 2, 3,. . . 25 ). Many students will continue this for all 26
terms. However, some more astute students may observe that the number of
telephone lines required for n offices is given by the expression
n{n - 1)
2 '
This is not a simple observation to make. The teacher may have to lead the student
to this discovery.
For senior high school students, this is an excellent opportunity to consider
the method of finite differences. This can be demonstrated as shown in Table 7.4.
Reduce, Expand, and Look for a Pattern 155
Table 7.5
Size of Squares
lxl 1 1
2x2 4 1 5
3x3 9 4 1 14
Size of Board
4x4 16 9 4 1 30
•
•
•
8x8 64 49 36 25 16 9 4 1 204
===^=
Since the constant appears in the A2 column, we know that the formula will
be of second degree.
Example 4
The first reaction to this problem is that there are 64 squares. However, these
are only the 1 x 1 squares, and there are, indeed, 64 of them. But what about the
board itself? This is an 8 x 8 square. Further analysis reveals that there are 2 x 2
squares, 3x3 squares, and so on. Reduction and expansion is the obvious strategy.
There are 204 squares on an 8 x 8 checkerboard. Once again, the more ad¬
vanced students should continue this problem and develop a formula for the sum
of the squares. Table 7.6 illustrates the method of finite differences, revealing that
the constant occurs in A3, indicating a cubic relationship. The resulting formula is
156 THE ART OF PROBLEM SOLVING
Example 5
The new school building has exactly 1,000 lockers and 1,000 students.
On the first day of school, the students meet in the schoolyard and
agree on the following plan. Student 1 will run through the building and
open all of the lockers. Student 2 will then enter the building and close
all the even-numbered lockers (2, 4, 6, . . . 1,000). Student 3 will then
enter the building and reverse every third locker beginning with number
3 (3, 6, 9, . . . 999)—that is, if the locker is open, he will close it, and if
it is closed he will open it. Student 4 will then reverse every fourth locker
(4, 8, 12, 16 . . . 1,000). The students will continue in this way, until
student 1,000 has reversed locker number 1,000. Which lockers now
remain open?
Table 7.7
Locker # 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Student 1 o O O o o o o o o o o o o o o o o o o o
2 c o c o c o c o c o c o c o c o c o c
3 c c o o o c c c o o o c c c o o o c
4 o o o o o c c o c o c c o o o o o
5 c o o o c o o c o c o o o o o c
6 c o o c o o o o c o o o c o c
7 c o c o o o o o o o o c o c
8 c c o o o o o o c o c o c
9 o o o o o o o c o o o c
10 c o o o o o c o o o o
11 c o o o o c o o o o
12 c o o o c o o o o
13 c o o c o o o o
14 c o c o o o o
15 c c o o o o
16 o o o o o
17 c o o o
18 c o o
19
c o
20
c
this is still impractical (what if there had been 10,000 lockers?). Reduce, expand,
and look for a pattern!
Observation shows that the lockers whose numbers are the perfect squares
remain open. But an interesting question remains. Why the perfect squares? Here
is an opportunity to look at some number theory concepts. This problem is based
on numbers and their factors. A locker will only be touched by a student whose
number is a factor of a locker's number. But factors always occur in pairs (open-
close):
12 contains 1, 2, 3, 4, 6,12
10 contains 1, 2, 5,10
so all of the lockers should be closed. But why did the perfect squares remain open?
The factor 4 has itself as a "mate." Thus, only the perfect squares have an odd
number of factors, a fundamental theorem in number theory.
158 THE ART OF PROBLEM SOLVING
Note
The solution to this problem gives rise to the "Fibonacci sequence," 1,1, 2, 3, 5,
8, 13, 21,... where each term after the first two is the sum of the two terms
immediately preceding it.
8
ALFRED S. POSAMENTIER
WEI LEE
Logical counting has fascinated people through the ages. When the counting
is straightforward, we use schemes such as those described in the Introduction.
However, sometimes problems ask for an ordering—that is, for the data to be
arranged according to certain rules. The inspection of this ordering can best be
captured by the Pigeonhole Principle.
Before considering the Pigeonhole Principle in a formal sense, let's consider
a few examples that will demonstrate how the Pigeonhole Principle manifests
itself.
There are 5 pigeonholes in the tree. There are 6 pigeons flying into these
5 holes. Therefore, there is at least one hole containing 2 pigeons.
159
160 THE ART OF PROBLEM SOLVING
Principle 1. Put k + 1 objects into k holes; at least one hole will contain 2 or more
objects.
Principle 2. Put m > 1 objects into n holes. If n divides m evenly—that is, itm = nq
(q e N)—then there is at least one hole containing at least q objects.
It follows that if each of the n holes has less than q objects, then the number
of objects all the holes contain will be less than nq.
Principle 3. Put m > 1 objects into n holes. If n cannot divide m evenly—that is, if
m = nq + r(0<r< n)—then there is at least one hole containing at least q + 1 objects.
It follows that if every hole has fewer than q objects, then the number of
objects all the holes contain is less than nq, which is contradicted by the assumption
that the total number of objects is nq + r > nq.
,
Principle 1 is a special case of principle 3. Actually, if m = n + 1 then princi¬
ple 3 becomes principle 1.
The Pigeonhole Principle, comprising the above subprinciples, provides a
rather effective problem-solving tool for dealing with a specific type of (some¬
times) difficult problem.
The Pigeonhole Principle 161
Examples
On selects 5 cards from a set of cards colored red or blue. Prove that,
whichever way you select these colored cards, there are at least 3
cards of the same color.
Proof. Consider the red and blue colors as two different holes. Assume that the red
cards are put into the "red hole." Similarly, the blue cards are put into the "blue
hole." According to the Pigeonhole Principle, there are at least 3 cards in one hole.
This proves that there are at least 3 cards with same color.
This example is simple. But considering colors as holes is usually not easy
for students to comprehend at the start. Additional examples are provided to help
secure an understanding of this very important and useful principle.
There are n people in a meeting. Every one of them knows at least one
of the other n - 1 people. Thus, among these n people there are at
least 2 people who know the same number of people.
Proof. Since everyone knows at least one of the other n -1 people, then the number /^\
of people every person knows is 1,2,3,..., or n-1. Consider "the number of people V_y
every person knows" as holes. In other words, we make n-1 holes:
Now we can put these n people, according to the number of people they
know, into these n -1 holes. According to the Pigeonhole Principle, there is at least
one hole that contains 2 people. This proves that within these n people there are
at least 2 people who know the same number of people.
The Pigeonhole Principle can also be used to solve problems related to integer
points (or lattice points) on a coordinate system. An integer point is a point with GEOM
integers as coordinates. If Pfxv yx) and P2(x2, y2) are two points in the X-Y plane,
then the coordinates of the midpoint M of the line connecting these 2 points is
xi + x2 , _ vi+ .v2
2 2
162 THE ART OF PROBLEM SOLVING
Clearly, if Pi(xi, yi) and Pi(x2, yi) are integer points, and both x\ and X2, and yi and
1/2 have the same parity, then the midpoint is also an integer point.
Similarly, if P1(x1/ xv za) and P2(x2, y2, z2) are integer points in the three-
dimensional plane, and x1 and x2, yx and y2, and z1 and z2 have the same parity,
then the midpoint of the line connecting these points is also an integer point.
In X-Y plane, pick up any 5 integer points. Prove that there must exist
2 of these integer points where the midpoint of the line connecting
these 2 points is an integer point.
Proof. (Method 1). Assume the 5 integer points are Pi(xi, yi), P2 (X2, y2), P3(*3, yd),
?4(x4, y4), and Ps(x5, ys). First consider their x-coordinates: xi, X2, X3, X4, and X5.
According to the Pigeonhole Principle, at least 3 of these 5 numbers have the same
parity. Assume these three numbers are xi, X2, and X3.
Consider the y-coordinates of the 3 integer points Pi(xv yx), P2(x2, y2), and
Pfx3> yf)'- yv yi> and y3. According to the Pigeonhole Principle, 2 of these 3 numbers
have the same parity. Assume these numbers are yx and y2. Now we have 2 integer
points, P1(x1, y2) and P2(x2, y2), where both x1 and x2 and y1 and y2 have the same
parity. Therefore, both
Xl + X2 .V1+.V2
and
1~ 2
are integers. This proves that the midpoint of the line connecting Pi and P2 is an
integer point.
Proof. (Method 2). Use the following method to construct the holes. Consider the
ordered integer pair (x, y). A total of four odd-even cases are possible for the
ordered integer pair: (odd, odd), (odd, even), (even, even), and (even, odd). We
construct 4 holes and put all 5 integer points into these 4 holes. Therefore,
according to the Pigeonhole Principle, there is at least one hole that contains 2
integer points, say Pi(xi, yi) and P2(x2, y2). Therefore, both xi and x2 and yi and y2
have the same parity. This proves that the midpoint of the line connecting Pi and
P2 is an integer point.
We can use the first proof method of the previous problem to solve this
example, but this requires us to use the Pigeonhole Principle three times for the
three pairs of numbers (x-coordinate, y-coordinate, and z-coordinate). This is a bit
tedious, so we shall use the second proof method of the previous example.
2 x 2 x 2 = 23
different kinds of even-odd combinations of the ordered triple (x, y, z). Therefore,
there are 8 holes, and we have 9 integer points. Again, according to the Pigeonhole
Principle, there is at least one hole that contains 2 integer points, say Pi(xi, yi, zi)
and ?2(X2, yi, 2:2). Then both x-coordinates, xi and X2, have the same parity.
Similarly, it is true for y-coordinates yi and y2 and z-coordinates zi and Z2. This
proves that the midpoint of the line connecting Pi and Pi is an integer point.
Prove: For a set of 27 different odd numbers, each of which is less than NUMB
Proof. There are 50 odd numbers less than 100: 1, 3, 5,..., 49, 51, 53,. . ., 97, 99.
Within these 50 numbers, there are only the following 24 pairs of numbers whose
sum is 102: {3, 99}, {5, 97},..., {49, 53}; that is, {2k - 1,103 - 2k) (k = 2, 3,..., 25).
Consider the following number sets as 26 holes: Al = { 1}, A2 = {3, 99},
A3 = {5, 97},..., Ak = {2k - 1,103 - 2k),..., A25 = {49, 53}, A26 = {51}. (N.B. A1 and
A26 must be included to provide for the possibility of having the numbers 1 and
51 selected, although they cannot fit the program successfully.) And now we put
these 27 odd numbers (less than 100) into these 26 holes. Then, according to the
Pigeonhole Principle, there is at least one hole that contains 2 numbers. Since the
sum of 2 numbers in every hole is 102, we have what we want to prove.
Randomly pick 50 numbers from the natural number set {1,2,3,... ,100}
and prove that there must exist 2 numbers (of 50) such that one is the
multiple of the other.
Proof. From the above several examples we should consider the following. In
order to prove that there are 2 of n objects that satisfy the requirement, we need to
construct n- 1 holes. For this problem we need to construct 50 holes, and every 2
numbers in the same hole have the property that one is the multiple of the other.
164 THE ART OF PROBLEM SOLVING
First, consider the 50 odd numbers from the natural number set {1,2,3,, 100}:
1,3,5,..., 99. Clearly, any number in the natural number set {1,2,3,..., 100} can be
represented as the product of one number from the odd number set {1, 3,5,..., 99}
and some power of 2. In other words, we can separate the natural number set
(1, 2, 3,..., 100} into the following 50 groups:
Aso = {99}.
Every number in the natural number set {1, 2, 3,..., 100} belongs to one of
the 50 groups. We can consider every group as a hole and then randomly pick any
51 numbers from the natural number set {1, 2, 3,..., 100} and put them into these
50 holes. Therefore, there must be at least one hole that contains 2 numbers. For
any 2 numbers in every hole the bigger number must be a multiple of the smaller
number.
There are 9 points in a square whose sides have length 1. Prove: For
GEOM the triangles formed by these points, there must exist a triangle whose
area is less than 1/8.
In this problem, what is going to be considered as a hole and how many holes
exist is not yet clear. Therefore, first we need to construct holes. Before we do that,
we need to prove a corollary.
Corollary 1. For any three points in a rectangle with area a, prove that the area of the
triangle formed be connecting these three points is less than a/2.
Proof Because the largest triangle that can be drawn inside a given rectangle is
one whose vertices lie on the sides of the rectangle (see Figure 8.1) and has an area
half that of the rectangle, any triangle formed joining three points inside the triangle
must have a smaller area than the previous triangle and thus less than half the area
of the rectangle.
The Pigeonhole Principle 165
From Corollary 1 we know that for this example we only need to prove that there
exists a rectangle of area 1/4 that contains at least three points.
Proof. Draw three lines parallel to one side of the square, and separate the square
into four equal rectangles (Figure 8.2). Consider each small rectangle as a hole.
Now we have 4 holes. Put these 9 points into these 4 holes. According to the
Pigeonhole Principle, there is at least one hole that contains at least 3 points. From
Corollary 1, the area of the triangle formed by connecting these three points is less
than 1/8.
Proof. We may use the same method we used in the previous example. Connect
the midpoints of the sides of the equilateral triangle. The original equilateral
triangle has then been separated into four equilateral triangles (Figure 8.4). Let s
consider each equilateral triangle as a hole. Then, according to the Pigeonhole
Principle, there are at least 2 points in one of the holes. Since we know that all 5
points are in the original equi- lateral triangle, and the length of sides in the small
equilateral is 1/2, the distance between the 2 points in the same hole is less than
1/2. This proves that, given 5 points, there are at least 2 of these points whose
distance apart is less than 1/2.
There are 15 seats around a round table, and there are 15 customers’
names posted around the table. Assume that no customers notice the
name signs until they sit around the table and realize that no one is
sitting in front of his or her own name sign. Prove that we can rotate
the table so at least 2 customers are sitting in front of their own name
signs.
Proof. We can separate the round table into 15 equal blocks and use a\, ai,. .., a\s
to represent the 15 customers' name signs (Figure 8.5a). Pick any customer from
GAME
these 15 customers and use A to represent her. Assume that A initially sits in front
of post a\ (Figure 8.5b). From Figure 8.5b we should be able to establish the
following: Whenever the table rotates one notch clockwise, there will be a name
sign pointing to the seat where A sits.
The Pigeonhole Principle 167
Figure 8.5b
After the round table has moved 14 notches clockwise, the 14 name signs on
the table (except apointed at different times to the seat where A sits (Figure 8.6).
These rotation movements have all 14 name signs (except the one in front of
which the person originally sat) point to every customer at the table, one at a time.
According to the initial assumption, no customer sits in front of his or her
name sign. In other words, at is not customer A's name sign. And customer A's
name sign is among the other 14 name signs besides a1 Then among these 14
rotations is one that has customer A sitting in front of her post.
This proves that every one of these 15 customers was situated in front of his
or her name sign in one of these 14 rotations.
Now consider every one of these 14 rotation positions as holes, and the zth
(i = 1,2,..., 14) rotation position is called the "z* rotation hole." If at the zth rotation
position customer A is sitting in front of her name sign, then put A into the zth
rotation hole. According to the Pigeonhole Principle, there are at least 2 customers
in the same hole. In other words, there must exist at least one rotation position
such that 2 customers are sitting in front of their respective name signs.
168 THE ART OF PROBLEM SOLVING
Figure 8.7
Before we solve this problem, let's analyze the question: According to the
coloring requirements, every block must be colored by exactly one of the colors,
black or white.
Figure 8.7 shows a checkerboard with 4 rows and 7 columns. The dashed
outline delineates a rectangle having all four comers the same color. The four
like-colored blocks (in this case, white) shown in Figure 8.7 are positioned in the
second and third rows and the first and fourth columns. The dashed rectangle may
be called a "same-color-comer rectangle."
Part 1
Proof. (Method 1). Whichever way we color the checkerboard, we can always find
the desired rectangle considering only three rows on the checkerboard—that is, a
3x7 checkerboard.
The Pigeonhole Principle 169
We shall first consider the first row. When we color the 7 blocks in either
black or white, then according to the Pigeonhole Principle there are at least 4 blocks
having the same color. Let's assume that these 4 blocks are colored white. Without
loss of generality, we can assume that these 4 white blocks are positioned at the
left side of the first row (Figure 8.8).
Now let's consider the other 2 rows. Since we are interested in the same-color
blocks, we will consider only the first, second, third, and fourth columns and not
consider the fifth, sixth, and seventh. If there are at least 2 white blocks in one of
the other 2 rows, then these 2 white blocks, together with the 2 corresponding
blocks in the first row, will be the four comers of the rectangle. This is a white
same-color-comer rectangle. If every row has at most 1 white block, then delete
the column(s) where they have the white block(s). The remaining 2 columns will
form a black same-color-comer rectangle.
w w w w
Figure 8.8
Proof. (Method 2). There are 21 blocks in the 3x7 checkerboard (Figure 8.8).
Consider the black blocks as black holes and the white blocks as white holes.
Therefore, according to the Pigeonhole Principle, there are at least 11 blocks with
the same color (say, white).
Now, let's consider every row as a row hole. Put these 11 white blocks into
these 3 row holes. There are at least 4 white blocks in one of the row holes. In other
words, at least one row contains 4 white blocks. Assume these 4 white blocks are
positioned in the front of the ?'th row. Now, as in the previous proof (method 1), we
can conclude that there must exist a same-color-comer rectangle.
Part 2
We begin by coloring 2 of the 4 blocks in white in each column, and the other
two blocks will then be colored black. There are =6 different colorings possible.
This applies exactly to the 6 columns. Figure 8.9 is one such example, where there
is no same-color-comer rectangle there.
170 THE ART OF PROBLEM SOLVING
Figure 8.9
Exercises
1. Put any 6 integers into the 6 blocks (Figure 8.10). Prove that there exists
a rectangle for which the sum of the four corner numbers is an even
number.
7. Assume that everyone in New York City has at least 1 friend. Prove that
there are at least 2 people in New York City who have the same number
of friends.
8. Select any n + 1 numbers from the integers 1 through 2n. Prove that there
exist two of these n + 1 numbers whose difference is n.
9. If there are more leaves than trees, then there are at least 2 trees with
the same number of leaves. Is this true?
10. Choose any n + 2 numbers from the integers 1 through 3n. Prove that
when n > 1, there must exist 2 numbers whose difference is between n
and 2n.
11. On a 3 x 9 block checkerboard, color the blocks red and blue. Prove that,
whichever way you color them, there must exist 2 columns that have the
same coloring.
12. Using 4 colors, color a 5 x 41 block checkerboard. Prove that, whichever
way you color the blocks, there exists at least one same-color-corner
rectangle.
13. Color the blocks of a 5 x 5 block checkerboard red and blue. Prove that
there exists at least one same-color-corner rectangle.
14. Separate the positive integers from 1 through 67 into 4 parts. Prove that
there must exist a positive integer in one part that is equal to the difference
of two other numbers in the same part.
15. Select any n + 1 numbers from the positive integers 1 through 2n. Prove
that there must exist a pair of numbers in the selected group such that
one number is the multiple of the other.
9
EVAN M. MALETSKY
One of the great contributions of the early Greek mathematicians around the
time of Pythagoras, Plato, and Euclid was to move an imperfect, hands-on mathe¬
matics based on experience and induction to a perfect and ideal, abstract mathe¬
matics built on deductive reasoning. They were very effective in establishing this
view, and it has remained essentially unchanged to this day. Unfortunately, this
unique mathematical strength may also be the cause for much of our current
concern in mathematics education.
The National Council of Teachers of Mathematics (NCTM), in its Agenda for
Action in the 1980s, identified the primary focus of the mathematics curriculum in
the schools to be problem solving. Its Teaching and Evaluation Standards for School
Mathematics of the 1990s broadly expanded this same theme by adding reasoning,
communication, and connections. These recommendations give direction for the
ideal experiences in the mathematics classroom, and they clearly incorporate
handling and seeing as well as thinking.
Unfortunately, many believe these three terms form an ordered sequence.
173
174 THE ART OF PROBLEM SOLVING
Furthermore, it is seeing that serves as the vital link between handling and
thinking. If we are to enhance the problem-solving skills of our students, one thing
we must improve is their ability to see properties, patterns, and processes, whether
they be in numerical, algebraic, or geometric form.
How often we assume, without question or doubt, that our students see what
© we see. How often we assume that because the numbers are there in a pattern to
be seen, they see the pattern; that because all the words are there to be read, they
GEOM
see the problem; that because all the parts are there in a diagram to be seen, they
see the whole of it. Many years of experience in the mathematics classroom have
revealed an increasing deficiency in our students' ability to see and think in a
visual, geometric form. It is our responsibility as mathematics teachers, regularly
and repeatedly, to express numerical and algebraic relationships geometrically.
Not only will this develop a better understanding of how these different fibers are
interwoven into the fabric of mathematics, but it will also make geometric visual¬
ization a familiar and ready tool for problem solving.
Benoit Mandelbrot, considered by many to be the father of fractal geometry,
writes: "Today, I believe deeply that science could accommodate more variety,
and badly needs more people with a sharp and happy roving eye."
This is no idle wish. It is an apparent and frightening observation and should
be a major concern to teachers of mathematics, especially when one considers the
role the sense of sight can play in problem solving. The dilemma can be simply
stated: Too many students look but do not see.
Do we train our students to look for answers with their eyes? And if we do,
do we also encourage them to rove around and find what else they can see? Stu¬
dents learn from the way they are taught. They also see by the way they are shown
to look.
Perhaps most important of all is the ability to see numbers in geometry and
ARITH geometry in numbers. As a case in point, consider the following question from a
recent National Assessment of Educational Progress (NAEP) test at the junior high
level.
Surely, percentages remain one of the difficult topics to master at this level,
but should this question, with these friendly numbers, be missed by more than
half those trying to answer it correctly? Could it be that they looked but saw
nothing? Could it be that the various algorithms used to teach percentage prob¬
lems of this type did not include any that were visual? Perhaps this is why so many
students failed to recognize these numbers as being friendly and, hence, could not
associate the problem or its solution with any simple geometric figures.
The figures below all connect to the percentage problem above. It is the
responsibility of the teacher, especially in the middle-school years, to show this
connection not only through pictures to see but through pattern block, tangram,
and paper-folding activities where things can be handled as well.
Not only should the connection to geometry be made when doing friendly
computation problems with percentages, fractions, decimals, and whole numbers,
but the connection should also be made to these computations when doing the
related work in geometry. Furthermore, this visual connection helps focus in on
the fact that both the 75% and the 12 are friendly numbers for computation. One
wonders how many of those taking the NAEP test even noticed this fact. We need
to look carefully at how we teach what we teach and how we help students see
when they look. Could it be that many students would see no difference between
these two percentage problems and how to solve them?
One of the great skills in problem solving is the ability to look for differences
as well as similarities and to know what to do with them when they are found.
The following problem appeared as a calendar problem in the NCTM Mathe¬
matics Teacher. It is used here to illustrate how we can enhance our students ability
to see numbers and patterns in what appear to be straightforward problems in
geometry.
176 THE ART OF PROBLEM SOLVING
GEOM
Figure 9.5
Two different skills are required here. One is the ability to see that there are
three different sizes of triangles. The other is the ability to count systematically all
the possible positions for triangles of each size. Cutting out a triangle of each size
and moving it around on the figure can be a helpful strategy here. Some triangles
point up while others point down. It may be less obvious that there are the same
number of each kind in each size, but the horizontal axis of symmetry forces this
to be true.
Table 9.1
What else is there to see and do here? These follow-up questions illustrate
just how many additional problem-solving strategies can easily be brought into
play from this single problem.
What fraction of the area of one large triangle is shared with the other?
If a triangle is chosen at random, what is the chance that it is a small one?
What is the relationship among the perimeters of different-sized triangles?
How are the areas of the different-sized triangles related?
Do all pairs of middle-sized triangles share the same percentage of their areas?
Does the figure contain more triangles or trapezoids?
Are there any special number patterns in the figure?
the triangular pieces in a set of pattern blocks. In the figure on the left, the entire
triangle is built with blocks. On the right, only those blocks that point up are used.
Have students search for geometric patterns. Then have them set up tables
and look for numerical patterns that relate the number of blocks in given rows of
the figures. Finally, have them extend the patterns for more rows. On the left the
odd numbers are generated. On the right the counting numbers appear.
Table 9.2
Row 1234512345
NUMB
Have students look again for patterns, this time among the numbers giving
totals from the top through a given row. Surprisingly, the perfect squares are
generated on the left while the triangular numbers are generated on the right.
Table 9.3
Through row 1 2 3 4 5 1 2 3 4 5
Number of blocks 1 4 9 16 25 1 3 6 10 15
Once the separate number patterns are explored, let students investigate how
they relate to one another as the sequence is extended. For example, as more and
more rows are added to the array, what happens to the ratio of triangular holes to
triangular blocks? Does it remain the same, increase, or decrease? If it changes, is
there a limiting value approached as the number of rows increases without bound?
Table 9.4
Number of rows 1 2 3 4 5 6
Holes 0 1 3 6 10 15
Blocks 1 3 6 10 15 21
Ratio of holes 0 1 1 3 2 5
to blocks 1 3 2 5 3 7
Figure 9.8
178 THE ART OF PROBLEM SOLVING
As the array gets larger and larger, the ratio of triangular holes to blocks
increases toward 1.
Note how the dynamics of mathematics have been built into this last activity
by looking not only at specific ratios but at how these ratios change. This sequenc¬
ing of stages is the same kind of dynamics that can be built into the six-pointed
figure from the original counting problem. Is it a static figure or is there some action
involved? The answer, of course, depends on what you see. So let us look at it
again.
Figure 9.9
Challenge students to do the mental visualization first. Then have them cut
the figure from paper, fold in the petals, and open it in their hands. Many students
need to see the essence of a problem by drawing a picture or making a model. This
flowering bud can serve as the basis around which some interesting questions can
be built. Here are two examples.
GEOM If the triangular petals may or may not open, how many flowers are
possible in all, and in how many different shapes can they occur?
There are two choices for each of the six petals, open or closed. Thus, in all,
there are 26 = 64 possibilities. Visualizing the 13 possible shapes is a bit more
challenging.
If there is only a 90% chance that a petal actually opens, what is the
probability that the flower has the shape of a trapezoid?
First, see the she different ways a trapezoid can be formed from two petals
open with one closed in between. Then multiply by the corresponding prob¬
abilities for the open and closed petals.
Experiences in Mathematics 179
Imagine some paper cut in the shape of regular hexagons. With one,
fold an opposite pair of vertices to the center. With a second, fold alter¬
nate vertices to the center. With a third, fold all vertices to the center.
What shapes do you get? What fractional part of the original area
remains in each case?
Visualize the foldings first. What mental images do you see? Then fold with
your hands and look with your eyes. Again, models can help in solving problems.
Next, build a supporting argument for why the shapes are what they are. Finally,
look at them numerically and find the resulting areas as fractional parts of the
original hexagons.
GEOM
ARITH
soning.
180 THE ART OF PROBLEM SOLVING
A Sample Problem
COMB A 9-inch piece of wire is bent at two points such that its ends come
GEOM
together to form a triangle. If the bending points must be on the inch
marks, how many choices are possible?
Bring a piece of wire for the students to handle. Let one bend the wire so
others can see a solution. Then let them think in groups about how to find all
possibilities.
A typical strategy might be to find the different integer dimensions possible
for the triangles by systematically accounting for all possibilities in a list.
Table 9.5
Length Classification
Isosceles 1-4-4
Scalene 2-3-4
For some, the work ends here with an answer of 3. There are three different
triangles that can be formed. For others, a careful rereading of the problem reveals
the need for a second step that leads to an answer of 10 possibilities. The list shows
10 different points where the wire can be bent in forming these triangles.
Table 9.6
Here are the three choices of isosceles triangles. In each case, the wire has
been bent about the shortest 1-inch base, which has remained fixed in a horizontal
position.
Experiences in Mathematics 181
The first approach erroneously implies the three classifications are equally
likely. The second approach shows that the chances are 3 in 10 that an isosceles
triangle is formed, assuming, of course, that all 10 choices of bending points are
equally likely.
As we as teachers do more with problem solving in the classroom, we become
more comfortable with extending and adapting problems. Invariably, one good
problem leads to another.
What if the initial wire were 10 inches long? Would there be fewer, the
same number, or more folding choices for triangles?
The answer may come as a bit of a surprise. The wire is longer but there are
fewer choices. The 9-inch wire gives 10 choices while the 10-inch length gives only
6 choices, all of which are isosceles.
Table 9.7
Our goal as mathematics teachers is not only to improve our students' skills
in problem solving but also to motivate and encourage them to search out, recog¬
nize, and explore problems on their own. The triangle problems with the 9- and
10-inch wires were only the appetizers. The main course comes when we look for
an underlying rule for a wire of any integer length.
182 THE ART OF PROBLEM SOLVING
NUMB
How and where does one begin? One approach is to guess and test some
different ideas. Another is to quickly solve some simpler analogous problems with
COMB very short lengths, such as with 3-, 4-, 5-, and 6-inch wires. The respective answers
of 1, 0, 3, and 1 may not reveal any clues. There are many different approaches to
try and many interesting connections to observe along the way.
Do you see anything special? For many, this means looking, in their minds,
at the triangles formed. Far fewer look at the answers of 10 and 6, along with the
1,0,3, and 1, and see if they can recognize or learn anything from them. Are these
special numbers? Is there a conjecture that you are willing to risk making and
checking?
The numbers 1, 3, 6, and 10 are all triangular numbers. Could there be any
connection? Try some other cases and see. An 8-inch wire gives 3 choices, and an
11-inch wire gives 15 choices. That's both interesting and encouraging. They, too,
are triangular numbers, but the order in which they appear seems strange.
This is the place to send students off on their own for some more counting
and have them return (we hope) with an observation or maybe even a generaliza¬
tion. This is also the place to encourage them to try adopting different points of
view, using numbers, algebra, or geometry with tables, formulas, or diagrams.
One possibility is to do some additional counting for other lengths to com¬
plete a sequence. Then collect the data and present them in the form of a table.
Here is a table of all choices for lengths from 3 through 15 inches. There is a striking
pattern to be found, but not everyone will see it.
Table 9.8
Length 345678 9 10 11 12 13 14 15
Choices 1 0 3 1 6 3 10 6 15 10 21 15 28
1 = 1
3=1+2
6= 1 +2 +3
10 = 1 + 2 + 3 + 4
15=1+2+3+4+5
Had we approached the initial listing process from a different point of view, /'J'n
we might have more quickly seen the relationship to the triangular numbers.
Instead of listing by classification according to sides, this method exhausts all those
with an initial bending point before moving to the next possibility. Notice how the
10 choices for the 9-inch wire are readily seen as the sum of 1, 2, 3, and 4
possibilities.
12 3 4
5 X X X X
6 XXX
7 X X
8 X
184 THE ART OF PROBLEM SOLVING
A Geometric View
The study of graph theory can help students view situations in geometric
form. Graphs are pictures that offer a highly visual approach to presenting infor¬
mation, and they can be very useful in solving many problems.
Consider again the triangle-folding problem with the 9-inch wire. The three
vertices of the triangle come from the two bending points on the wire plus the
joining of the two endpoints into a single third vertex. The 10 choices of bending
points are show here as connected graphs on circles divided into 9 units corre¬
sponding to the 9-inch length of wire. The three arc lengths on each circle corre¬
spond to what the actual lengths of the three sides would be.
Why look at this problem in another way with pictures when the general
pattern has already been discovered? The reason is that a different point of view
often sheds new insight into a problem. It was the great mathematician and
educator George Polya, who has been quoted as saying, "It is far more important
to do one problem five different ways than to do five problems all one way."
As we look at the essence of problem solving in the mathematics classroom,
it is wise to keep this idea in mind. Be careful not to measure success only by how
many problems are done, but measure it also by what is done with those we do.
Consider again the 10 graphs showing the solutions for the 9-inch wire. What
separates these from any choice of 3 of the 9 points on the circle? First, we use the
point 0 to note where the two ends of the wire come together. The wording of the
problem specifies choosing only two bending points on the length of the wire.
Combinations can be used to count the 28 possible choices of number triples
that contain 0 as one of the choices. To do this, count the number of pairs that can
be chosen from the remaining 8 points that can be connect with the point 0.
Experiences in Mathematics 185
8x7 COMB
28
2x1
v y
What distinguishes the 10 solutions from these 28 possible triples? Look at
their graphs again, carefully. What do your eyes see? The sense of sight can be so
important in the problem-solving process. We look, but what do we see?
Some students' eyes only see the metric, measurement aspects of geometric
figures. Others see graphs only in terms of points on a coordinate system. We seek
different views here. The eyes may first spot that the center point of each circle lies
within the triangle. Seen through another set of glasses, all these triangles are
acute. What is the connection between these two observations, and how do they
relate to the original problem?
Deductive reasoning comes into play here, recalling that the sum of the
lengths of any two sides of the triangle must be greater than that of the third side.
This means that the two bending points cannot be on the same side of the midpoint
or at the midpoint of the wire, since the midpoint is not a unit point. Translated
into our graphs drawn on circles, this requires the center of the circle to be enclosed
within the triangle; hence, the triangle on the circle must be acute. Right triangles
on the circle would contain the center point, since the hypotenuse would neces¬
sarily be a diameter. Obtuse triangles on the circle would leave the center point
outside.
Does this mean that you can never form a right or obtuse triangle by bending
a wire at the inch marks? The answer, of course, is no. Like many diagrams in
discrete mathematics, our graphs on circles show connections and relationships in
a discrete sense. The triangles shown on the circles are not the actual triangles
folded from the wire.
For example, bending points at 2 and 5 form a triangle with sides of 2,3, and
4 inches. The graph showing this choice of points may appear to be isosceles, but
the actual triangle is scalene, since no two sides are the same length.
The shortest wire length for a right triangle is 12 inches. Six of the 10 choices
are scalene right triangles formed from the Pythagorean triple 3-4-5.
186 THE ART OF PROBLEM SOLVING
Table 9.9
The reader is asked to find the shortest integer length needed for an obtuse triangle.
As a final view of the problem, consider drawing all the triangular-connected
graphs on the same circle. Better yet, since all triangles must contain a vertex at 0,
only connect the remaining two vertices from each of the 10 graphs. Each chord
now identifies a unique choice of bending points on the 9-inch wire.
1 + 2 + 3 + 4 = 10 choices in all.
Figure 9.28
This graph can give us insight into the connection between the triangular
number pattern for odd lengths and the one for even lengths.
The even length that also has 10 choices is 12 inches. When the wire length
is odd, the first possible bending point is at point 1. But when the wire length is
even, the first bending point is at point 2. Furthermore, when the length is even,
the midpoint of the wire is a marked point but cannot be used. So 3 of the 12 points
on the circle cannot be used: the first and the last points (1 and 11) and the midpoint
(6). Thus the counting problem for the 12-inch wire reduces to that for the 9-inch
wire where all 9 points can be used.
Notice that the graph for the 12-inch wire has the same characteristics and
the same number of solutions as the one shown above for the 9-inch wire.
12-inch wire 0
1 + 2 + 3 + 4 = 10 choices in all.
Figure 9.29
7 T 5
6
Experiences in Mathematics 187
In each case, even lengths yield the same results as odd lengths 3 inches
shorter.
Table 9.10
The diagrams illustrate the connection between the two triangular number
patterns for odd and even lengths of wire. Another approach might be to show
algebraically the two corresponding formulas. They yield the correct numbers but
obscure the connection both to the triangular numbers and to each other.
The k,h triangular number is the sum of the first k positive integers.
and vision of mathematics is likewise expanded and they are more likely to stretch
to extend their reach yet again.
Manipulatives offer hands-on, visible objects that later become visual images
for mental thought. They open up a whole new realm of geometric thinking, a skill
that many problem-solving strategies require.
Manipulatives also help give students a dynamic view of mathematics. They
encourage a sense of action, motion, and change—key components in the problem¬
solving process, where flexible views and alternate approaches must be freely
tapped and tried.
The best manipulatives to use as settings for a problem-solving experience
are usually the simplest. In this example, they serve to verify and reinforce as well
as to explore and discover.
Have students think first and write down what they visualize as possibilities.
Let then share and discuss their views and conjectures next. Then let them cut and
fold and explore with their hands.
The initial challenge is to visualize in their minds the different possibilities.
It is unlikely that many students will identify all three classifications of solutions.
It many even be that they do not find them at the second step when cutting and
folding with their hands. This step, with the actual manipulatives, offers insight
and reinforcement and perhaps a surprise. But most of all, it motivates further
investigation. Some students will study the shapes they get as results and look for
the cause. Others will immediately start cutting and folding more triangles, hoping
for an underlying rule to reveal itself to them. The reader is encouraged to explore
and analyze this problem before continuing.
Pentagon
For many students, this will appear as a rather abstract analysis. A more
concrete observation, made from a hands-on exploration folding paper, deals with
side lengths.
Let sides a and b contain the vertex to be folded with a < b. Let c be
the opposite side.
It usually comes as a surprise that the result depends on the lengths of the
sides of the triangles and not on the sizes of the angles. Each of the three cases of
quadrilateral, triangle, and pentagon can be folded from an acute, right, or obtuse
triangle. But, needless to say, different vertices may fold into different results.
190 THE ART OF PROBLEM SOLVING
GEOM
This certainly appears to be a hands-on activity. Just cut out a lot of triangles
STAT at random and see what proportion are acute. If the question is asking about actual
results from actual cutting, collect some data and study the results. This is often a
very powerful problem-solving strategy.
Alas, we so often require in mathematics a theoretical approach that is ideal
and abstract, deductive rather than inductive. After all, how else can randomness
really be considered without personal biases? However beautiful such solutions
are, they remain, from one point of view, impractical. Yet they illustrate the essence
of mathematical thinking so we offer one approach here. (There are others, some
of which yield different results.)
Here is one analysis based on deductive reasoning involving algebraic rela¬
tionships. Let the three angles of a randomly chosen triangle have degree measures
ALG
of A, B, and C. That implies these conditions.
What do these inequalities look like when shown on a graph? The sample
space is a large right triangle where every point in its interior corresponds to an
ordered pair of values (A, B). Assuming all such points are equally likely, selected
areas can be compared to find appropriate probabilities. The areas, associated with
the three separate conditions, are shown first, with their intersection given at the
far right.
Experiences in Mathematics 191
Many students will be able to establish the necessary inequalities but may
not know how to use them to get the needed probability. Others will draw the
graphs but may not be able to translate the results into a probability because they
do not see all that is there when they look at the graph. The four small triangular
regions are all equal in area but only one of them contains points that satisfy the
inequalities required for an acute triangle. It follows that the theoretical probability
of randomly cutting an acute triangle under these assumptions is 1/4. In similar
fashion, we conclude that the probability of cutting an obtuse triangle is 3/4. Now
this very same sense of sight that was so useful thus far challenges us to accept 0
as the probability of randomly cutting a triangle that is a right triangle.
The ability to visualize in the mind comes, in large measure, from the ability
to see with the eyes. To a great extent, visualization is an extension of the sense of
sight. There are many times in mathematics when perfect, abstract things are seen
192 THE ART OF PROBLEM SOLVING
•in the mind as extensions of imperfect, concrete things that we can handle with
our hands and see with our eyes. The key components of Euclidean geometry—
points, lines, and planes—are examples, as are circles and squares and spheres and
cubes. Indeed, the very triangles that have been mentioned throughout "this
chapter are examples as well. We view them once again, but this time as part of a
dynamical system. Students need experiences seeing changes that result from
dynamical, iterative processes, especially since sequencing is, in fact, another
problem-solving strategy.
Start with a triangle cut from paper. Connect the midpoints of the
triangle to form four smaller triangles. Keep the three corner triangles
but discard the middle one. Then repeat the rule with every new triangle
at each and every stage.
GEOM
Stage 0 1
Here are some counting, measuring, and geometric views of this sequence.
A Counting View
Table 9.11
Stage 0 1 2 3 4 Stage n
Triangles 1 3 9 27 81 3”
3” -1
Holes 0 1 4 13 40 2
30 + 31 + 32 + 33 + 34 = 1 + 3 + 9 + 27 + 81 = 121
The ratio can be computed at any stage, but another approach is required to /'Jn
find its limiting value. Each successive ratio can be shown to be closer and closer V©/
to the fixed point attractor of 1/2.
A Measurement View
Again, a tabular array is used to investigate these relationships. Note that for
convenience, an arbitrary unit of area and of perimeter has been assigned-to the
stage 0 figure.
Table 9.12
Area
Perimeter
1
1
3/4
3/2
9/16
9/4
27/64
27/8
81/256
81/16
(3/4)n
(3/2)n
©
At stage 5, the total area is (3/4)5 = 243/1024 = 0.237. At stage 5, the total ©©
perimeter is (3/2)5 = 243/32 = 7.594.
The dynamics of change are intriguing here. While the area is decreasing
from stage to stage, the perimeter is increasing. As the iteration continues, ad ©©
infinitum, the area shrinks towards zero while the perimeter increases without
bound.
194 THE ART OF PROBLEM SOLVING
A Geometric View
We can draw the first few stages with our hands. We can see and extend the
geometric pattern with our eyes. We can devise a new three-step transformation
algorithm for generation.
But can we stretch our minds to visualize the strange attractor that is being
approached by these successive stages?
Here we must see with our minds what we cannot see with our eyes. As the
iterative sequencing continues, more and more of the area of the original triangle
is eaten away. Each and every triangular region at each and every stage is being
reduced in successive stages. At the final limit state, the original triangular region
has been completely eaten away with holes so no area at all remains. Each small
triangular region is ultimately reduced to a point with an area of 0. But there still
is something there!
Infinitely many line segments remain, forming a delicate, endlessly repeat-
mg/ ever-diminishing, highly complex fractal framework. To help in this visual¬
ization, notice that the segments that form the boundaries of the triangular regions
at each stage remain as part of this final figure. This strange attractor, the resulting
fractal, is called the Sierpinski triangle. It is named after the Polish mathematician,
Waclaw Sierpinski, who discovered it in 1917.
One of the striking properties of the Sierpinski triangle is that it possesses
strict self-similarity. Take any small piece of it, and embedded within that piece
will be an exact but reduced replica of the original figure. Another property of this
and all fractals is that its fractal dimension can be measured. The Sierpinski triangle
is more complex that a line segment with dimension 1 and less complex than a
region of the plane with dimension 2. Hence, we would expect its fractal dimension
to lie between 1 and 2, as it does with a value of 1.58.
Aspects of fractal geometry may soon be in the secondary school curriculum.
Not only is this topic accessible at these levels, but it also captures a new view of
mathematics as dynamic, active, and alive. However, it requires substantial pre¬
requisite experiences in the classroom, especially those that utilize the action and
motion of manipulatives. It builds on the handling, seeing, and thinking activities
of good problem-solving experiences of earlier years.
Fractal geometry offers students an excellent opportunity to search and rove
with their eyes and to stretch and flex with their minds. To offer a final illustration
of this point, study these results of an iterative process applied to squares. At each
Experiences in Mathematics 195
stage, every square region is replaced with three smaller ones such that the total
area is repeatedly reduced to three fourths that of the preceding stage.
Stage 0
Each and every figure consists of shaded square regions, increasing in num¬
ber and decreasing in area. But what is the strange attractor that is being ap¬
proached here with the square regions in these successive stages? Surely, it is some
self-similar fractal framework based on squares. As much as the eye tells us that
it is, it is not. The resulting fractal is the same Sierpinski triangle described earlier.
When viewed as a transformation algorithm, this one with squares behaves
exactly the same as did the one that started with triangles. The rebuilding code is
shown in the squares. Whatever the figure inside the square, when reduced to half
size, replicated three times, and rebuilt by that code, the resulting fractal will,
indeed, be the same Sierpinski triangle. The figure inside the square could be a
triangle or any other shape whatever. Look with your eyes at the first few stages
using the letter A and this same rebuilding code. Then see if you can visualize the
next few stages in your mind and recognize the strange attractor.
A
A
A A A
AA
A A
AA AA
A A
A A
AA
AA AA
A A A A
AA A A AAAAAAAA
Keep in mind the way the rebuilding is repeated at each new step. After the
reduction to half size and the threefold replication, the rebuilding code in¬
volves only translations of the squares containing the letters A. When rotations
and reflections of these squares are involved in the building codes, new shapes are
formed with new and different strange attractors.
196 THE ART OF PROBLEM SOLVING
d b d b d b
d F
F 1
d
d F
b J db lb
=1 F 1 FI F bd
d
IF IF dF 1
F
bd
How does one solve the problem of getting a glimpse of the general shape of
these final fractals? We can resort to yet another problem-solving strategy, that of
using a computer. With a computer, enough additional stages can be drawn so
that the eye can begin to see and perceive in those images the characteristics of
these fractals, the beautiful mathematical abstractions of the imagination, that are
evolving.
The computer-generated figures below show higher stages in the sequences
for the two square rebuilding codes that were used above. Already the general
shapes of the final fractals are clearly emerging. However, the computer is limited
to some small finite pixel size. Remember that it is only the final limit states of
these figures that are the real self-similar fractals.
Figure 9.56
Experiences in Mathematics 197
Figure 9.57
The real study of these types of fractals begins with their limit figures and
their self-similarity. Hidden in them are the building codes. The real visual expe¬
rience is in finding and identifying these transformations from their limit figures.
Fractals stretch the imagination because they are so highly visual in their
repetitive, diminishing, complex patterns. Computers become useful tools to help
us see into the depths of these figures and study the details that we cannot
otherwise see with our eyes. The focus of this chapter has been on the sense of
sight and its role in problem solving, and this visual experience with fractals
stretches that experience to the limit.
Conclusion
Fractals offer excellent examples of what the NCTM means, in its Curriculum
and Evaluation Standards for School Mathematics, by connections within mathemat¬
ics. They also bridge the gap between mathematics and the sciences and the arts.
Within the past several years, fractals have found applications in virtually every
discipline. They offer a new view of the world around us, a model perhaps more
198 THE ART OF PROBLEM SOLVING
powerful and appropriate even than the Euclidean models of ancient times. But
most of all, they develop two very critical skills in problem solving: insight into
geometric thinking and visualization.
As we teach, we want our students to work with their hands, to see with their
eyes, and to think with their minds. Our goal is to ultimately help them think
critically and to solve problems, not just in mathematics but in all areas. We want
them to apply what they know to real problems in everyday life dealing with real
things that are often imperfect, approximate, and concrete rather than perfect,
ideal, and abstract. We can help them achieve this goal though problem-solving
experiences in the classroom that focus on handling, seeing, and thinking. These
are what the mathematical experience is about.
Mathematics must tickle the senses as well as stretch the mind.
10
Problem Solving as a Continuous
Principle for Teaching
HANS HUMENBERGER
HANS-CHRISTIAN REICHEL
1. Introduction
199
200 THE ART OF PROBLEM SOLVING
important "for the test." This should start with the beginning of mathematics
classes and should recur as a continuous class principle. Furthermore, it seems
important to us that the students get enough time to deal with the problems
themselves and that various possibilities are discussed thoroughly and are really
understood by most of the students. There must be time for mistrials as well, and
these must not simply be considered as "wrong"; there has to be an explanation
why certain methods do not lead further—the problem must really be "digested"!
It is of only little use for students to be rushed from one problem to another, or to
have the teacher distinguish himself or herself with brilliantly demonstrated
solutions (which appear out of nowhere) and where the student cannot even guess
where the statement came from or why the teacher tried one way and not another.
Learning to solve problems surely does not depend on which problems the
students have to solve, as much as on how the lessons are shaped, what activities
are expected from the students, and what the climate is in the classroom.
But the treatment of problem solving as a "leading idea" for classwork is not
without difficulties, and we hope our chapter and this book will alleviate some of
these problems:
Working with over- and underdetermined problems is, in our opinion, a fine
means to "sharpen the spirit," to promote the ability of the students to better
recognize the structure (the essential things, the gist), and to translate (colloquially
formulated) texts into questions that can be answered in a mathematical way. This
capability is combined with the ability to apply mathematics—an ability every¬
body will need to solve his or her (mathematical) problems (in school and later,
too) and which we should therefore master!
A Continuous Principle for Teaching 203
Example 2.1
Here the students first have to find out which data are not necessary (fresh¬
man class, 4,300 inhabitants, 120 km, $500). Then they have to recognize the
mathematical structure of the task: The total spending consists of the fees for the
bus and rope-walk. The students should be able to find the formula for the
student's fee p basically by themselves: p - (360 - 110)/25 = 10.
Such examples can be embellished and extended almost limitlessly, and there
are no bounds for the teacher's imagination. The more sophisticated and complex
the story is and the more details (variables or concrete numbers) that appear in the
text, the more difficult it is, in general, to find the unnecessary data, particularly
if several quantities are asked for. (Many students can visualize a situation much
better with concrete numbers.) The students could also be asked to develop such
"task-stories" themselves (e.g., in smaller groups where students then solve each
other's tasks or as homework.) This may be a small contribution to promoting the
student's motivation, especially the very young student's (10 years old).
Analogous to this, tasks with too little information for the solution—"under¬
determined" problems—would be valuable too. Here the students would have to
figure out why the problem is not solvable, and they could make suggestions as
to what additional information is needed to solve the problem.
Reality—in contrast to math textbooks—often does not provide the full
amount of data needed to solve given questions. On the other hand, we often have
much more information than necessary. Therefore, to educate in the spirit of
applied mathematics, textbooks should copy reality in this respect (even artifi¬
cially) so that students have to figure out which and how many data have to be
known to answer the questions asked before they start working. Students should
also be able to say explicitly which additional data would be necessary. Such
problems are called over- and underdetermined problems in this chapter.
204 THE ART OF PROBLEM SOLVING
Example 2.2
Solution: A sketch showing the distances that were walked and cycled may
be very helpful (Figure 10.1).
© home
3
1 km walking
work
cycling
cycling
Figure 10.1 A cyclist who had some bad luck on his journey
To solve (a): He had to push his bike from the scene of the accident to his
work, but on his way back, he rode his bike exactly the same distance as he had to
walk earlier (therefore the distance does not matter). The distance from home to
the scene of the flat tire is covered twice by bike, once on the way out and once on
the way back, so he went 2 x 1 km = 2 km more by bike than he walked.
To solve (b): Everything is unnecessary except the 1 km! The value in
question (a) does not even depend on the total distance of the trip.
A Continuous Principle for Teaching 205
3
(i) —h = 12 min
(ii) 28 min
2
(iii) 4 —km/h
Example 2.3
Mr. Mayer usually takes the 7:25 bus from home to his office, and in
the afternoon he walks home (for health reasons). This takes him alto¬
gether 1 hour and 10 minutes. If he were walking both to and from the
office, it would take him 1 hour and 50 minutes, because on average
he walks only 15 km/h slower than the bus travels, with its scheduled
stops, traffic lights, congestion, and so forth.
Solution: For (a); As walking both ways takes him altogether 1 hour 50 min,
walking one way takes him 110/2 = 55 min. Therefore, one way takes the bus
70 min - 55 min = 15 min. Hence, both directions would take the bus 30 min.
(ii) 51 km/h
206 THE ART OF PROBLEM SOLVING
(iv) 5^ km.
The latter three solutions are to be expected at the earliest from 12- or
13-year-olds.
The following example will show, among other things, that it is possible, even
ALG
at an elementary level, to demonstrate how important it is to mark the right
quantities with variables.
Example 2.4
Forty-two birds are sitting on three trees. If 3 birds fly from the first tree
to the second one and 7 birds fly from the second tree to the third one,
there are twice as many birds on the second tree as on the first one
and twice as many on the third tree as on second one. How many birds
were sitting originally on each tree?
x + y + z = 42
2(x - 3) = y - 4
2(y - 4) = z + 7.
Their solutions would be x = 9, y = 16, and z = 17. But such equation systems
are solvable only for students who are at least 14 or 15 years old (at least in Austria).
So one could ask the students to look for a solution that does not involve solving
a system of equations. Alternatively, if this problem is stated already for 12- or
13-year-old students, then they are forced to find a different method anyway.
The situation, arithmetically, gets much easier if one looks at it first after the
birds' flights: If there were a birds on the first tree, there were 2a birds on the second
tree and 2 (2a) = 4a birds on the third tree. That makes altogether 7a birds, whose
total number is still 42. Therefore a—the number of birds on the first tree after the
birds' flight—must be 6 (42/7 = 6). So after the birds' flight, there were 6,12, and
A Continuous Principle for Teaching 207
Example 2.5
M,
1 800m
A X s' B
_77* b
M,
Figure 10.2
Again, we want to present two solutions that vary from each other in a similar
way as in the above mentioned example. Maybe it should be emphasized that a
sketch of the situation is very important for both solutions. A great part of the
solution is already here in a suitable depiction.
Solution 1. The described situation may pose quite a dilemma for most students,
as only very few things seem to be given and it may be not easy to find a
"statement."
208 THE ART OF PROBLEM SOLVING
Let us suppose that Ben, who is starting from B, is the slower one. Then the
first meeting point Mx must be nearer to B than to A; M1B - 800 m and M2A - 400 m
(Figure 10.2). It seems reasonable to let x be the length of the unknown piece,
M1 M2 of AB, the more so since the length of M2 M1 is required. FJowever, neither
of the two runners' speeds is known, nor are any times! As both runners started
at the same time, it takes them the same time to get to the two meeting points, their
running times from the start to Ml must be equal, and likewise from M1 to M2\ Let
v1 be Anton's average speed and v2 Ben's; then we get the following equations:
400+ x _ 800
V\ V2
But there are three variables in these equations! It is not possible to get three values
from two equations, is it? Did we not think of all the information? Is this problem
really solvable?
Indeed, in general, it is not possible to extract three variables (here x, vv and
v2) from two equations, and we did not forget anything. It is immediately clear
that the runners speeds cannot really be defined by the information given above:
If both runners, for instance, were running twice as fast, then they both would be
at the meeting points in half the time, but the meeting points would not change.
So one can expect only a statement about the relationship of both speeds (but this
is not really asked here.)
We surely cannot compute all three variables, but perhaps we could find at
least one—how about the value of x? This is indeed possible, by dividing the first
equation by the second equation (no occurring terms are zero); both speeds then
disappear from the equation and we get an equation in x;
400+ x _ 800
1,600+ x “ 800 + x
So we get the quadratic equation 320,000 + l,200x + x2 = 1,280,000 + 800x, with the
only positive solution being x = 800m. So the alley is 400m + 800m + 800m = 2,000m
long. The relationship of the two speeds we get by inserting:
1,200 800 3
- = - => V\ = —V2
vi v2 2
Solution 2. This solution does not need quadratic equations or, indeed, any equa¬
tions at all, so it could be seen as particularly "elegant" in that way. Arithmetically,
it contains, again, almost only elementary mathematics, but one has to see the
situation in a suitable way:
A Continuous Principle for Teaching 209
"At their first meeting both sprinters together ran the whole length of the
way exactly once." Have we already gained something by this? Not completely,
but we've obtained a basis for the next steps! What is the situation at the meeting
point M2? At M2, they have been running exactly three times the length of the alley
(see Figure 10.2)! Now, as both sprinters run at a constant speed, not only has the
total sprinted distance been tripled, but so have the individual distances run by
each of the sprinters. Therefore, Anton and Ben are running from the start to M2
three times the distance just to Mv That means that Ben has to run a distance,
from the start to M2, of 3 x 800 m = 2400 m. Of that distance, he was running already
400 m back toward his starting point. Therefore, the length of the alley has to be
2400 m - 400 m= 2000 m!
Although "text problems" and translating colloquial texts into the language
of mathematics refer, respectively, in most cases to setting up and solving equa¬
tions, not all described situations have to lead to equations. The latter kind of
problems are extremely suitable for classes, as they stimulate the activity of the
students and show the great variety of ways of thinking mathematically. Two
connected examples follow.
Example 2.6
There are six patients (A, B, C, D, E, and F) sitting in the waiting room
of a dentist. The dentist knows (on the basis of previous treatments)
what needs to be done for these particular patients and can therefore
guess how long it will take him to treat each one. The treatment of A
will last about 15 minutes; for B he estimates 30 minutes; for C and D
approximately 10 minutes each; for Eabout 20 minutes; and for Fonly
5 minutes. The dentist wants to keep the total waiting time of all patients
as low as possible and is thinking of an order in which to see the
patients. In this sense, is there an optimal order?
Students could (should) try some orders, but it will hardly be possible to
register all 6! = 720 possibilities in a systematic way.
How do the single waiting times and the total waiting time come about?
Already in the example above it can easily be seen: If a patient Y has his turn after
patient X, then the waiting time of Y obviously consists of the sum of the waiting
time of X and the treatment time of X! The patients Pv P2, ■ ■ ■, Pn have the times
for their treatments tv t2, ... , tn. If they take their turns in that order, the follow¬
ing waiting times emerge:
P 1:0
Pi: h
P3' h + ti
Pn- i: t] + t2 + . . . + tn-2
Pn' h + t2 + . . . + tn - 2+ tn -1-
The time for the treatment of the patient who is being treated first (fi) is
multiplied by the biggest factor (n - 1), thus giving it the "greatest weight" for
the total waiting time, as this particular time has to be waited by all following
patients. Now it can immediately be seen that if T shall become as small as possible,
then the biggest factor (n - 1) has to be multiplied by the smallest possible value
of t, (n - 2) has to be multiplied by the second smallest, and so on.
Result. The total waiting T time is the least if each time the patient with the shortest
treatment time—among all still-waiting patients—has his turn first. The (one) op¬
timal order for the initially asked problem is therefore
with a total waiting time of T = 5 x 5 + 4 x 10 + 3 x 10 + 2 x 15 + 20 = 145 min = 2 h
25 min.
So there is a waiting time saving of 2 hours compared with the alphabetical
order—a considerable difference!
Remark. If there are patients with the same time for treatment, there is, of course,
more than one optimal order. In our problem, the order
would also be an optimal one, as C and D have the same treatment time. We see
that there may be more than one optimal solution to a problem, so, often, one
cannot speak of the, but of one, solution.
Remark. Especially for young students, it is very important to learn that problems
can have more than one solution. On the other hand, it is difficult to create such
problems in an interesting way before equations of second or higher degree are
taught. But certainly, this is too late!
A Continuous Principle for Teaching 211
Example 2.7
for the consultation (compare with Table 10.1). In what order should
the senator call in the groups in order to keep the total waiting time of
all the individual members as low as possible?
The situation is again similar to the one at the dentist's office (but now with
groups rather than individuals).
Table 10.1
1 4 20 min.
2 8 10 min.
3 5 30 min.
4 10 15 min.
5 6 25 min.
in single persons and the group times into single times (times per person). The
above example gives a solution to the problem of an optimal order of single
persons: A person with the shortest possible consultation time has to take his or
her turn first. So the times per person determines the single order and in the sum,
therefore, also determines the order of the groups, because in an optimal order of
the individuals, the members of a group would certainly have their turn for the
consultation one after another according to the above principal and could be seen
in this respect as a group again. In other words, to call all persons with the currently
shortest consulting time into the room is the same as calling in the whole group in
its entirety first!
The consulting time per person in the single groups is indicated in Table 10.2:
Table 10.2
Group 1 2 3 4 5
Time per Person (min.) 5 1V4 6 lV2 4X/6
Therefore, the optimal order for the groups, according to the increasing times
per person, is G2 -) G4 -> G5 -) Gj -4 G3.
Solution 2. The following would be a slightly different way of dealing with the
problem (for instance, in the case where the example above was not used). Let
us reduce the problem first to the two groups Gi and G2. Is it better to take Gi first
or G2?
If Gx has its turn first, then the 8 persons of G2 have to wait 20 minutes each
(i.e., the consulting time of Gj), with a total waiting time of 160 minutes. If G2 were
to go in first, the 4 persons of G1 have to wait 10 minutes each, with a total waiting
time of 40 minutes, which is considerably less.
Let us now consider the general problem again: n different groups Gv ..., G„
with gp ... ,gn members and the consulting times tv ... ,t „ are sitting in the wait¬
ing room. Let us assume they take their turn in the order Gv ..., Gk, Gk+V ..., Gn to
see the senator, and this results in a total waiting time T1. What happens if any two
consecutive groups are interchanged? If one interchanges, for instance, the two
groups Gk and Gk + V we get the order Gv . . ., Gk+1, Gk,. . ., Gn and a new total
waiting time T2. There is obviously no change in the waiting times of any of the
groups not involved in this particular change {Gv ..., Gk _ t and Gk+2,..., G„); nor
does the combined waiting time that Gk and Gk+1 wait together for G1,...,Gk_1
change. Let the sum of these (in every case) constant waiting times be S. Then
consists of S and the waiting time of Gk+ 3 for Gk: T1 = S + tkgk + l
T2 is derived analogously if Gk + ^ has its turn before Gk: T2 = S + tk+l gk.
Under what conditions is Tk smaller than T2? Obviously, when
So if tk/gk (the consulting time per member of the group Gk) is smaller than
tk+\/gk+v it is more favorable to ask the group Gk to the consultation first.
Also from this we can deduce in general: If it should occur in a certain order
tk ifc + l
that the inequality — >-holds true for any k, it is possible to reduce the total
Sk Sk+ 1
waiting time by exchanging the two groups.
t1 f2 tn
Result. The order is optimal if and only if T- - T- - • • • - 7~- So the result here as
well is that the consulting time per person is decisive in finding the best order.
Especially for such problems, it is extraordinarily important to give the
students enough time for presumptions and trials and to talk about all suggestions
and arguments. When necessary, the students may be given little hints—the
teacher should not disclose too much and should encourage students to work
independently. The teacher, of course, has to lead or channel all problem-solving
processes to a certain degree, since the students will probably often choose wrong
approaches or will just not know what to do next. But this leading should not take
the form of an especially elegant solution that appears from nowhere.
This section will deal with some problems that illustrate the essence and the
importance of certain mathematical themes and that contain easily recognizable
"typical features" of these theories. Such (or similar) examples may sometimes
even serve as an entry into certain mathematical topics if they are suitable for
discovering the essentials of the new topic and if they motivate students to deal
thoroughly with the particular issue. Such entry problems must not be too sophis¬
ticated and should show clearly that an increase in knowledge (e.g., precision,
theoretical investigation, systematization, etc.) is necessary or at least possible.
Example 3.1
The task is now to move this “n-tower” from one rod to another while
obeying two rules:
What is the minimal number H(n) of single steps (moving of discs) for a
tower of n discs?
It is certainly a great advantage to have a few real trials with "model" towers of
objects that get smaller (e.g., books, pieces of paper). By trial and error one can
immediately make the following table for small n:
Table 10.3
n 1 2 3 4
H(n) 1 3 7 (15)
Many students will already begin having problems with n = 4, but for
students with even a little interest, H(4) = 15 will not be too hard to work out. But
now the question seems to be quite obvious, whether there is a "system," and
whether it is possible to get H(4), H(5), H(6),... by a thought experiment (e.g., from
the known values of H(l), H(2), and H(3), respectively). Let us go back one more
step: Is it already possible to get H(3) = 7 by consideration, without really experi¬
menting? Put a different way: How can I move a 3-tower if I know how to move
a 2-tower?
One can generally determine an answer easily from the concrete experi¬
ence in experimenting: To move a 3-tower from rod A to rod C, one first has to
move the upper 2-tower to rod B, then the lowest (largest) disc from A to C, and
finally the 2-tower "stored" in B to C (and that by employing the same scheme
used to move it before from A to B; see Figure 10.3). As it is known how a 2-tower
has to be moved and how many steps it takes—H(2) = 3—it follows that H(3) =
H(2) + 1 + H(2); 7 = 3 + 1+3, as the experiment confirmed already.
A Continuous Principle for Teaching 215
Now the structure of the problem becomes clearly visible: The same princi¬
ple must now be applicable for H(4): H(4) - H(3) + 1 + H(3) = 2 x H(3) + 1.
Analogously, to move a 4-tower, one has to move a 3-tower first, then the lowest
disc, and finally again the 3-tower. So now it is not especially hard to discover:
H(n) = 2H(n - 1) + 1.
The value H(n) is easy to determine if H(n - 1) is known, this in turn if
H(n - 2) is known and so on, until one gets to a known H(k). By this, one can
determine step by step: H(k + 1) = 2H(k) + 1, H(k + 2) = 2H(k + 1) + 1,. .., H(n) =
©
2H(n - 1) + 1.
In this manner, the above table can be conveniently continued: The last num¬
ber with which it seems sensible to conduct the experiment in reality is n = 6,
because starting at n = 7 the number of necessary single steps is already very high.
Table 10.4
n 1 2 3 4 5 6 7 8 9 10
One can see that this example very naturally leads to the principle of re¬
cursion or iteration. So it can be a gateway to the accompanying theory if the
student has not heard of it. Another theory demonstrated here would be that of ALG
the difference equations (such as linear difference equations with constant co¬
efficients), which are increasingly important for the description of dynamic sys¬
tems, especially those of first order (an = kan _ 1 + d) and perhaps those of second
order (an = k1an_1 + k2an_ 2 + d).
The principle of proof by mathematical induction can also be illustrated and
ALG
even motivated by this problem. Generally, it seems that proofs by induction
especially are often introduced without any natural motivation, without "discov- @
ery" and through a relatively dry lecture that doesn t help students really under¬
stand the necessity or the principle itself!
For big n (e.g., n = 50), of course, this step-by-step calculation of the values
of H(n) is relatively clumsy (at least without a programmable calculator or com¬
puter). So the question arises whether it is possible to express H(n) directly by a
formula without having to start at H(2) or H(3) and coming forward only gradu¬
ally. For this we should look again at Table 10.4. The powers of 2 (2,4, 8,16,32,64,
128, .) are very significant and impressive numbers, so one will soon notice that
the values of H(n) are always 1 smaller than 2n. Therefore, the assumption : H(n) =
2” - 1 seems justified. According to Table 10.4, it is true at least for n < 10.
Now, what does the case n = 11 look like? We know H(10) = 210 - 1, and we
further know that H(ll) = 2(210 -1) + I= 2x210 - 2 + 1 = 2 -1.
So our assumption is true as well for n = 11. We proved the validity of the
formula for n = 11 by the validity for n = 10. For n = 12 we therefore get H(12) =
2H(11) + 1 = 2(2n - 1) + 1 = 2 x 211 - 2 + 1 = 212 - 1.
216 THE ART OF PROBLEM SOLVING
One can already see that the validity of the formula is "inherited" from one
number to another (one could also use further concrete examples). It is already
perceptible, from these concrete examples, that this hereditary character does not
depend on the particular choice of n (e.g., n - 11 or n = 12). We could repeat the
above step to prove thereby the validity of the formula for any fixed n. But it is much
more convenient to show this hereditary character in general (i.e., detached from
concrete numbers): We want to show the heredity of the formula for all ne N. Let
us assume the validity of the formula H(n) = 2" - 1 for any n. It then follows that
H(n + 1) = 2H(n) + 1 = 2(2” -1) + 1 = 2” +1 - 1, and this is the validity of the formula
for the next number n + 1.
But if a formula is true for n = 1, and if the validity for any natural number
ensues from the validity for the previous natural number, then the formula must
obviously be true for all natural numbers:
We believe with this example, students could discover the theories of recur¬
sion and complete induction almost by themselves in a genetic way, and could
therefore perhaps understand them better (see Stowasser & Mohry, 1977, 7ff).
Other possibilities of proving the validity of H(n) = 2” - 1 would be, on the
one hand, through the sum formulas of geometric series (another theory) or, on
the other hand, by the following consideration:
We know the recursion equation H(n + 1) = 2H(n) + 1 with H(l) = 1 is true.
From that we get, step by step:
H{ 2) =2xl+l=2+l=2+ H( 1)
H(3) - 2(2 + 1) + 1 = 22 + 2 + 1 = 22 + H(2)
H( 4) = 2(22 + 2 + l) + l = 23 + 22 + 2 + l = 23 + H( 3)
By equating both terms for H(n + 1), one gets 2H(n) + 1 = 2" + H(n), from
which immediately follows H(n) = 2" - 1.
or projects. One of the most essential aims of this technique is to determine the
earliest possible completion date of a project, which consists of several "partial
tasks."
Mathematics can be seen in a certain way as a language, and its special
structure can contribute decisively to the solution of many problems. Its means
are, among others, variables, equations, and graphics (illustrations). Graphic pre-
sentations of all kinds especially serve to describe situations as they can also be
described by variables, equations, formulas, and the like; but graphic presenta¬
tions can often be realized better, more easily, and more quickly than texts or
representations with variables. "A picture is worth a thousand words."
The network technique is an excellent opportunity to make these graphic
aspects clear. Here, not only are graphic representations worked out, interpreted,
and described, but there is an essential additional factor: One can also see this
technique as an "intelligent" graphic representation in the sense that logical
connections, formalized relations, quantitative connections, and the like are illus¬
trated in such a way that the representation already gives a huge part of the
solution. Here is a very simple example that can be especially suitable as an entry-
level problem.
Example 3.2
Solution. The given information about the single PiPj contains information both
about which stages have to be reached before others and about how long each
single working process lasts. The graphic translation of the above information
leads to a netlike representation (Figure 10.4) called a "network," in which the
single working process is marked as an arrow between the single stages. Such an
representation gives a good survey of the situation and allows, in simple cases, a
solution by trial and error. (See Chapter 15.)
218 THE ART OF PROBLEM SOLVING
Since, for completion, all processes have to be done, one has to look for the
longest path from P1 to P6 to get the shortest total time. The other processes can
then easily be "fit in." This may present a slight paradox for some students. They
intuitively might have looked for the shortest path, but it is not difficult to see that
the jobs that last longer could not be completed then. By trying all possible paths
from Pl (start) to P6 (end)—adding all the working times along the paths—in
simple "networks" such as the one in Figure 10.4, one will quickly see that here
the "critical path" is > P2 -+ P3 —> P5 —> P6 ("critical" only because the jobs along
this path obviously must not be delayed to prevent an extension of the shortest
total time). The processes P2 —» P5 and P2 —> P4 -4 P6 can be fit in parallel fashion,
as they are shorter. The shortest time for the realization of the project (e.g.,
construction time) is therefore 4 + 6 + 7 + 3 = 20 time units.
Remark. If students have never been confronted with this kind ot task, even simple
ones such as the above can pose a problem in the sense that students have to look
for a (for them) new kind of solution.
Also, in slightly more complicated "networks" (e.g.. Figure 10.5), it is possible
to find one solution by trying different paths. As this example shows, there may
be several solutions (i.e., the critical or longest paths). In the network of Figure
10.5, the two paths P0 -> P1 -> P5 -+ P7 and P0 -+ P3 -> P4 -> P7 are both critical ways
with a total length of 15 time units each (see Reichel et al., 1989-1992, p. 171).
For the solution of very complex building projects, of course, this method of
trial and error will not be sufficient anymore. Some algorithms have already been
developed to determine the critical (longest) paths with the help of highly efficient
computers. These algorithms are not at all so complicated that they could not be
dealt with in school. There is really a theory behind such tasks, and even students
can gain a first insight into it.
A Continuous Principle for Teaching 219
P 5 P,
P, P
7
P P
We do not know if all students develop the right idea of the importance and
the nature of irrationality by the current examples in school (e.g., proof of the
irrationality of V2). What can it mean or what effect does it have if a number cannot
be expressed as a fraction (i.e., the relation of two whole numbers)? The following
problems, along with the solutions, will show some surprising results.
Example 3.3
Of course there are many straight lines that do not contain a lattice point in
the shown sector of Figure 10.6, but whether they do not meet a lattice point
"anywhere" is surely an exciting question!
220 THE ART OF PROBLEM SOLVING
Another formulation could be: Is it possible to fire from the origin an ideal¬
ized "arrow" (with diameter zero) through a "forest" with "trees" at the lattice
points (also with diameter zero) without ever hitting a "tree" (the "arrow" may
eternally maintain its original direction)?
Solution. If a straight line contains a lattice point with the coordinates (m, n) m,
n e N, it has the slope n/m. This means that if it is not possible to express the slope
of a straight line as a quotient of two natural numbers, then it cannot contain any
lattice points. The inverse is, of course, true as well: If a straight line has the slope
n/m, it will contain a lattice point, such as (m, n). So only the straight lines with
irrational slopes do not contain lattice points, and for this there are many more
possibilities than for straight lines with rational slopes. Although there is an
infinitude of straight lines with both rational and with irrational slopes, one can
formulate: "The majority of straight lines through the origin do not contain lattice
points" (noncountable infinitude in contrast to countable infinitude of straight
lines with lattice points).
To continue, it could be shown: If a straight line through the origin contains
one additional lattice point (other than [0, 0]) it must contain an infinite number
of them! What is the distance between two neighboring lattice points?
A Continuous Principle for Teaching 221
Solution. If (a, b) is a lattice point, then (2a, 2b), (3a, 3b), . . . are, of course, lattice
points, too. These are neighboring points if a and b are relatively prime. In this case
we get, for the distance, Vfl2 + b2 .
Example 3.4
Take any circle and mark a point P0. Then translate (rotate) P0 by the
angle a to obtain the point Pv Repeat the process to obtain P2, P3, P4,
. . . (see Figure 10.7). One of the points Pn may be equal to point P0 again.
In that case, our sequence of points is finite. On the other hand, the
sequence obtained by the process described can be infinite, too. Then it
consists of pairwise different points, of course. Now the question arises:
For which angles a do we get finite sequences, and for which a infinite
ones? In other words, for which a is there an index n such that Pn = P0?
Po
Solution. If Pn = Po, for some n, some multiple of a is a multiple of 360 , too; in other
words, there are natural numbers k and l so that the following is valid:
Example 3.5
A circle /C, with the radius r, touches another circle K2 with the radius
r2, and the points that touch are called P0 e /C, and P e K2, respectively
(see Figure 10.8a). By rolling along the circumference of Kv the point P
describes an arched curve called an epicycloid (see Figure 10.8b), where
the “next” points that touch (when P meets again exactly with /C,) are
called P1,P2, . . . Under which conditions of r, and r2 will the curve of the
point ever close? (When will the point Pever return to the starting point P0?
When is there some re Nwith Pn= P0?)
a) b)
Solution. Here one could rush to the conclusion: The curve will be closed only if
r\ and r2 have rational values, otherwise not. The curve will obviously be closed if
and only if both circumferences (u\ and w2) have a common multiple—that is, if
there are natural numbers k and / with (2ri7i) x k = (2r2n) x /. This is equivalent to
13J n x k = r2 x l or, alternately, r\/r2 = l/k. Therefore, it is not the irrationality of n or
r2 individually that is decisive here, but that of u\/u2 or r\/r2. If, for instance, r\ = 3k
and r2 = n, then the curve obviously closes exactly after one "round" of Ki or,
equivalently, after three resulting "arches" (e.g., see Figure 10.8b), though both
radii have irrational values.
A Continuous Principle for Teaching 223
Result. If the fraction r\/rz is irrational, then the curve of P will never close. If
ri/r2 is rational, so that r\/ri = m/n (m and n relatively prime), then the curve will
close after exactly n rounds of K\, or equivalently, after m arches.
To better appreciate the full flavors of rationality, we shall consider a few
more problems.
Example 3.6
Most students will know the following proposition: The square root of a
natural number is either again a natural number or an irrational number. The same
holds true for the kth root instead of the square root (to be proved analogously).
Indirect proof: Let a = m/n (a £ N) and b = k/l (b <£ N) then the above condition states
^*1 —
'm'
— v y = c (c e N).
n
\ 7
This yields:
6 m^
k = cl or, equivalently, — = Nc1
«
V 7
As c1 e N, then, according to the above proposition, k^lcr is either a natural number
itself (contradiction to m/n = a g N) or an irrational number (contradiction to
m/n = HJ). That is why such numbers cannot exist.
Example 3.7
and have to answer the question posed with Yes; if not, we have to continue
working. So let us assume that V2^is not rational and continue trying. We form,
for instance.
(VT5)^ = V22 = 2
and see that the result, 2, is not only rational but also a natural number, though
the exponent (V2) is irrational and (according to the assumption) the base (Vz^)
is irrational, too.
Example 3.8
ALG
Are there a, b e C, but a, b e R, with ab = c e R?
Students generally learn Euler's formula e^ = cos (j) + i sin <(> without an exact
13^ foundation by expansion into series. But the exact derivation and the knowledge
ZJ of all problems around the exact definition of the complex exponential function
are not really necessary for the following considerations.
It is quite simple to insert some values into Euler's formula and just wait with
excitement for the result. For <() — k one can already get the first interesting result:
em = cos n + i sin 7t = -1.
The numbers e and k are irrational (even transcendental), i x k is a complex
number, and em is nevertheless an integer—really surprising! For <j> = 2n one gets
e2m = 1—a natural number.
A Continuous Principle for Teaching 225
.K
Now let us take <|) = k/2 and substitute it; we get e 2 = i—also an interesting
identity. The simplest power where base and exponent are not real numbers is i1.
We get
.2 7t
1
a real number, which is expressed again by the numbers e and n, which are
extremely important for mathematics in general.
With help of the Euler's formula we can also get an insight to the periodicity
of the complex exponential function and can find as well solutions to problems
like the following.
Example 3.9
As seen already, Euler's formula yields, for <}> = 2n, the length of the identity
e2ni = 1. Since the functions sin and cos are periodical with a period 2n, the formula VII/
also holds true for <|) = 47t, 67c,..., 2kn,...:
We see the equation ex =1 has not only the solution x = 0 but even an
infinitude of solutions x = 2kni (k e Z). Now to the equation of the posed problem
1* = 3, which can be described—as we have just seen—in the following way, too:
ln3 ln3
x i(ke Z).
X 2kni 2kn
The given equation does not have a solution in R, but in C it even has infinite
solutions, so this is an especially interesting proof of the "mathematical abun¬
dance" of complex numbers. Of course, one cannot carry on complex analysis in
school, but a little insight into the possibilities of mathematical extension by
complex numbers seems to us very interesting for students and beneficial to their
motivation, especially in advanced courses.
Analytic geometry also offers some surprises and good opportunities to
incorporate complex numbers into certain considerations.
226 THE ART OF PROBLEM SOLVING
Example 3.10
Given any circle and a point P outside of it, we easily can find the
ALG equation of the tangent through P. But what if P lies inside the circle.
Is there still a tangent? Formally, the point Pcan be inserted into the
tangent-equation. But is there any meaning in doing so?
It surely needs no explanation that this is not possible in the range of real
numbers. But in the realm of complex numbers, even this is possible! We want to
try, for instance, the point with coordinates (0, 0), the center-point of the circle
2 2 2 ^
x + y = r , to put formally/arithmetically the tangents on that circle. From ana¬
lytical geometry we know that the line y = kx + d will touch the circle x2 + y2 = r2 if
and only if r2(l + k2) = d2.
As the tangent shall be put on the circle from the point (0, 0), d - 0, and
because of r * 0 it follows: 1 + k2 = 0, so k = ± i.
Therefore, the equation of the tangents is y = ± ix. As expected, these are
imaginary lines, purely complex ones. These straight lines have some other sur¬
prising properties as well.
It is well known that two straight lines are orthogonal if for their slopes k\
and ki, the following is valid: k\ k2 = -1. For the straight line y = ix, k\ = z; what is
wanted now is a kz with ikz = -1 (i.e., k2 = i). Therefore, the straight line that is
orthogonal to y = ix is y = ix itself (!)—and this is not a misprint; in the range of
complex numbers, there are straight lines that are orthogonal to themselves
(analogously: y = -ix).
Take any pair of points P, and P2 of the line y= ix (or y = -ix, respec¬
tively). What distance will they have?
P\(xi, yp and Pz(x2, yz) will be two points on the line y = ix. The length
(Euclidean distance) of segment PiP2 always equals zero:
(P1P2)2 = (X2 - Xi)2 + (ix2 - ix 1)2 = (X2 - Xi)2 + i2(xi - x\)2 = 0. (Since z'2 = -1)
A Continuous Principle for Teaching 227
Example 3.11
Calculate the sum of all natural numbers up to 300 that are divisible by
neither 8 nor 6.
This is a typical task that can easily be divided into subtasks. First one could
calculate the sum of all natural numbers from 1 to 300:
Now the sum of those numbers that are divisible by 8 or 6 is to be subtracted. The
sum of all numbers (< 300) that are divisible by 8 is
37 x 38
8 + 16 + 24 + ... + 296 = 8(1 + 2 + 3 + ... + 37) = 8 x - = 5,624.
The sum of all numbers (< 300) that are divisible by 6 is:
50x51 _,rn
6 + 12 + 18 + 24 + ... + 294 + 300 = 6 x —= 7,650.
228 THE ART OF PROBLEM SOLVING
If one now forms the sum 5,624 + 7,650 = 13,274, then one gets the sum of all
numbers that are divisible either by 8 or by 6—or do we? No, there has to be a little
correction: The number 24, for instance, was counted twice (see above), because it
is divisible both by 8 and 6. We see that all numbers with this quality were counted
twice, which means we have to subtract them once. Which numbers are these that
are divisible by 6 and 8? It is not particularly difficult to see that they are the
multiples of 24.
Therefore the sum
has to be subtracted from 13,274: 13,274 - 1,872 = 11,402. Now this number has to
be subtracted from 45,150, and as the result we get 33,748.
Example 3.12
Beginning with 1, the natural numbers are added until one gets a
three-digit number with three equal digits. How many numbers have to
be added (cf. Baron & Windischbacher, 1990, pp. 14, 89f)?
Solution. One solution would be to simply add up step by step until one gets the
demanded form. But we want to choose another method. We have to add n
numbers, and this sum
n(n +1)
£■' = 2
i= 1
©may not exceed 999. By inserting, or by solving a quadratic equation, one can see
easily that n may not exceed 44. We look, therefore, for an n < 44, so that
n(n +1)
---= xxx (1 < x < 9)
Now, unfortunately, the expression "xxx" is not very usable; therefore, we are
looking for a more suitable and simpler expression of "xxx," such as xxx = x x 111,
and therefore
n(n +1)
= x x 111.
2
But the number 111 can be written even more basically as a product of primes
37 x 3 = 111; then we get
Example 3.13
9x10
300(1 + 2 + ... + 9) = 300 x 13,500.
2
230 THE ART OF PROBLEM SOLVING
The following example from Kranzer (1989, pp. 157, 223ff) should not be
interpreted as a practice-oriented example. The wording here is only one aspect
that should support the surprise from the result of this thought experiment and
may contribute to motivation.
Example 3.14
Solution. Let us first consider how the length of the band changes in the course of
the time. At each full second the band is homogeneously stretched by 1 kilometer.
Let the time immediately before the completion of the nth second be n, and the
one immediately after the completion of the nth second be n+; between both the
band has been stretched by 1 km. At the time n~, the band is n km long, but at the
time n+, its length is n + 1 km, caused by the consistent homogeneous stretching
of the whole band by the factor (n + 1 )/n "as quick as lightning." Also, the distance
already covered by the snail is being increased.
Let us look first at the individual 1 cm pieces the snail covers during the 1st,
2 , • • •, n second, in particular at the stretches that have an effect on these pieces.
The 1 cm piece of the first second is at the transition 1“ —» 1+ being stretched
to its double length, at the transition 2“ -4 2+ to the 3/2-fold length,..., in general
at the transition n -> n+ to the (n + l)/n-fold. So at the time n+ it has the length
(in cm):
,234 n +1 n+1
1XTX2X3X"'x n 1
The 1 cm piece along which the snail creeps during the second second is in the
transition 1 -4 1 + not yet stretched as a covered distance; it is being stretched only
at the transition 2 -4 2+ to the 3/2-fold length, at the transition 3“ -4 3+ to the
4/3-fold length,..., in general at the transition n -4 n+ to the (n + 1)/n-fold length.
It therefore has at the time n+ the length (in cm):
i 3 4 n+1 n+1
lx-x5x . X-
n 2
A Continuous Principle for Teaching 231
The 1 cm piece along which the snail creeps during the third second is being
stretched only by the third stretching onward, for the length of this piece at the
time n+ we get analogously (in cm):
„ 45 n+ln+1
1 X - X - X ... X-= —I—.
3 4 n 3
Finally we look at the 1 cm piece that is covered in the nth second, which is
stretched exclusively at the transition n~ —» n+ with the stretching factor (n + l)/n.
So this piece has at the time n+ a length of (n + 1 )/n cm.
The sum of all these stretched 1 cm pieces indicates the total distance of the
snail from the starting point A. This distance is therefore at the time n+:
(n +1) x
1 v
3 n
The negative fractions in the second parentheses are exactly the terms of the
harmonic series, from which one knows that the series is divergent that is, it
grows beyond all limits. The numerical value of the series
n
1=1
is for a very big n therefore also bigger than 10^' and the value of the expression in
the second parentheses is therefore negative; but if the distance that remains to be
covered is no longer positive, then this means that the end has been reached
already. It takes, of course, a very long time for this to happen. To get at least a
very rough approximation of the order of magnitude of the necessary period of
time, a formula for the approximation for a very large n could help. The exact
foundation of the following limit-statement (Euler-Mascheroni's constant):
will hardly be possible at school, but the teacher could just state it, without proof,
to give an idea of the necessary period. For large n, we therefore have:
This is the time in seconds the walking-tour takes the snail. But even expressed
in years the number does not look much smaller—it would be approximately
1.6 x 1043/422 years, an unimaginably huge number and an even more unimaginable
period of time (according to today's knowledge, about 2 x 1010 years have passed
since the Big Bang)!
References
STEPHEN E. MORESH
Part One
The field of combinatorics has always been very interesting to both the
mathematician and the lay person. Problems in this area are often related to the
real world and are usually a great deal of fun. Unfortunately, students studying
this discipline, whether at the secondary or university level, often complain that
the problems are very difficult. When asked why they find it so hard, the usual
response is that the type of reasoning involved in combinatorics is different from
that in other mathematics courses they have taken. In fact, some students, trying
to classify problems by "types," often give up in distress, after ending up with
scores of different types of problems.
If you ended up with so many types of problems, you might want to give up,
too; however, the analysis of most problems in combinatorics can be narrowed
down to two basic categories. This chapter will focus on combinatoric problems a
little differently, presenting two processes that can be used to solve most problems.
The only assumption being made here is that the reader has had some experience
233
234 THE ART OF PROBLEM SOLVING
(limited as it may be) with permutations, combinations, calculating nPr, nCr, and
doing simple probability problems.
Combinatorics is usually introduced through a simple problems such as the
following:
In how many ways can a man choose a shirt and a tie if he has three
shirts and four ties?
Figure 11.1
It is clear that since this "tree" has 12 different paths, each one representing
a different outfit, the answer to the problem is 12. And so we have Combinatorics
Problem-Solving Technique 1: Draw a tree diagram.
Another View of Combinatorics 235
Traditional Approach. Since Mr. Appel is selecting toppings, order does not count.
Therefore, we can consider this to be a problem involving a combination, a
©
COMB
selection of objects without regard to order.
Mr. Appel has 6 toppings from which he can choose 0, 1, 2,. .., 6 toppings
for his pizza. Thus, there is a total of 1 + 6 + 15 + 20 + 15 + 6 + 1 = 64 different
pizzas, so the answer to the original problem is 64 days. Most people are amazed
at this, since they intuitively expect the answer to be at least several hundred. (Note
that the symbol, in some texts written as nCr or C(n, r), or n is used to mean the
r
number of ways r items can be selected from n items.)
AND process approach. Although "and" does not even appear in the original
problem, an AND process is implied. If we can find a way of rethinking the
problem to find such an AND process, it will greatly simplify the solution.
Picture Mr. Appel sitting in the restaurant, deciding on the type of pizza he
will have today. He starts by looking at the different toppings and asks himself.
Should I have pepper? AND Should I have onions? AND Should I have sausages?
AND Should I have mushrooms? AND Should I have broccoli? AND Should I have
anchovies?
For each question, he has to decide whether or not to include the specific
topping. For example, for the first question (Should I have pepper?), the answer is
Yes or No. This gives 2 possibilities for the first stage of the process. After deciding
GEN about pepper, he goes on to the next stage (Should I have onions?), and after
deciding about onions, he decides about sausages, then mushrooms, and so forth.
This can be summarized as a six-stage AND process. Since there are 2 choices at
each stage, there are 26 = 64 possible pizzas. Note that this includes the plain cheese
pizza, in which the decision at each stage was not to take the given topping. A
major advantage of this approach is that it can greatly cut down the amount of
calculations. If the problem had originally included 15 different toppings for the
pizza, the traditional approach would be quite messy indeed. This approach
would yield 215 = 32,768 different toppings immediately. An AND process could
be thought of as any given situation in which there is a sequence of stages in order
to accomplish what is being described. This problem is equivalent to finding the
number of subsets there are for a given set.
In many places, the custom is to give a “tip” for good service, whether
COMB in a restaurant, where a waiter has taken your order, at the airport,
where a porter has carried your suitcase, or in a building, where the
doorman helped you carry some packages. In the United States and
other countries, the tip is not an automatic part of a bill but given at the
Another View of Combinatorics 23 7
discretion of the person being served. Ms. Smith decided to give her
doorman a little tip for helping her with a package. If her purse has only
a quarter, a half dollar, a silver dollar, and a five-dollar bill, how many
tips are possible?
Here the money denominations replace the pizza toppings. Using the ap¬
proach of the pizza problem, Ms. Smith looks at her quarter and decides whether
or not to give it to the doorman, giving two possibilities, AND then does the same,
one by one, for each coin and bill. Thus, there would appear to be 24 = 16 possible
tips. This includes the case of her being very generous and giving everything to the
doorman; however, it also includes the extreme opposite case of giving nothing at
all. Since the problem states that she definitely gives a tip, we must exclude that
case from the 16, giving an answer of 15 possible tips.
Note that we had to look back to see if any special or extreme cases had to be
handled. This is very important, since it may be implied in a problem. Sometimes
it is not clear at all, and two answers must therefore be given to take into account
the possibility of special cases. We state this as Combinatorics Problem-Solving
Technique 3: Look back to check for special cases.
Why is this more difficult than the previous problem? We have to be extra
careful not to overcount the number of possibilities. For example, a 6-cent tip is
just one possibility, but it can be achieved in four ways. Use the nickel and the first
penny, the nickel and the second penny, and so on. Just shifting our thinking
slightly will enable us to incorporate an AND process into the picture. Some types
of coins have multiple possibilities! Rather than looking at each coin and deter¬
mining whether or not to take it, we can envision the process as looking at the type
of coin, or denomination, and seeing the number of possibilities.
First, decide about the pennies: Take 0,1,2,3, or all 4 of them —> 5 possibilities.
AND once you're finished, decide about the nickel: Take it or leave it —> 2 pos-
sibilities. AND once you're finished, decide about the dime: Take it or leave it
2 possibilities. AND once you're finished, decide about the quarters: Take any¬
where from 0 to 6 quarters —> 7 possibilities.
This would normally give 5x2x2x7=140 possible tips.
We must now go back to see if there are special cases. The extreme cases pp.
should always be checked first. The case in which the person gives the porter every
238 THE ART OF PROBLEM SOLVING
An assistant helped Ms. Streett retrieve her coat after she ate in a fancy
restaurant. When she looked in her purse, all she had was a twenty-
dollar bill, two dimes, two quarters, and a nickel. She decided definitely
to give a tip, but not the twenty-dollar bill. In how many ways may she
give the tip?
There are clearly 17 different ways of tipping, which the problem asked for.
Had the question asked how many different tips are possible, the answer would be
different. Listing the values of the tips (in cents), we get:
Note that 25- and 50-cent tips appear twice. Why does this happen? In the
original problem, combinations of some of the coins add up to other coins.
Specifically, the two dimes and a nickel add up to a quarter. Two dimes and a
nickel are a different way of tipping from leaving one quarter; however, they have
the same total value. In the previous tipping problems, this difficulty did not occur.
In this problem, because of the double occurrence of 25- and 50-cent tips, we must
subtract 2 from the 17 tips listed above, thus giving us 15.
This last example illustrates that we must read all problems very carefully.
Counting problems are infamous for the number of nuances that can occur. And
so we come to Combinatorics Problem-Solving Technique 4: Read the problem very COMB
carefully until you understand exactly what is being asked, and be sure to clear
up any ambiguity before attempting to solve the problem.
If we were going to be a little more realistic, we would add a condition that
a certain minimum tip was being considered. Other variations and approaches to
the last tipping problem can be considered.
Finding the number of committees that can be formed is a popular type of
question for students studying combinations. The following problem illustrates
how AND processes can make a difficult committee problem much easier.
Method 2. This method also uses an AND process. Here, however, we do the
complementary problem; that is, we will make sure that at least 1 husband-and-
wife team are represented. Then subtract this number from the number of com¬
©
mittees possible with no restriction.
Although the first method was easier here, the second method was presented
since it will be the easier approach in other problems. In addition, it gives us
© Combinatorics Problem-Solving Technique 5: Do the complementary problem and
subtract this answer from the total number of outcomes in general.
Exercises
GEN For each problem, (a) interpret the problem using an AND process; (b) solve
the problem; and (c) if possible, find another (more traditional) solution.
COMB 1. A certain locksmith makes keys using any one of 15 brands of keys. Each
brand uses a different length and thickness. No matter what brand is
chosen, there are 8 different positions from which metal can be removed
or left alone. For the first 5 positions, there are 3 different depths at which
the metal can be removed. For the remaining positions, there are 4
different depths. How many different keys are possible if metal must be
removed in at least one position?
COMB
2. How many subsets does a set of 15 elements have?
COMB 3. Many different kinds of desserts have been put out for people to choose
from at a party. Judy is watching her weight and decides to have at least
one piece of fruit and no other type of dessert. If the fruit bowl contains
exactly 6 apples, 5 plums, 3 oranges, 4 bananas, and 5 pears, how many
different combinations of desserts can Judy have?
COMB
4. In how many ways can Peggy make a selection of one or more new fish
for her tank from 6 identical coral fish, 7 identical angel fish, and 3 identical
blue pullers?
COMB
5. A postal carrier is late and wants to hurry up to go home. He decides to
just put any letter into any mailbox. If there are 12 letters and 20
mailboxes, in how many ways can the letters be placed if (a) each letter
is to go in a different mailbox; (b) any number of letters may be placed in
a mailbox?
COMB
6. In how many ways can the letters in exercise 5 be placed if there are 20
letters and 12 mailboxes?
COMB
7. Problems 5 and 6 above can be generalized. In how many ways can a
set of m items be matched with a set of n items if m < n and (a) each of
the m items can be matched with at most one of the n items; (b) there is
no restriction on the number of items with which any items can be
matched?
COMB 8. There are 5 different algebra books, 4 different biology books, and 6
different calculus books. How many selections of books can Roslyn make
if (a) she decides to take at least one book; (b) she decides she must
include a specific algebra book; (c) she decides she must include an
Another View of Combinatorics 241
algebra book but it could be any one of them; (d) she decides she must
take at least one algebra and at least one biology book; (e) she must have
at least one of each type of book?
9. (a) In how many ways can a triangle be named using letters of the GEOM
alphabet? (b) In how many ways can a triangle be named using letters of
the 24-letter Greek alphabet? (c) In how many ways can a quadrilateral
be named using letters of the alphabet? (d) In how many ways can an
n-gon be named using letters of the alphabet? What restriction is there
on the value of n?
10. After breakfast in a diner, Mrs. Stux looked for change in her purse to give COMB
the waiter a tip. She had 6 dimes, 2 quarters, 4 dimes, and 3 pennies. How
many tips can she give if (a) she decided to definitely give a tip; (b) she
did not like the service so she will give at most 5 coins; (c) she did not
like the service so she will give at most 5 coins but will definitely give a
tip; (d) she did like the service so she will give at least five coins; (e) she
liked the service and intends to give at least 75 cents.
Part Two
How is the following problem different from previous ones that we have
considered?
selection down to 6 books. In how many ways can she make her
selection?
The major difference here is that we have what we will refer to as an "OR
process." She can take 2 OR 3 books from the library. She cannot take exactly 2
AND exactly 3 books at the same time. Thus, there are different cases to consider.
In this problem, the most important word is "or." If you can recognize that
LOGIC
a problem involves an OR process, it may greatly simplify its solution. An OR
process is any situation in which either one thing will be done or another thing
will be done, but not both of them.
In order to solve the vacation problem, analyze each case as if that case were
the entire problem:
©
Case 1: Brenda chooses exactly 2 books. She can do this in
v2;
-15 ways.
©
Brenda chooses exactly 3 books. She can do this in
V = 20 ways.
Case 2:
v3y
242 THE ART OF PROBLEM SOLVING
These two cases cannot occur simultaneously (she will choose either exactly
2 books OR exactly 3 books but not both, since both events cannot happen at the
same time). This leads to the following definition: Two events are said to be
COMB
"mutually exclusive" or "disjoint" events if they cannot possibly happen at the
same time.
The events described in cases 1 and 2 are obviously mutually exclusive
events. Since each of the 15 outcomes in case 1 contains 2 books and each of the 20
outcomes in case 2 contains 3 books, there is no overlap. In other words, the 15
outcomes in case 1 are all different from the 20 outcomes in case 2. Thus, the total
number of selections that Brenda can make is 15 + 20 = 35.
We now have Combinatorics Problem-Solving Technique 6: For any situation
that can be interpreted as an OR process, handle each possible case separately and
then add to find the total number of possibilities.
A hint that an OR process is involved is the presence of mutually exclusive
events. Thus, there is another way to state Combinatorics Problem-Solving Technique
6: If two events are mutually exclusive, and the first event can occur in x ways and
the second can occur in y ways, then the total number of ways that the first event
or the second event can occur is the sum, x + y.
Very often, problems involve an OR process in a way that is not obvious. Let
us focus on some of these.
ARITH How many whole numbers less than 1,000 can be created if the digits
must come from the set {3, 5, 6, 7, 9}?
If the number is to be less than 1,000, there are three mutually exclusive cases
to handle:
(J) Case 1: The number created has one digit. There are 5 possibilities here.
Case 2: The number created has two digits. There are 5 x 5 = 25 possibilities
here. (Note that repetition of digits is permitted, since the original
problem had no prohibition against this.)
Case 3: The number created has three digits. There are 5 x 5 x 5 = 125
possibilities here.
Since these three cases are mutually exclusive, we add the number of ways
in which each case can be accomplished, giving 5 + 25 + 125 = 155 numbers.
A problem that has interested state and federal officials has been that of
creating enough license plates. When a region runs out of license plate numbers,
new rules are often formulated in order to create enough new numbers to allow
for the growing number of vehicles.
Another View of Combinatorics 243
A certain state has license plates which consist of three letters of the
alphabet followed by a three-digit number greater than or equal to 100.
It also produces, for a fee, “vanity” plates, such as “DAVID,” “MARY
1,” or “NURSES.” The Bureau of Motor Vehicles, in order to provide for
more cars, trucks, and buses, produced new license plates consisting
of three letters followed by a four-digit number greater than or equal to
1,000. How many nonspecial licenses can this state produce?
In the above problem, there is an OR process implied, since there are three
cases—special licenses, those that consist of six characters using the old rules, and
those with seven characters using the new rules. Since we are not interested in the
special licenses, we need only focus on the last two cases.
Thus, the total number of licenses possible is 174,002,400. Had the state used
the last rule in the first place, it probably would not have needed a change, since
this rule alone provides for many more vehicles than it would ever have. In
actuality, when states figure out how many licenses are possible, there are many
special cases that must be excluded from consideration. For example, in some
states the following three-letter combinations are not allowed for obvious reasons:
KKK, LSD, SEX, JEW, and so forth.
License plate problems are very typical of many OR process problems in that
each case involves an AND process. Each case is therefore handled using multi¬
plication, whereas the final result is handled by adding the amounts given by the COMB
various cases. We state this as Combinatorics Problem-Solving Technique 7: Each case LOGIC
of an OR process can often be interpreted as an AND process.
244 THE ART OF PROBLEM SOLVING
Exercises
ARITH
1. How many whole numbers less than 1,000 can be created if the digits
NUMB
must come from a set of 8 different non-zero one-digit numbers?
ARITH 2. How many whole numbers less than 10,000 can be created if the digits
must come from a set of n different non-zero one-digit numbers?
COMB 3. How many license plates would a state have using the rule that the license
have (a) four letters followed by a two-digit number; (b) four letters
followed by a two-digit number or a two-digit number followed by four
letters; (c) six characters which can be digits or letters?
COMB 4. How many licenses are possible if a state allows licenses to contain three
letters followed by three digits or three digits followed by three letters?
COMB
5. A state creates license plate numbers by using as a general rule three let¬
ters followed by three digits. Every possible license number has been used!
Since it is expensive to create new licenses for everybody, it is probably
cheaper to add a new character for new licenses. Explain how to generate
new licenses so that the maximum number of plates is available.
COMB 6. How many different selections of at least one book can Erica make from
7 different books? (a) Do this problem using an AND process. (b) Do this
problem using an OR process. For this problem, which process produces
an easier solution?
COMB
7. A committee of at least 2 people is to be made from 8 people. How many
different committees are possible?
COMB
8. Radio stations in the United States can have three or four call letters. East
of the Mississippi River, they start with the letter W, while west of it they
start with the letter K. How many different radio stations are possible using
these rules?
COMB 9. Tony’s Tattoo Parlor offers the following designs: a heart with a dagger
through it, an eagle, a rose, a panther, a skull, or a design with MOM,
DAD, your name, or your girlfriend’s name on it. No one gets the same
tattoo twice, (a) Yesterday, Bob was seen going into Tony’s. No one
knows how many tattoos he got. He may have even lost his nerve and
gotten none! How many different selections of tattoos could Bob now
have? (b) Suppose you heard that Bob definitely got a design with “MOM”
put on his right arm. How many different selections of tattoos could Bob
now have? (c) Suppose you heard that Bob definitely got five tattoos
including one with his girlfriend’s name on it. How many selections of
tattoos can Bob now have?
COMB
10. A certain diagram has 20 points in a plane, no 3 of which are collinear.
How many different (a) triangles can be connected? (b) quadrilaterals can
be connected? (c) hexagons be connected? (d) polygons can be con¬
nected? (e) Redo part d another way.
Another View of Combinatorics 245
Part Three
An urn contains three red marbles and one blue marble. If two marbles
are chosen randomly, find the probability they are both red if (a) the PROB
first marble is returned to the urn before the second one is chosen; (b)
the first marble is not returned to the urn before the second one is
chosen.
Both parts of this problem involve AND processes. The problem turns out to
be more difficult than it appears! Two reds can occur in many ways. Both parts of
the problem are very important, since replacement problems illustrate an impor¬
tant problem-solving technique related to AND processes.
What are the different problem-solving strategies used to solve the two parts
of this problem?
First, to simplify the problem, we envision each of the red marbles having
the number 1, 2, or 3 on it. We will let R1 stand for the first red marble, R2 the ©
second, R3 the third, and B the blue marble. Thus, the sample space is {RjR^ RXR2,
R1R3, RxB, R2Rj, R2R2, R2R3, R2B, R3Rv R3R2, R3R3, R3B, BRv BR2, BR3i BB). Of the
16 possibilities, 9 have red in both cases (R^R2, for example). Thus, the desired
probability is 9/16. Note the following probabilities associated with this problem:
P(red the first time) = 3/4, P(red the second time) = 3/4. P(red both times) turns
out to be the product of the two probabilities.
In the second strategy, the sample space is different, since certain cases such
as R3R\, for example, cannot be included. What happened in the first event affects
what happens the second time. The sample space is therefore {RiR2, R3R3, R\B,
©
R2Ri, R2R3, R2B, R3Rv R3R2, R3B, BRv BR2, BR3). Of the 12 possibilities, 6 have red
in both cases. Thus, the desired probability is 6/12. Here the associated prob¬
abilities are: P(red the first time) = 3/4. P(red the second time) = 9/12 = 3/4. P(red
both times) does not turn out to be the product of the 2 probabilities.
What is different about these cases that allows us to multiply probabilities
the first time but not the second time? The difference is that in the first case, the
first outcome has no effect on the second outcome. In the second case, what
happens the first time does affect the second. Here, for example, blue for the first
selection disallows blue in the second selection.
246 THE ART OF PROBLEM SOLVING
Maile randomly pulls a card from a standard deck of 52. (a) What is the
probability that she pulls a seven or an ace? (b) What is the probability
that she pulls a seven or a red card?
Exercises
For each problem, see if AND or OR processes can be used to find the given
probability. Then find the desired probability.
PROB
1. A die is rolled 5 times. Find the probability that the number 2 shows up
all 5 times.
PROB
2. A nickel and a dime are tossed simultaneously. Find the probability that
(a) both show heads; (b) both show tails; (c) the nickel lands tails and the
dime lands tails; (d) one of them lands heads and the other lands tails.
PROB
3. Helen tosses a coin 6 times. Find the probability that she gets heads the
first 3 times and tails the last 3 times.
PROB
4. The numbers 7, 8, 11, 12, and 15 are written on 5 slips of paper and put
into a hat and stirred. Two numbers are picked, one after the other,
without replacement. The two numbers are then noted. Find the prob¬
ability that the sum of the numbers is odd.
PROB
5. A box contains 7 marbles: 3 red and 4 blue. Two marbles are drawn, one
after another. Find the probability that both are red if (a) the first marble
is replaced before choosing the second one; (b) the first marble is not
replaced.
PROB
6. Roz bought 12 cups of yogurt. Unfortunately, 4 of them are spoiled. Find
the probability that the first 3 cups she chooses turn out to be spoiled.
PROB
7. Blake is a senior in high school. He estimates the probability of being
accepted at college A to be 0.75 and the probability of being accepted at
college B to be 0.5. He then multiplies, and claims that the probability of
being accepted at both colleges is 0.375. Explain why Blake’s argument
is or is not correct.
248 THE ART OF PROBLEM SOLVING
Part Four
More Examples
1. Glenda rolls a die 5 times, with “getting a number less than three” being
considered a success. Find the probability that Glenda has 2 successes
followed by a failure, a success, and a failure. What are the probabilities of
success and failure on one throw of the die? Note that p = 2/6 = 1/3 and
q = 2/3. Since the results of the five rolls of the die are independent of one an¬
other, we can treat this problem using an AND process. Using combinato¬
rics problem-solving technique 8, we can multiply to get the desired prob¬
ability. This gives P(SSFSF) = p x p x q x p x q = pV = (1 /3)3 x (2/3)2 = 4/243.
2. Josh, like Glenda, rolls a die 5 times, with “getting a number less than
three” being considered a success. What is the probability that Josh has
2 failures followed by 3 successes? P(FFSSS) = cfp3 = p3cf = (1/3)3 x
(2/3)2 = 4/243, the same as in the example above.
3. Adam, like Glenda and Josh before him, rolls a die 5 times, with “getting
a number less than three” being considered a success. What is the
probability that Adam has alternating successes and failures, starting with
a success? P(SFSFS) = px qx px qx p = p3cf = (1/3)3 x (2/3)2 = 4/243,
the same as before!
What do these three examples have in common? Of course, they are examples
of Bernoulli trials with the same p's and of s. An AND process was implied in each
example. In addition, each person had 5 trials, with 3 successes and 2 failures. To
generalize these results, we ask the following:
250 THE ART OF PROBLEM SOLVING
In rolling a die 5 times with “getting a number less than three” being
PROB
considered a success, what is the probability of getting exactly 3
successes?
Table 11.1
0 FFFFF (1 way) i 5 i 0 5
M = lp 9 =
0
V /
5 SSSSS (1 way) 5 0
ipV= 51 P <7
5
Another View of Combinatorics 251
The above table will enable us to answer slightly more complicated prob¬
lems. Try the next example before reading the solution.
In rolling a die 5 times with getting a number less than three being
considered a success, what is the probability of getting exactly 3 OR
4 successes?
The events getting exactly 3 successes and getting exactly 4 successes are
mutually exclusive. As a result, we can use combinatorics problem-solving tech¬
nique 9 and add to get the desired probability, giving 10p3q2 + 5pV = 40/243 +
20/243 = 60/243 = 20/81.
Study the table above. Where have you seen the entries in the last column
before? Recall:
Deanna did not study for her test because she thought she could pass
it without studying; to use the vernacular, she thought she could “wing
it.” If the test consists of 10 multiple-choice questions, each with 4
choices, and 70% is passing, what is the probability that she passes
the test by just guessing?
Here, careful thought must take place before realizing that, although there
are 4 choices each time, each outcome is either a success (getting the right answer)
or a failure (getting the wrong answer). Her getting a question right or wrong is
independent of what she did on the previous questions. Since she is just guessing,
the probability of getting a correct answer to any question is p = 1/4 each time,
and the probability of getting the wrong answer is q = 3/4. The only difficulty here
is that we have to consider 4 mutually exclusive cases. After dealing with these
cases, we then use combinatorics problem-solving technique 9 and add the 4
results.
''l0'N
Case 2: P(exactly 8 successes) = psq2 = 45(l/4)8 (3/4)2
8
v y
= 45(1/65,536) (9/16) = 405/1,048,576.
Thus, the probability that Deanna passes her test is the sum, or 3,676/1,048,576,
which is approximately 0.0034. Our only advice to poor Deanna is to study, study,
study!
The beauty of this last problem is that it illustrates the importance of what
we did in parts 1 and 2 of this chapter. Each case itself involves an AND process,
while all the cases considered together give us an OR process. By viewing the
problem this way, it really is not too difficult, except for doing some messy
calculations.
We end this chapter with a problem that comes up every October in the
United States and Canada.
Before solving this problem, there are several considerations. Even though
the World Series is supposedly a 7- game series, it may end in fewer games. If a
team wins the first 4 games, or 4 out of the first 5 or 6 games, there is no need to
continue on to a 7-game series. As a result, we have to be very careful. First, let us
focus on a few related questions.
First, is the World Series an example of a Bernoulli trial?
1. There are two outcomes for each “trial”: success if your favorite team PROB
3. We will accept the fact that the probabilities do not change, even though
this may not be true as explained in (2), above. In fact, with slightly
different lineups of players, the probabilities likely do change, if only
slightly. Again, we may very well oversimplify matters by assuming that
we have Bernoulli trials.
PROB Second, what are the binomial probability parameters in this problem? Since
the problem says the teams are equally matched, the probability of each team
winning (and losing) is 1/2. Since the problem asks us to find the probability that
the New York Mets win, we accept the Mets winning as success and the Blue Jays
winning (the Mets losing) as failure. Thus p = q= 1/2.
The number of successes is obviously 4, since it takes that number of games
to win. The biggest difficulty is the number of games in all. A World Series can be
played in 4, 5, 6, or 7 games. If we were to make these separate cases, we run into
the danger of counting several combinations more than once. For example, the
winning combination SSSS is included in each of the other cases! FFFF is embed¬
ded in SSSSF, SSSSFF, SSSSFFF. We can simplify the work by realizing that the
winner must win the very last game, and the only flexibility is in choosing the
successes and failures among the games before the last one.
Method 1
Case 2. 5 games: Since the Mets will definitely win the last game, we must
find P(winning 3 of the first 4 games and winning the last game) =
© p\p = 4(1/32) = 1/8.
v3y
Case 3. 6 games: We must find P(winning 3 of the first 5 games and winning
the last game) = ©
pVp = 10(1/64) = 5/32.
3
v y
Case 4. 7 games: We must find P(winning 3 of the first 6 games and winning
the last game) =
©
pVp = 20(1/128) = 5/32.
3
v y
Thus, the probability that the Mets win is the sum 1/16 + 1/8 + 5/32 + 5/32 = 1/2.
(Is this the answer you would expect?)
Method 2
It would have been easier had there always been 7 games. We can envision
for the sake of this problem, that the teams do play 7 games. Even if the Mets win
Another View of Combinatorics 255
the first 4 games, they could have continued giving such combinations as SSSSFFF,
SSSSSFS, SSSSFSS, and so on. Thus, we can do the problem by finding the
probability that in 7 games the Mets win at least 4 games: This gives P(Mets win
Exercises
bilities. In other words, if two events are mutually exclusive, and the first
event can occur in x ways and the second can occur in y ways, then the
total number of ways that the first event or the second event can occur is
their sum, x + y.
7. Each case of an OR process can often be interpreted as an AND process.
8. For a problem involving an AND process and two independent events A
and B, P(AaB) = P(A) x P(B).
9. For a problem involving an OR process and two mutually exclusive events
A and B, P(AvB) = P(A) + P(B).
10. If a problem involves n repeated Bernoulli trials, use binomial prob¬
abilities.
11. Use a calculator to find most binomial probabilities.
References
Blakeslee, D., & Chinn, W. (1971). Introductory statistics and probability: A basis for
decision making. Boston: Houghton Mifflin.
Coppins, R., & Umberger, P. (1986). Applied finite mathematics. Reading, MA:
Addison-Wesley.
Elgarten, G., Posamentier, A., & Moresh, S. (1986). Using computers in mathematics.
Menlo Park, CA: Addison-Wesley.
Garfunkel, S. (1988). For all practical purposes: Introduction to contemporary mathemat¬
ics. New York: W. H. Freeman.
Goldberg, S. (1960). Probability: An introduction. Englewood Cliffs NJ- Prentice-
Hall.
Kemeny, J., Snell, J., & Thompson, G. (1966). Introduction to finite mathematics (2nd
ed). Englewood Cliffs, NJ: Prentice-Hall.
Kenny, M., & Hirsch, C. (1991). Discrete mathematics across the curriculum, K-12:1991
yearbook. Reston, VA: National Council of Teachers of Mathematics.
Mosteller, F., Rourke, R„ & Thomas, G., Jr. (1961). Probability: A first course.
Reading, MA: Addison-Wesley.
Niven, I. (1965). Mathematics of choice, or how to count without counting. Washington,
DC: New Mathematics Library, Mathematical Association of America.
Rabinowitz, S. (1992). Index to mathematical problems 1980-1984. Westford, MA:
MathPro Press.
Soifer, A. (1987). Mathematics as problem solving. Colorado Springs: Center for Excel¬
lence in Mathematical Education.
Szekely, G. (1986). Paradoxes in probability theory and mathematical statistics. Buda¬
pest: Akademiai Kiado.
12
Problem Solving
by the Use of Functions
WOLFGANG SCHULZ
257
258 THE ART OF PROBLEM SOLVING
Problem 1
This matter can be modeled with the help of a function. The burning time t
can be related to the length of the candle f(t). Assuming that the candle will always
become 2 cm shorter each 20 minutes, we can presume a linear function. This
idealizing is very important here; it now makes a very elementary solution
possible.
The question for the burning time of the candle is answered if the zero of the
function is determined. This problem can be solved through different methods.
Making a Picture
If we plot the points Px (0,15) and P2 (60, 9) into a system of coordinates and
draw a straight line through these points, we can read the zero. Hence, it follows
that/(150) = 0. So the candle is burning altogether 2.5 hours.
The question of how long the candle will remain burning can be answered
graphically as well. Candle length 13 cm corresponds to the burning time 20
minutes (see point P3). Hence the candle will remain burning another 130 minutes.
It is remarkable in this example that it is not necessary to know an equation
for the linear function modeling the matter.
Problem Solving by the Use of Functions 259
Making a Table
The concept of functions can also help without a graphic representation. The
problem can be solved with a table of values. One considers the starting length of
the candle (15 cm) and the regular shortening (2 cm every 20 minutes).
Table 12.1
t = 150
The candle is burning altogether for 2.5 hours. How long will the candle
remain burning? We only have to subtract the elapsed time (20 minutes) from the
total burning time (150 minutes).
260 THE ART OF PROBLEM SOLVING
All of these methods are mathematically equivalent and prove that there can
be more than one promising way to solve a problem. It is an important aim of
mathematics classes to make the students aware of this frequently found variety
of possibilities.
Exercises
Figure 12.2
Figure 12.3
Problem Solving by the Use of Functions 261
Problem 2
The distance of the convoy from the barracks at the time t (in hours) can be
described with the help of the function/. The distance (in km) of the convoy from
the barracks at the time t will be stated by f(t). The distance of the motorcyclist
from the barracks at the time t (relative to the departure time of the convoy) can
be described with the help of the function g. So g(t) will state the distance of the
motorcycle from the barracks at the time t. The problem is solved when it is known
at which time t both functions have the same values. If the given speeds are
constant speeds—which is sensible if one interprets the speeds as average
speeds—the functions of/and g are linear. So, again, the situation is simplified by
idealizing. Problem 2 can be solved using the same variants of the solution used
in problem 1.
Making a Picture
The graphs of the functions of/and g can be plotted with two pairs of values
each. The pairs of values result from the given speeds and the starting times. The
motorcycle overtakes the convoy if the graphs of the functions / and g intersect
each other.
262 THE ART OF PROBLEM SOLVING
Figure 12.4
Result: The commander can be given his orders around 2:00 p.m.
The above graph makes one general shortcoming of graphic solutions clear:
The exactness of the result is not known. Furthermore, one cannot read the point
of intersection well if one chooses a clumsy scale, as is the case here. (It is valuable
for mathematics teachers to show the limitations of a method of solution, as is the
case here.) One should not get too impressed by the often only apparent exactness
of calculated results either. In our example, it can be traced back to the rough
simplification of the constant speeds.
Making a Table
This problem also can be solved with the help of a table of values. For this
purpose, the distances of the convoy and of the motorcycle from the barracks are
determined at different times—for example (and for convenience), at every full
hour. One has to consider that the motorcycle is traveling 25 km in half an hour
(traveling at a constant speed).
Table 12.2
The result can be read immediately in the last column. If this is not the case,
one can try calculating distances for shorter periods of time than one hour. Ac¬
cording to this calculation, the meeting time is exactly at 2:00 p.m. But because of
Problem Solving by the Use of Functions 263
the idealization that was made, it is sensible here as well to say that the convoy
and the motorcyclist meet at about 2:00.
Those for whom the graphic solution is not exact enough and the method of
trial and error is too clumsy can find an equation for both functions and then solve
the system of equations.
For the convoy,/(f) = 45f (f in hours,/(f) in km) applies.
For the motorcyclist, g(t) = 50(f - 0.5) applies (in consideration of the later
start time).
The meeting point is reached when both parties have covered the same
distance:
m=g(t)
4 5f = 50 (f - 0.5)
f=5
Result: The motorcycle overtakes the convoy 5 hours after the departure of
the convoy, that is, at 2:00 p.m.
The exactness of this result should, again, not be overrated. The result is only
true for the supposed constant speeds.
Exercises
4. An express train leaves Mountain Valley for Beach Town at 8:00 a.m.
It goes the distance of 300 km at a speed of 120 km/h. At 8:45 a.m., a
freight train leaves Beach Town for Mountain Valley. It goes at a speed
of 80 km/h. When do both trains meet? Do they meet nearer to Mountain
Valley or to Beach Town?
5. Mr. Smith is a sales representative. His earnings consist of a fixed salary
of $1,000 and a sales commission of 5%. He gets a new offer: a fixed
salary of $800 and a sales commission of 7%. For what amount of sales
is the new offer more favorable to Mr. Smith?
The first step in solving a problem is often to show patterns in the relationship
between the quantities involved. Such patterns often occur in functional interde¬
pendencies, provided one has enough data. In the following, two values x and y
will be involved, and k pairs of values will be known. We consider only numerical
values. One gets the following:
264 THE ART OF PROBLEM SOLVING
Table 12.3
X xi *2 *3 xk
y Vi Vi y3 Vk
If this is the case, then y = ax + b, where b = y- - axj for any j between 1 and k.
Proposition 4. The dependence ofyonx is a quadratic function if and only if there are
numbers a and b so that for all i * j holds true: (yf - yf/(xi - xj) = a(Xi + xj) + b.
If this is the case, then y = ax2 + bx + c, where c = y{-axj- bx{ for any i between
1 and k.
The propositions 1 through 4 are so strong that one can use them to decide
if there is such a pattern or not.
Problem 3
X 0 2 3 5 8
y 4 6 13 39 108
We are dealing with a function since there is exactly one value of y for each
value of x. It cannot be a direct proportion, because, according to proposition 1,
the pair (0, 4) must not occur. It cannot be an inverse proportion, because, accord¬
ing to proposition 2, it would have to be true, for instance, that x5 y5 = x2 y2', (i.e.,
8 x 108 * 2 x 6). It cannot be a linear function, because, according to proposition 3,
the quotients (6 - 4)/(2 - 0) = 1 and (13 - 6)/(3 - 2) = 7 would have to be equal; but
they are not. If one continues to form further quotients such as (108 - 13)/ (8 - 3) =
19 and (39 - 13)/ (5 - 3) = 13, one can check with proposition 4 if there is a linear
relation for the following table:
Table 12.5
Xi + Xj 2 5 11 8
Exercises
X -i 3 5 10
y -5 -3 -7 -17
X 0 2 6 7
b
y 3 -1 -9 -10
X i 2 4 5 6 8
Problem 3 could also be solved with the help of a system of equations. Let us
assume that there is a quadratic equation, y = uxf + bx + c, then the following must
be true:
266 THE ART OF PROBLEM SOLVING
yi = axi2 + bxi + c, so 4 = a x 02 + x 0 + c, so c = 4.
yi - axi + bx2 + c, so 6 = a x 22 + b x 2 + c, so 6 = 4a + 2b + 4.
y3 = flX32 + bx3 + c, so 13 = a x 32 + x 3 + c, so 13 = 9a + 3b + 4.
This system of equations has the solutions a-2,b- -3, and c = 4, from which
follows y = 2x2 - 3x + 4. One should check, then, that the other pairs of values fit
into the equation as well. This check is necessary. If one supposed, for instance, in
exercise 6b a linear function, then from the first two pairs it would follow that
a = -2 and b = 3. Hence, the relation could be y = -lx + 3. This relation is also true
for the pair (6, 9) but not for the pair (7, -10).
Exercises
1. There is a direct proportion if and only if the points are all on the same
straight line, which runs through the origin of the coordinates.
2. There is a linear function if and only if the points are all on the same
straight line.
If one confirms in this way that the table of values is based on such a function, then
one can, if necessary, determine the parameters for an equation of functions from
the pairs of values.
Problem Solving by the Use of Functions 267
Problem 4
The tables of values for exercise 7 are shown in the graphs below. Is it
possible to establish certain types of functions from these graphs?
k
y* k y* k •
- • - •
- • •
• •
3 30 3
• •
• 2 20 - • 2
• 1 10 1
^ i i i i T t i i i i k 1 1_1—1— 1 1 1 i—i■>
" -3 -2 -1 1 2 3 wx * -3 -2 -1 1 2 3 \ " -3 -2 -1 1 2 3 X
-1 -10 -1
-2 - -20 - -2 -
-3 -30 • -3
3r 3r 3r
Figure 12.5
One can see in (a) that all points are on the same straight line. Hence, there
is a linear function.
One can see that (b) does not resemble a linear function. It could be a
quadratic function.
One can see that (c) shows none of the functions described in the propositions
1 through 4, because the points are neither on a straight line nor on a curved line
that forms a parabola or hyperbola.
Like exponential and logarithmic functions, power functions have such
strong characteristics that one can recognize them. Those readers who want to deal
more thoroughly with this interesting mathematical problem should refer to J.
Aczel's Lectures on Functional Equations and Their Applications (New York: Aca¬
demic Press, 1966). Apart from functional equations, demands on continuity play
an important part here.
The following propositions renounce continuity and work with strict mono¬
tonicity, which seems more suitable for school, as one can work more easily with
the characteristics of monotonicity than with the mathematically demanding term
of continuity.
268 THE ART OF PROBLEM SOLVING
In the following, R denotes the set of real numbers and R+ the set of positive
real numbers.
If this is the case, the number a results from the functional value/(l) or from
the solution of an equation f(xj) = ax1 for an x1 * 0.
If this is the case, k can be determined by solving the equation f(xf) = x2k
for * 1.
Problem 5
Table 12.8
Time t (days) 0 1 2 3 4 5 6 7
Mass f(t) (mg) 1.0 1.5 2.2 3.4 5.1 7.6 11.4 25.6
From the table of values it follows that the function/is strictly monotonically
growing. One can check quickly that the conditions of propositions 1 through 4
are not met. Nor are the conditions of propositions 5 met, because there is a clear
difference if one calculates, for instance,/(2 x 3) = /(6) = 11.4 and compares this
number with f(2) x/(3) = 2.2 x 3.4 = 7.48. Hence, it cannot be a power function.
If one calculates f(2 x 3) = /(6) = 11.4 and compares this number with
/(2) + /(3) = 2.2 + 3.4 = 5.6, by proposition 7, we can conclude that it is not a
logarithmic function either.
Problem Solving by the Use of Functions 269
Table 12.9
t 0 1 2 3 4 5 6 8
One can say that this function is well suited as an approximation for the
process of growing. This function can now also be used to obtain further knowl¬
edge about the growth process. It follows, for instance, that 1.57 = 17.1, which is an
approximation for the mass of bacteria after 7 days. One should be careful of
long-term predictions. After 20 days, there results a mass of bacteria of 1.5 mg =
3,325 mg. This forecast comes true only if the conditions for the process of growing
do not change—for instance, there has to be enough space and food for the bacteria
to grow.
Exercises
8. The following table shows the number of people on Earth for certain
years.
Table 12.10
Table 12.11
10. Calculate the mass of the bacteria in problem 5 after 12 hours, after
9 days, and one day before the observation started, with the help of
(a) the functional equation f{x^ + x2) = f(x^) x f{x2) and already known
values, and
(b) the functional equation f(t) = 1.5f.
(Note: If one works with systems of coordinates with logarithmic scaling of one
or both axes, one can graphically recognize additional functions, because the
graphs of the exponential and logarithmic functions are then straight lines.)
Problem 6
Immediately after birth the length and weight of babies is determined. STAT
Is it possible to say “the longer the heavier”? Is there a linear function
GEN
describing the relation between the length and the weight?
The following data result from measurements on newborn babies:
Table 12.12
Length / (cm) 51 53 53 54 50 52 48 52 49 49
Weight w (g) 3,190 3,450 3,840 3,770 3,260 3,020 2,620 3,930 3,000 2,920
Table 12.13
Length l (cm) 48 49 50 51 52 53 54
These data yield a function. We can recognize a tendency for children who
are taller to be heavier as well.
272 THE ART OF PROBLEM SOLVING
If one now forms the quotient of proposition 3 in the above-mentioned way and
then calculates the numbers a and b, one will get the linear function w = 1741 - 5,600
(after rounding). If one inserts the measured values into this equation and consid¬
ers the deviations, one can judge whether or not the linear function found gives a
sufficient description of the matter. It is obvious that the consideration of a larger
amount of data will give a better result. If the investigation is based on a large
amount of such data, one gets a linear function that describes the correlation
between length and weight of newborn babies on average.
Exercise
11. With the help of the following data, one can examine whether there are
noticeable differences between newborn boys and girls. Which babies’
measurements differ significantly from the average?
Table 12.14
Girls
Length 1 (cm) 54 51 54 52 48 52 49 53 50
Weight xv (g) 3,620 2,800 3,590 2,800 2,210 4,170 2,980 3,770 2,950
1 Boys
Length l (cm) 51 53 53 56 54 52 51 50 47
Weight w (g) 2,400 3,900 4,350 3,750 3,920 3,930 3,240 3,870 2,610
From the techniques presented in this chapter, one should begin to get a genuine
appreciation for the role that functions can play in solving problems that might
normally be solved in other ways. At the elementary levels, the techniques should
be more familiar, as the beginning of the chapter showed. Yet these more familiar,
or common, uses have more sophisticated applications. The reader is encouraged
to pursue further applications according to individual interests.
13
DAVID SINGMASTER
273
274 THE ART OF PROBLEM SOLVING
is just long enough for the weight to reach the ground (dell'Abbaco, 1964, pp. 129-
133). Solving this is trickier than the preceding version. To make it even trickier,
suppose the rope is not long enough—then how high from the ground does the
weight hang? Consider Figure 13.1.
ALG
Figure 13.1
h +h=L
d\ + d2 — D
(hi-h)2 + d\ = l\
(h2-h)2 + d\ = l22
h=h
d\ d2
It took me several tries to solve this, but I have now found a reasonably direct
solution. Using a/b - c/d if and only if a/b - (a + c)/(b + d), we deduce LfD =
h/d\ — li/ d2.
Dividing the third of the given equations by d\, and using the previous
sentence, we have
fhi-h') 2 fh) 2
+1= —
di di D
v J K 9 k 9
Symmetry Saves the Solution 275
Symmetrically, we have
fh2~h 2
+ 1 =
rn 2
di D
\ v y
Hence
In —h hi - h I l1-!?
d\ di o2
and so
so
The simplicity of this result naturally makes one try to find a simple deriva¬
tion, and symmetry does it! Reflect one tower in the horizontal through W to get
Figure 13.2.
o
©
©
The two parts of the rope now form a straight line that is the hypotenuse
of a right triangle with sides D and (hx - h) + (h2 - h), so we immediately have
L2 = (/*! + h2 - 2h)2 + D2.
The earliest known example of the simpler version with the pigeons and the
fountain is in Fibonacci (1857, pp, 331-332,398-399). Vogel's (1970-1977) article on
Fibonacci says the problem is of Indian origin, but I have not located any such
versions.
276 THE ART OF PROBLEM SOLVING
Square Hunting
Klamkin, Breault, and Schwartz (1959-1960) posed and solved the following
problem:
ALG The first solver gave only the obvious solution: 50 miles south of the equator.
GEOM The later solver found the general solution but could not find the answers in closed
form.
TRIG
In 1983-1984, the problem and its solution appeared in the Journal of Recrea¬
tional Mathematics (Kakinuma, Barwell, and Collins), and Brian Barwell found a
better solution but did not quite carry it to its conclusion.
The only other version of the problem that I have seen is Perelman (1985,
pp. 18-19, 25-26), but I do not know when this first appeared. Perelman has a
dirigible (later a helicopter) traveling 500 km north, east, south, and west from
Leningrad—where does it land?
About 1986, the problem recurred to me, though I could not remember where
it had come from. I found the following closed-form solution by use of a standard
symmetry before I found the literature.
Figure 13.3
Symmetry Saves the Solution 277
or
for some integer k, which is the number of times the explorer circles the earth. The
solution of [1] and [2] or [2'] is not obvious and frustrated the 1960 solver.
Assuming we are not at the special cases Oi = 0 or <1>2 = ft, we can symmetrize the
problem by setting
Qi + <i>2
0 =
2
[3]
(jj2 — 0j
P=
so
01 = 0 - p, 02 = 0 + p . [4]
(Barwell [Kakinuma, Barwell, & Collins, 1983-1984] does the same.) Setting
a = d/2R and using several trigonometric identities, [2'] becomes
Assuming again that we are not in the special case k - 0, we can continue. Cross-
multiplying equation [5] gives a quadratic equation whose solution is
(Barwell [Kakinuma, Barwell, & Collins, 1983-1984] got to the quadratic but
thought the solution too messy to consider and went to consider approximate
solutions.)
This solution leads to a number of further questions and variations of the
problem, which I have treated in an unpublished paper. I give one example below.
None of this would have been possible without the basic symmetrization step of
changing from and <J>2 = Cq + v/R to <1> - (3 and <1> + [3. This process goes back to
the Babylonians, who used it to solve such problems as x + y = 10, xy - 24 by setting
x = 5 + (3, y = 5 - P, so xy - 25 - [32 = 24 and (32 = 1. (See p. 86)
This symmetrization has the effect of eliminating the first-order term in a
quadratic and is equivalent to completing the square. The square bear problem is
a trigonometric problem and hence more complicated, but the symmetrization is
even more essential.
As an example of a new problem inspired by the above, consider the follow¬
ing (Singmaster, 1993):
Hiawatha, the mighty hunter, has traveled far in search of game. One
morning he gets up, has breakfast, heads north and travels 10 miles for¬
ward in a straight line. Seeing nothing, he stops for lunch. After lunch,
he heads north and travels 10 miles forward in a straight line and finds
himself where he started in the morning. Where on earth is he?
If the value of the horse is known, this is a determinate problem. But the value
is often not given and the problem is then indeterminate, and one wants the
smallest integral solution or a general solution. Such problems date back to the
Chiu Chang Suan Ching (1968, pp. 86-88, problems 10,12,13), written c. A.D. 100.
The above example occurs there with an unspecified object worth 50. Diophantos,
c. A.D. 250 (Heath, 1964, pp. 139-149), gives a somewhat general approach. Histori¬
cally, these problems are significant as being among the first where negative
solutions are treated naturally—for example, by Fibonacci (1857, pp. 228-258,
327-349) in 1202.
Symmetry Saves the Solution 279
When one extends to more than two people, there are two forms of the
problem.
Type I: The 1th person says, "If I had a, times what all of you have, then I could
buy the horse."
Type II: The zth person says to the i + 1st person (considered cyclically), "If I
had fli times what you have, then I could buy the horse."
In both types, all known versions have all a, rational and we will assume this.
The Type I problem has an obvious symmetry, and symmetry will guide us
to a solution. Let be the amount the i * person has, and let T be the total amount
they have, that is, T = lx,. Let h be the value of the horse or other object being
purchased. Since T - x, is the amount that all except the zth person have, our
equations can be written as:
xi + - x,) = h . [7]
(The use of the symmetric expression T already makes the problem much simpler.)
Solving [7] for x; gives us
h - a{T
Xi = [8]
1-fli '
and adding these for all i eliminates the individual x/s, giving us
1 _ fl;
T = h'L-TL [9]
1 - a, 1-Ui
so
ai [10]
l+S = h
1 - ai 1 -at
One can readily find the smallest integers T and h satisfying [10], but a more
satisfactory expression arises from rewriting [8] as
h-T
Xi =T+ [11]
1 - ai
T = nT + (h-T)I, —- [12]
1 Cli
(n-l)T=(T-ft)Z [13]
1 -ai’
280 THE ART OF PROBLEM SOLVING
which shows that T and T -h are a simpler pair of unknowns than T and h. The
least integers T and T -h satisfying [13] correspond to the least integers T and h
satisfying [10]. For example, the Chiu Chang Suan Ching example has n = 2, a\ = 1 /2,
«2 = 1/3 so [13] becomes T = (T - h)7 / 2 or 2 T = 7 (T - h), and we can take T =7,
T - h = 2, giving h = 5,xi = 3, X2 = 4.
Fibonacci (1857) devotes a chapter of 30 pages to this problem, giving many
variations with up to seven persons or five horses, an inconsistent example, and
examples with negative solutions. He treats many of the problems by another
method in a later chapter. The numbers involved can be quite formidable—in one
case h = 35,839,901.
There is a c. 1390 version (Libro d'abaco, 1973, p. 141) with an extra mutual
friend, which leads to a two-parameter solution space, though only one solution
is given. The easiest example with a negative solution appears in a Provengale
manuscript of c. 1430 (Provengale Arithmetique, 1984, pp. 49-53) and is just our
problem for n = 5 people, with (at) = (1/2,1/3,1/4,1/5,1/6).
Unfortunately, integral values for T and h do not guarantee integral values
for all Xj, but one can always convert a solution to a solution with integral .
Type II problems give the equations:
Xi + ai xi + \ = h. [14]
x\ +a\xz = h,
X2 +U2X3 = h,
SO
X2 = h-az X3,
and
x\ + a\(h - «2 X3) = h
and
*3 + fl3 X4 = h,
X3 = h- fl3 *4,
XI - flia2(/z -a3X4) = h- a\ h,
getting
When n = 4, we have
Xl [1 + (—1)” + 1 fll fl2 • •• fln] = /l [1 — fll + fll fl2 - • • • + (—l)”fll fl2 • • • [15]
Trying to determine X2, X3, and so on from [14] and [15] gives a mess, but we
can simply note that the problem looks the same wherever we start from so we get
Xi [1 + (-1)” + xai at+1... a{ -1] = h [1 - fl, + a/ fl/ +1 - ... + (-l)"fli flf+1 • • • «»-1] [16]
Here we have again used symmetry, in a simpler way than in type I problems but
just as usefully.
Here the i* person says, "If I had the purse. I'd have fl, times what you have." ALG
Again, for more than two people, there are two types of problems, depending on
whether "you" means "all of you" or just the next person in sequence (taken
cyclically). In type I, we get the equation xt + p = fl, (T - x,), while type II gives the
equation x, + p = fl, x, + 1.
Here p is the value of the purse. We see that this is the same problem as men
buying a horse but with a horse of negative value and with negative multipliers!
Algebraically, the solution in this situation requires nothing new. However, there
are two ancient problems which are related to this form. The simpler gives the
following equations for n = 3.
Xl + X2 = «3
X2 + *3 =
X3 + Xl = fl2
282 THE ART OF PROBLEM SOLVING
Diophantos (Heath, 1964, p. 135) gives a general approach to this and ex¬
amples, with (a,) = (20, 30, 40) and (22, 24, 27, 30). The general form of this has
n equations stating that the sum of all but one of the values is a given value. If we
let T = Ex, as before, then the equations are T - x; = at. We proceed symmetrically by
adding all these to get nT -T= Har, hence T = —^-y Iat, from which Xj-T - ai are
readily found.
The other related problem is the "bloom" of Thyramides (see Heath, 1981,
pp. 94-96). This has n + 1 unknowns: x, xv x2, ■ . ., xn; and we are given x + x,- = a{
and x + x1 + ... + xn = S.
As before, let T = X x;- , so x + T = S. We again proceed symmetrically by
adding to get nx+T = 'Lai.
Combining this with x + T=S gives us (n - 1) x = £ a, - S, so x and all the x,
are determined.
The simplest version of this problem has two men and a bicycle, which cannot
- carry a passenger. The men wish to get to the next village, a distance D away, as
GEN soon as possible. Assuming the men have equal walking speeds and equal riding
speeds, it is easy to see that the optimum method is for one to ride halfway and
then leave the bicycle for the other to pick up. Thus, each person rides halfway
and walks halfway.
The earliest occurrence I've seen of this simple version is in Gaston Boucheny
(1939, pp. 77-78), but more complex versions go back to about 1900 (Loyd, 1960,
pp. 88, 160-161), and I suspect there must be older examples. [I have since found
1906 versions with two persons and a bicycle and with four persons and a car
which can carry two passengers (Laisant, 1906, pp. 127-132).]
Complexity can be introduced in several ways. First, the speeds of the various
individuals may differ. Second, we may have more people. Third, the vehicle may
carry more people (versions of the problem have a tandem bicycle, a motorcycle
with passenger seat, a taxi, a truck or van, or a pony). In all cases, the vehicle cannot
carry all the people, so the vehicle must shuttle back and forth—possibly with
different speeds when loaded and unloaded. I was led to investigate this type of
problem when I saw an elegant graphical solution in Menninger (1961, pp. 100-
101). Unfortunately this solution is incorrect, but it led me to the following
algebraic solution, and the graph is still the best way to see the equations.
Let us first consider the simplest case of two men and a bicycle. The men walk
at rates w, and ride at rates r,. We graph position d against time t for both men and
the bicycle as in Figure 13.4.
The first person rides at rate rx until he leaves the bicycle, and he then walks
© at rate wv The bicycle stays put until the second person gets to it, and then he rides
it the rest of the way. Obviously, the optimum solution has both people arriving
at the goal at the same time. Letting P1 = (tu dx), P2 = (t2, d2) and G = (T, D) in the
Symmetry Saves the Solution 283
Figure 13.4
diagram, and remembering that distance is velocity x time, we have five un¬
knowns—T, tv t2, dv d2—and the following five equations:
d\ = dz [17]
d\ = r\ ti [18]
dj = W2 t2 [19]
Although these have some symmetry and can be solved, the process is not
elegant, especially if one tries to deal with the more general problems. In particular,
if we have a vehicle which can return, then the diagram looks like Figure 13.5,
where r is the unloaded return speed of the vehicle.
Figure 13.5
284 THE ART OF PROBLEM SOLVING
After some trial and thought, I realized that the problem becomes much
simpler if one uses different variables that are less obvious but turn out to make
the problem much more symmetric and hence readily solvable.
We now let f,- be the time the 1th person rides. Then each person rides for time
f, and walks for time T - t{ to cover the total distance D, so we have
n U + Wi (T - ti) = D. [22]
(If we let di = n U, then the point P2 has coordinates (T - f2, D - d2), which is not an
obvious improvement!) We can solve [22] for f, as:
D - w,T
U = 23]
If we have a bicycle (or a pony which does not head back), then it also travels a
distance D, giving us
2>i U = D. [24]
2-!— (D - WiT) = D,
Ti - Wi
n ■ TjWj
D -l = T
n - Wi [25]
‘ ri - Wi
V
2>,f,-r(T-If,) = D. [24']
\
D S-Tl±L
L in + r)wj
+r [26]
Ti ~ Wi ti - Wi
Symmetry Saves the Solution 285
Concluding Remarks
References
JAN TROjAK
287
288 THE ART OF PROBLEM SOLVING
if a line p is mapped by T onto itself (so that p = p'), we say that p is a "fixed (or
invariant) line" in T.
An image p' of a line p in the transformation T can always be constructed by
joining the images of two distinct points of the line p—that is, (X, Y) e p implies
(X', Y) e p\
The fact that the circle k has the center S and the radius r will be denoted by
k = (S, r). An image of such a circle in a congruence transformation T is a circle
^ = (S , r) of the same radius r, the center S' of which is the image of the center S
of the circle k.
The inverse transformation T'1 to the transformation T maps every point X
of the plane to the point Y, the image Y' of which in T coincides with X.
The most trivial case of the congruence mapping is the identity transforma¬
tion, mapping every point of a plane to itself. All the points of the plane are
therefore invariant under the identity transformation.
Axial Symmetry
Rotation
In the general case, there is just one fixed point—the center of rotation and
there are no fixed lines. In the case, though, where the absolute value of the angle
of rotation is equal to n, the rotation has some special properties. The center of this
rotation is called the "center of symmetry," and a rotation of this kind is called
290 THE ART OF PROBLEM SOLVING
the "central symmetry." While the center of symmetry is still the only fixed point,
all the lines passing through it are fixed lines. In addition, every line p is parallel
to its image p' in the central symmetry (see Figure 14.3).
Figure 14.3
In the case in which the absolute value of the angle of rotation is equal to 2n,
the rotation turns into the identity transformation, with all points, and therefore
also with all lines, being invariant.
If the rotation T is given by its center S and by the angle of the magnitude co,
then the transformation T_1 inverse to T is again the rotation with the center S and
the angle of rotation of the magnitude -co. Every central symmetry is inverse to
itself: T = T-1.
Translation
Figure 14.4
In the case where the segment defining the translation is of zero length, the
translation is the identity transformation. Any translation that is not the identity
transformation maps every line p to a line parallel to p, the lines of the direction of
the translation are fixed lines, and there are no fixed pointsof this translation.
If the translation T is given by the oriented segment AB, then the translation
T_1 inverse to T is determined by the oriented segment BA.
As an example of a problem from the abovementioned category, let us
consider one whose solution demonstrates the use of a rotation.
Problem 1
GEOM
Two distinct lines p and q are given, and a point S. Draw a square
ABCD that satisfies the following conditions:
Solution
In the search for a solution, we assume first of all that the given problem
really has some solution, and we sketch its picture in Figure 14.5. From now on,
we shall be working backwards while looking for the relationships between the
known and unknown (but desirable) objects (in most cases, unknown points).
. Y'
Construction
71
1. Draw the image of the line p in rotation R (in point Si with magnitude —).
(See Figure 14.6):
(a) Construct the image X' of an arbitrary point X e p in the rotation R.
(b) Draw a line p' through X' perpendicular to p.
2. Denote the intersection point B =p' nq.
3. Find the point Ae p as a point that is mapped by the rotation R into B (or
as an image of the point B in the inverse rotation RA to R).
4. Complete the construction of the square ABCD, when its vertices A and B
and the center S are known.
Figure 14.6
294 THE ART OF PROBLEM SOLVING
When constructing the desired square, we can accomplish every step of the
construction in a unique way, with the exception of step 2: The first step yields a
unique line p' as an image of the given line p.
The second step, though, can result in one, none, or infinitely many points.
We are getting a single point B in case the lines p' and q are not parallel, to which
again just one point A e p exists, mapped by R to B and the construction results in
the unique solution.
If the lines p' and q are parallel and distinct, they have no common point. No
point B means no point A and, therefore, no solution of the problem.
If the lines p' and q are identical, then any point of the line q = p' can be taken
for B, and every such point B provides us with one of infinitely many squares
satisfying the conditions of the given problem.
The third condition of our problem, slightly altered, changes our problem
into another one which provides the opportunity to employ a central symmetry
(or, if the reader prefers, a rotation of the magnitude tt) to get a solution easily.
Problem 2
Two distinct lines p and qand a point Sare given. Draw a square ABCD
that satisfies the following conditions:
Solution
an intersection point A = CSnp and complete the construction of the square ABCD
without further difficulty.
Construction
We can consider the given lines p and Cj, and the given point S, drawn already
in our plane. (See Figure 14.8)
1. Draw the line p' as an image of p in the symmetry S (reflection in point S).
(a) Draw the image X' of an arbitrary point X e p in the symmetry S.
(b) Through the point X' draw the line p || p.
2. Denote C = p' nq.
3. Draw the image A of the point C in the symmetry S.
4. Complete the construction of the square ABCD when the vertices A and
C and the center S are given.
296 THE ART OF PROBLEM SOLVING
In an analogy with the previous problem, we can conclude that once the point
C exists (the lines q and p' are not parallel), a unique resulting square ABCD is the
only solution, while there is no solution when the lines q and p' are parallel and
distinct, and there are infinitely many solutions in the case of the coinciding lines
q = p'. In the last case, any point of the line q = p' can be taken for a point C and
used as the vertex of square ABCD with the center in S and the vertex A lying on
the given line p.
The character of the next problem suggests the use of a reflection in a line
because the geometrical object we are looking for is required to be symmetrical
with respect to the given line.
Congruence Transformations in Problem Solving 29 7
Problem 3
Given two lines p and q and a circle k= (S, r) with the center S and
the radius r. Draw an isosceles A ABC having C ep, A ek, and B eq,
for which CA = CB = d is a given distance and p _L AB.
Solution
Figure 14.9
298 THE ART OF PROBLEM SOLVING
Construction
Lines p and q and a circle k are given in the plane; d is a given distance. (See
Figure 14.9)
1. Draw the image q' of the line q in the reflection R through line p.
2. Denote A = k n q'.
3. Draw a circle m = (A, d).
4. Denote C = m n p.
5. Find B as an image of A in the reflection R.
6. Draw the triangle AABC.
Figure 14.10
Congruence Transformations in Problem Solving 299
Let us look through the construction again, surveying each step in order to
find out whether that particular step can always be accomplished and, if this is the
case, in how many ways it can be done.
The first step can certainly always be realized, and the resulting line q' is
unique.
In the second step, looking for the points common to line q' and the circle k,
we can find two distinct intersection points (Av A2), a single point in which the
line q' touches the circle k, or no common point at all.
In the third step, therefore, we can have two circles mx = (Av d) and m2 =
{A2, d), only one of them, or none.
In the fourth step, every circle m (if there are any) can have, in general, two
distinct points, one point, or no point in common with the line p. Thus, the fourth
step can result in up to four distinct points in the most favorable case, shown in
Figure 14.10, where they are denoted as Cv C2, C\, C2.
To every point A of the base of AABC there exists just one point B constructed
in the fifth step of our construction as an image of A in the reflection R.
In step six, we can always construct AABC except in the case in which the
points A, B, and C are collinear, which will be the case when C is the point of
tangency of the line p and the circle m.
Summarizing the results, we can say the following:
a. There are four distinct solutions in the case in which step 2 results in two
distinct points and step 4 yields four distinct points (see Figure 14.10).
b. There are two solutions in the case in which
(i) only one point A was found in step 2 and the single circle m = (A, d)
intersects the line p in two distinct points (C, C) or
(ii) there are two points Av A2 provided by step 2; but of the two circles
mx = (Av d) and m2 = (A2, d), only one intersects the line p in two
distinct points.
c. In all other cases, there is no solution.
Problem 4
Draw a triangle A ABC, if its side c = AB, its median ta, and an acute GEOM
angle w = Z{ta, b) between its median ta and its side b are given.
300 THE ART OF PROBLEM SOLVING
Solution
Let us suppose that such a triangle really exists and that it looks like the one
sketched in Figure 14.11. Let us denote by SA the midpoint of the side a. Then
fa = ASa.
Figure 14.11
Points B and C are symmetrical with respect to the point SA, and since B lies
on the circle with the center A and the radius c (which we shall refer to as k), we
can look for the point C on the circle k', which is the image of the circle k in the
already mentioned central symmetry S with the center SA. The point C is also the
endpoint of the side b, the angle of which with the median fa is known to be co.
Thus, if the angle to = Z(b, fa) is positioned as ZKASa, the point A' found as
an image of the point A in the symmetry S, and the circle k' = (A', c) drawn, the
point C is an intersection point (if it exists) of the circle k' and the side AK of the
angle to.
Congruence Transformations in Problem Solving 301
Construction
Figure 14.12
302 THE ART OF PROBLEM SOLVING
From the above discussion it follows that there are two solutions of the
© problem when AK intersects the circle k' in two distinct points, as in Figure 14.12.
There is just one solution when AK is tangent to the circle k', while there is no
triangle with the given properties if AK has no common points with the circle k'.
Problem 5
Two circles k, = (S^ r,) and k2 = (S2, r2) and a pair of the distinct points
GEOM
A and B are given. Draw a segment XY parallel to AB, the endpoint X of
which lies on ku the other endpoint Yof which lies on k2 and XY = AB.
Solution
Assuming again that a solution exists for the given circles k2 and k2 and the
© pair of points A and B, let us sketch one of the possible configurations (Figure
14.13).
The point Y can be considered as being the image of the point X in the
translation T determined by the oriented segment AB. While the point X runs on
along the circle kv the point Y traces another circle k\ = (S\, rx), the image of kx in the
translation T. The circles k1 and k\ are congruent, and the center S'a of the circle k\
is the image of Sj in T. The point Y can be found as an intersection point of the two
circles k2 and k\, and the point X is the point of the circle kx mapped by T on Y.
Congruence Transformations in Problem Solving 303
Construction
Let us consider that the circles k^ and k2 and the points A and B are already
positioned in the plane.
In the most favorable case (shown in Figure 14.14), we can have four distinct
resulting segments, when each of the pairs of the circles k2, k\ and k2, k\ has two
distinct intersection points. In the less favorable cases, there can be three, two, one,
or no intersections, resulting in three, two, one, or no solution segments.
Figure 14.14
304 THE ART OF PROBLEM SOLVING
Problem 6
Given two distinct lines pand q, a circle k = (S, r), a point C, and a
distance d. Draw a triangle A ABC, such that A e k, Be q, AS II p, and
AB= d.
Solution
(©© We assume again that our problem has at least one solution, the picture of
which is shown in Figure 14.15. Let us investigate relationships between the given
objects and those we need in order to be able to construct the resulting triangle.
Because we know the length of the side AB ( = d) of AABC, as well as the
direction of the segment AB (AB || p), we can take the vertex B for an image of the
vertex A in a translation T of the magnitude d and of the direction given by AB.
But also, on the contrary, the vertex A can be taken for an image of the vertex B in
the translation T_1 (inverse to T) of the same magnitude d and of the opposite
direction, given by the opposite ray BA.
Let us consider, in the first case, the use of the translation T. Provided the
point A lies on the circle k, point B has to lie where all the images in T of all the
points of the circle k lie—that is, on the image of the circle k in the translation T.
This, as we already know, is again a circle k' = (S', r), where S' is the image of the
point S in T. Since our problem requires the point B = A' to lie on the given line q,
we can find B as the intersection point of this line with the circle k'. Two such points
of intersection can, in general, be found, and still another two may be constructed
as intersection points of the image k" = (S ", r) of the circle k in the inverse translation
T-1 with the given line q.
" ~ " B d
Figure 14.15
Congruence Transformations in Problem Solving 305
Once the vertex B is determined, we shall find the last vertex A of AABC as
the point of the circle k mapped by the corresponding translation to the point B.
Construction
Let us suppose that the given objects p, q,k- (S, r), and C are positioned in a
plane and a non-zero distance d is known. Then the desired triangle can be
constructed in the following way:
1. Draw an image k' = (S', r) of the circle k in the translation T, where T has
magnitude d on direction AB. (See Figure 14.15)
2. Draw an image k" - (S", r) of the circle k in the translation T-1.
3. Denote {Bi, Bz) = k' n q and {B3, B4} = k" n q.
4. Determine the points A\, Ai, A3, A4 of the circle k mapped to the points B1,
Bz, B3, B4, respectively.
5. Draw the triangles AA1B1C1, AA2B2C2, AA3B3C3, AA4B4C4, if they exist.
Already in the discussion of our problem we have accounted for all the
solutions possible, and we have found that in the most favorable configuration we
can have four solutions, when the given line q intersects both of the circles k and
k" in two distinct points, as is the case in Figure 14.16.
All the other configurations of the given elements can result in three, two,
one, or no solutions, depending on the relative positions of the line q and the circles
k' and k".
Figure 14.16
306 THE ART OF PROBLEM SOLVING
Problem 7
When a line p, two distinct circles /c, = (Su p) and k2 = (S2, r2), and a
GEOM distance d are given, draw the line parallel to p cutting the circles /c,
and /c2 in the chords X, V, and X2Y2, respectively, in such a way that the
sum Xy Y, + X2Y2 is of the given length d.
Solution
Q Provided that a solution of the problem exists, we can assume that the
configuration looks like the one in Figure 14.17.
Let us consider a case when the circle k2 is so shifted in the direction of the
line p that the two points Yy and X2 coincide. Let us denote this shifted position
of the circle k2 by k2 and its center by S2. Then the shifted position of the point
Y2 e k2 can be taken for the image X\ of the point Xj e 1] in a translation T
determined by the direction of the line p (and of whatever orientation) and by
the length d. The point X\ has to lie on an image k\ = (S^, ry) of the circle ky in
the translation T. The center S2 of the circle k2 lies on the axis o of symmetry of
the shifted chord X2Y2, which is the axis of symmetry of the segment S1S,1 at the
same time. The circle k2 = (S2, r2) can therefore be drawn as an image of k2 in the trans¬
lation T+ determined by the distance of the center S2 from the axis o in the direction
parallel with p (i.e., perpendicular to o) and orientation from S2 to the axis o.
Thus, the point Yy (if it exists) can be found as an intersection ky n k2. The line
through Yy parallel to p may intersect ky in another point Xlr the circle k2 in another
point X\ (which is also the point of k’) and the circle k2 in the points X2 and Y2
Since the chord Y1X'1 of k2 is the image of the chord X2Y2 of the circle k2, they are
of the same length and so d = XyYy + YyX\ = X1Y1 + X2Y2.
Figure 14.17
Congruence Transformations in Problem Solving 307
Construction
The given elements p, klf and k2 are placed in the plane, and the distance d is
known.
1. Draw the image ST of the center Si in the translation T and the image circle
fcT = (ST, ri) of the circle fci. _
2. Draw the axis o of the segment SiS'i-
3. Construct the point S2 such that
(a) S2 e 0
(b) S+2S2\\p.
4. Draw the circle kt = (S2, rz).
5. Denote Yi =k\ nE
6. Draw the line q through Yi parallel to p.
7. Denote the points Xi, Yi, X2, and Y2 for which {Xi, Yi} = q n k\
and {X2, Y2} = q o kz.
The first four steps of the construction can always be accomplished just one way.
The fifth step, though, can result in two, one, or no points YL In Figures 14.18
through 14.21, we have two distinct intersections {Ya, = *T n kf The line q
through Y1 and the line q through Vx yield two pairs of chords, XxYv X2Y2 and
Z V Z V2 respectively.
In step 7, some intersections of q and/or q may coincide. Thus, in Figure
14.19, one chord of the circle k2 is of zero length, while in Figure 14.20, the chord
308 THE ART OF PROBLEM SOLVING
Figure 14.19
of zero length is one of the circle kv and in Figure 14.21, both of the chords of the
circle k2 have lengths equal to zero.
Step 5 results in one solution in Figure 14.17 and in no solution in Figure 14.22.
It can be observed that (besides some other, more intricate cases) there is always
no solution in the case in which r1 + r2< d/2 or the line p is of such a direction that
no line q parallel to p can be found intersecting both of the circles kx and k2 at the
same time.
Figure 14.20
Congruence Transformations in Problem Solving 309
The general subject of geometry was characterized for the first time by
German mathematician Felix Klein in his Erlanger's Program, published m 1872.
The subject of Euclidean geometry, in particular, then appeared to be the investi¬
gation of those properties of figures in a Euclidean space that do not change when
a figure is transformed by any congruence transformation (also called isomet y )
of Z space. Since a congruence transformation preserves the shape as well as the
310 THE ART OF PROBLEM SOLVING
,K2
Figure 14.23
The first thing we do when constructing the triangle is choose the position
of one side, say BC, by placing the points B and C somewhere in a plane so that
Congruence Transformations in Problem Solving 311
WEI LEE
Introduction
The Swiss mathematician Leonhard Euler, who wrote a paper in 1736 solving
the popular Konigsberg Bridge problem, was the first mathematician to work in
the area of graph theory. In recent years, graph theory has become an important
area of mathematics where a great deal of research is being conducted. The current
interest in graph theory among mathematicians is perhaps due to the fact that it
has a large number of applications in both academic and business areas.
Although it can be a powerful tool, graph theory is not often suggested as a
method for problem solving. In this chapter, I will try to offer some insight into
using graph theory in problem solving, with the hope that it will be used more
often.
AUTHOR'S NOTE: I would like to give my special thanks to Dr. Stefan A. Burr and Mrs.
Bessie Burr, who put in a tremendous amount of time proofreading this chapter. Their
corrections, comments, and encouragement made it possible for me to hntsh this chapter,
313
314 THE ART OF PROBLEM SOLVING
Graph theory has also played a very important role in mathematics contests.
Some contest problems are graph theory problems in disguise. When you look at
these problems, you may not know where to start. But if you familiarize yourself
with some basic graph theory definitions and theorems and their applications,
these problems will become more solvable.
This chapter consists of six sections. Section 1 introduces graph theory defi¬
nitions. Section 2 presents selected problems that illustrate several strategies
incorporating graph theory and problem solving. Sections 3 and 4 use these graph
theory tools to solve the problems. Section 5 introduces directed graphs to solve
problems involving contests. Section 6 provides additional exercises for readers to
practice what they have just learned about graph theory.
1. Basic Terminology
The graphs in graph theory consist of points and lines that connect these
points. The graph in Figure 15.1a consists of 8 points vv v2 , • . ., Vg, and 9 lines
eve2,..., e9. The graph in Figure 15.1b consists of 5 points vvv2, ■ ■ ■, v5, and 8
lines eve2,..., e8.
There is a difference between these two graphs. The one in Figure 15.1b has
arrows on each line, representing the direction of the line. We call this kind of
graph a directed graph. We call the graph in Figure 15.1a an undirected graph.
We use capital letters to represent graphs. Sometimes we add subscripts or
superscripts—such as Gv G2, G1, and so on—to represent graphs. In a directed
graph, we assume that if there are two lines between two points, these two lines
must have opposite directions. In Figure 15.1b, there are two lines with opposite
directions between the points vx and v2.
V4 e5 V5
Figure 15.1a
v2 e5 v3
Figure 15.1b
Graph Theory 315
The points in a graph are called vertices. We call lines in an undirected graph
edges, and the lines in a directed graph arcs. The edges in a graph may be
represented by writing both endpoints of that edge. For example, vx v2 means the
same as e1 in Figure 15.1a.
The most important thing in a graph is its number of vertices. Questions
such as whether the lines are straight or curved are not important. Let's look at
graphs Ga (Figure 15.2a) and G2 (Figure 15.2b). It looks as if they are two different
graphs, but actually the two graphs are isomorphic (i.e., they are the same graph).
Both have six vertices (vv v2> v3, v4, v5, v6), and exactly the same pairs of vertices
are joined by edges.
Figure 15.2a
Figure 15.2b
The definition of some common terms used in graph theory will be helpful
here. In an undirected graph, if an edge e{ connects the vertices v} and vk, then we
say Vj and vk are the endpoints of the edge eu and the vertices Vj and vk are incident
with the edge et Moreover, we say Vj and vk are adjacent vertices. Similarly, if two
edges e, and e- share a common vertex, then we call et and adjacent edges.
In an undirected graph, if there are edges between every two vertices in that
graph, then we call it a complete graph. If the graph has n vertices, we denote the
complete graph as Kn. Figure 15.3 shows the graphs K2, K3, K4 and K5. Note that in
K4 and K5 the intersection of the two crossing lines is not a vertex in the graph.
316 THE ART OF PROBLEM SOLVING
Figure 15.3
v2
Figure 15.4
Vertices with degree 0, like v7, are called isolated vertices. The following
theorem is concerned with the degrees of vertices.
Theorem 2. In any graph, the sum of the degrees of the vertices equals twice the number
of the edges.
Graph Theory 317
Before we prove this theorem, let's see if the theorem applies to the graph in
Figure 15.4. The sum of the degrees of the vertices of this graph is2 + 3 + 3 + l +
4 + 1 + 0 = 14, and the number of edges in the graph is 7, which agrees with the
theorem.
Proof of Theorem 1. Let's imagine that we take away all the edges in the graph. Then
the degree of every vertex in the graph is 0. Therefore the sum is also 0. Then we
put the edges back into the graph one by one. Subsequently, every time we add
an edge to the graph, the degrees of the two endpoints of that edge are increased
by 1. Therefore, the sum of the degrees of the vertices is increased by 2. Therefore,
the sum of the degrees of the vertices is exactly equal to twice the number of the
edges.
Corollary 1. In any graph, the number of vertices with odd degrees is even.
Proof Let's use the contradiction method (induction) to prove this corollary.
Assume there is an odd number of vertices with odd degrees in a graph. Then the
sum of the degrees of the vertices is an odd number. But theorem 1 states that the
sum of the degrees of the vertices equals twice the number of the edges. Therefore,
it's an even number. This is a contradiction.
y3
G:
Figure 15.5a
318 THE ART OF PROBLEM SOLVING
v4
G
• v6
Figure 15.5b
v\
Figure 15.6a
v4 v3
Figure 15.6b
Graph Theory 319
There are many terms in graph theory. We have just listed a few of them here.
We will introduce some more when we need them later.
2. Examples
Example 1
Prove that in any collection of six people either three of them mutually
know each other or three of them mutually do not know each other.
Proof. Let's pick any vertex in Ke, say v\. The five edges between vertex v\ and the
other five vertices v2, vs, v4, vs, V(, are either red or blue. According to the Pigeon¬
hole Principle (see Chapter 8 for detail), at least three edges of the five have the
same color. Let's assume that v\ v2, v\ vs, v\ v4 are red edges (see Figure 15.7). Now
consider the triangle v2 vs v4. If one of the edges v2 vs, v3 v4, or v2 V4 is red, then we
will have a red triangle. Otherwise, if v2 v3, vs v4, v2 v4 are all blue, then the triangle
v2 Vs V4 is a blue triangle. This proves that there must exist a triangle all of whose
edges are colored by the same color.
320 THE ART OF PROBLEM SOLVING
Figure 15.7
There is another way to translate the puzzle into a graph theory problem
without using edge coloring. We still use six vertices to represent the six people.
If two people know each other, then we connect these two vertices by an edge. If
two people do not know each other, then we don't connect these two vertices. For
5^ example, if everybody knows each other, then it becomes a complete graph K6. Or
if nobody knows each other, then the graph becomes a graph with six vertices
without any edges. Of course, those are the extreme cases. Usually there will be
some vertices connected with edges and some without. At any rate, the problem
we want to solve becomes this:
Prove that for any undirected graph with six vertices, there must exist
either a triangle or three vertices without any edges between them
(three isolated vertices).
If we use the terms subgraph and complementary graph, then the puzzle can be
translated to this:
Prove that in any undirected graph G with six vertices, either G has a
subgraph K3, or G’s complementary graph has a subgraph K3. (See
exercise 1)
Example 2
First, we need to translate this problem into a graph theory problem. Again,
we use vertices vh v2,..., v\7 to represent the 17 scientists. If 2 scientists discuss
the first topic, then we use red edges to join these two vertices. Similarly, for the
second topic we use blue and for the third topic we use yellow. Then we have a
complete graph K\7 and edges colored red, blue, and yellow. What we need to
prove is that there exists a triangle with all edges the same color.
Proof. Choose a vertex, say v7. The 16 edges from v7 to the other 16 vertices are
colored with red, blue, or yellow. According to the Pigeonhole Principle, at least
6 edges have the same color. Let's assume the edges vi v7,v7v7,..., ve v7 are
colored red. Then let's look at the complete graph Ke with the vertices
v\,V2,..., ve■ Obviously, if there is a red edge, then we have a red triangle. And
if all the edges in Ke are not red, that means the edges in Ke are either blue or yellow.
Then from the result of example 1, we know that there exists either a blue or a
yellow triangle in the graph Ke. That finishes our proof.
Figure 15.8
322 THE ART OF PROBLEM SOLVING
Let's look at some more examples related to the degrees of the vertices.
Example 3
Proof. Since we have the terminology and corollary 1 from section 1, this problem
© becomes obvious. Let's assume every vertex represents a person. If two people
know each other, then we put an edge between these two vertices. The number of
people a person knows is just the degree of that vertex. Using corollary 1, we have
the solution to this problem.
Example 4
Just by looking at it, you can see this problem is different from examples 1
and 2. But after careful analysis, you will notice the following facts: (a) This
problem actually is a graph theory problem. Whether or not the 9 points are on a
circle is not important, (b) The graph representing this problem is a complete graph
K9 (the reader should try to find out why this is true), (c) If we can solve the
following example 5, then example 4 will be trivial.
Example 5
Prove: Color all the edges in a complete graph K9 either red or blue.
COMB Then either there are three vertices for which all three edges between
these three vertices are blue (a blue K3) or four vertices such that all
six edges between these four vertices are red (a red K4).
Graph Theory 323
The condition in example 4 is that there is at least one red line in any
three-point triangle. Translated to graph theory language, the condition will be,
"There doesn't exist a blue triangle." If we proved example 5, then it suggests that
there must exist a red K4 in the example 4. This is exactly the result of the exam¬
ple 4.
Proof. Use v\, Vi,..., V9 to represent the vertices in K9. Every vertex is the endpoint
of eight edges that have been colored either red or blue. Now let's consider several
possible cases for the problem:
Case 1. If there exists a vertex, say v\, such that at least four of the eight edges
adjacent to it are colored blue.
Let's fix these four blue edges, say vx v2, vx v3, vx v4, vx v5 (see Figure 15.9).
Then of the six edges joining the vertices v2, v3, v4, v5, if one of them were blue then
we will have a blue triangle. Otherwise, if all of these six edges are red, then it's a
red K4.
Therefore, the result in example 5 exists in case 1.
blue
Figure 15.9
Case 2. If there exists a vertex, say v\, such that of the eight edges joined to it at
least six of them are colored red.
Let's fix these six red edges, say vx v2, vx v3 ,..., vx v7. Consider the edges
joining the six vertices v2,v3,..., v7> and the complete graph K6 with these six
vertices. The edges in that K6 were colored either red or blue. From the result of
example 1, there must exist a triangle with its edges the same color. If there is a
blue triangle, then we are done. Otherwise, if we have a red triangle, assume a red
triangle v2 v3 v4. Since the six edges joining the vertices vv v2, v3, v4 are all red, we
therefore have a red K4.
324 THE ART OF PROBLEM SOLVING
We have proved that we will have the result if case 1 or 2 holds. Now we
need to prove that whatever way we color the K9, at least one of the two cases will
13s) hold. Actually, if neither case holds, then every vertex has exactly five red edges
connected to it, because if there are less than five, we get case 1; and if there are
more than five, we get case 2. Now let's look at the subgraph of our K9 having just
its red edges. This graph has nine vertices and every vertex has degree 5. In other
words, all nine vertices have odd degrees. But this contradicts corollary 1. This
contradiction shows that at least one of the two cases should hold.
The proof in example 5 looks long and complicated. But if you explain the
proof by drawing a graph on the blackboard, then you will find it's actually a very
easy way to visualize and solve the problem.
The reader should analyze the differences in the proofs of example 5 and
examples 1 and 2 and study the way we use the result of corollary 1 in the proof
of example 5. That should give a better understanding of using graph theory as a
tool to solve problems.
Example 6
Proof. First, we need to construct a graph G representing this situation. The graph
G contains nine vertices v\, Vi,..., V9 representing the nine mathematicians. If two
people can speak the same language (no matter what language), then there will be
an edge between these two vertices.
Assume that there do not exist three mathematicians who can speak the same
language. We will prove that if this assumption exists, then every vertex in the
graph G has at most degree 3.
Actually, if there is a vertex of G, say vv with degree > 4, then pick four edges
joined to vv say vx v2, v3 v3, v1 v4, vx v5 (see Figure 15.10). Since vx can speak at most
three languages, then among v2, v3, v4, v5, there are at least two people who can use
the same language to speak to vv say v2 and v3. But now, vv v2, v3 can speak the
same language. This contradicts the assumption. This proves that the degree of
every vertex in G is less than or equal to 3.
Graph Theory 325
Figure 15.10
Now let's fix vv According to what we just proved above, vx has at most three
edges joined to the other three vertices. Since G has nine vertices, at least 9-1-3
= 5 vertices have no edges joined to vv Let's assume v5, v6, v7, vs, v9 have no edges
joined to vv Then, similarly, pick v5 and its degree < 3. Therefore, among
v6, v7, vs, v9, there is at least one vertex not joined to v5. Let's assume that vertex is
v9. Now we find three vertices vv v5, v9, and there are no edges between these three
vertices (see Figure 15.11). But this contradicts the condition of the problem, which
says that there are two people who speak the same language among any three of
them. That means that among vv v5, v9, there is at least one edge between two of
the three vertices. From this contradiction, we solve the problem.
v9
Figure 15.11
We will use the following graph G (Figure 15.12) to introduce some more
important terminology in graph theory.
326 THE ART OF PROBLEM SOLVING
v2 e4 v3 e8 v\ e13
Figure 15.12
Consider the sequence of the combination of vertices and edges. For instance,
in Figure 15.12, Pl = {vv ev v2, e4, Vg, e7, v8, ew, v9}. Here, for every edge, there are
two vertices that are the endpoints of that edge. This kind of sequence is called a
path (see Figure 15.13a). We can say this path starts at v1 and ends at v9 or that this
path connects vx and v9.
Usually we use a capital letter—for instance P or Py P7> and so on to represent
a path.
In a path, vertices and edges may be repeated. For instance, in Figure 15.13b,
P2 = [v2, e4, v3, e7, Vg, e5, v2, e4, v3, e8, v4}. Since P2 satisfies the condition that the
vertices before and after every edge are the endpoints of that edge, P2 is a path. In
this path, e4, v2, v3 occurs twice. A path like Plr having no vertex or edge repeated,
is called a simple path.
v2 e4 v3
Figure 15.13a
e4 v3 e8
P 2=
Figure 15.13b
Graph Theory 32 7
If the first and last vertices of a path are the same, then we call it a cycle.
Let's use Figure 15.12 as an example: P3 = {v3, e8, v4, e12, v9, ew, v8, e7, v3] is a
cycle (see Figure 15.14a). There are also cycles in Figure 15.14b, such as P4 -
[v2, e5, v8r e9, v4, eu, v9, e10, v8, e6, v6, e3, u2). Similarly, if no vertices and edges repeat
in a cycle (except the first and last vertices), we call it a simple cycle.
v3 e8 va u2 v4
P4:
Two distinct vertices or edges in a graph G are independent if they are not
adjacent in G. A set of pairwise independent edges of G is called a matching in G.
For example, in Figure 15.15a, Mx = (e2, e5] and M2 = {ev e3, ee) are matchings
in G. Usually we use M, Mv M2, and so on to represent the matchings.
Figure 15.15b
328 THE ART OF PROBLEM SOLVING
Figures 15.16
For any graph we can consider its matchings. Matchings in bipartite graphs
are very interesting and powerful tools for problem solving.
Theorem 2. If a bipartite graph has a perfect matching, then X and Y must have the same
number of vertices.
Example 7
There are some balls. Every ball is colored either red or blue, and every
ball weighs either one pound or two pounds. We know that both red and COMB
blue balls appear and both one-pound and two-pound balls appear. Prove
that there exist two balls with different colors and different weights.
Let's translate the problem into a graph theory problem. We can construct a
bipartite graph as follows: X consists of two vertices a\, a2, representing the colors
red and blue; Y consists of two vertices b\, b2/ representing the weights one pound
and two pounds. If there exists a red one-pound ball, then we put an edge between
fli and b\. If there exists a red two-pound ball, then we put an edge between a\ and
b2, and similarly for the other two cases. Now we have a bipartite graph. The
conditions of the problem indicate that (a) both red and blue balls exist, which
means that both vertices a\ and a2 have edges incident to them, and (b) both
one-pound and two-pound balls exist, which means that both vertices b\ and b2
have edges incident to them. What we want to prove is that there exist two balls
with different colors and weights, which means there exist two edges with differ¬
ent endpoints. In other words, there is a perfect matching.
Example 7 has now been translated into a graph theory problem as the
following: We have a bipartite graph with parts X and Y. Both X and Y contain
two vertices, av a2 and bv b2, respectively. We also know that no vertex has degree
0 (i.e., each vertex has at least one incident edge). Prove that this bipartite graph
has a perfect matching.
Proof. Let's pick a vertex from X, say a\. According to the assumption, there is at
least one incident edge, say a\ b\. Now let's consider a2; similarly, there is at least
one edge incident to a2. If the edge incident to a2 is a2 b2, then a\ b\ and a2 b2 will be
a perfect matching in the graph, and the problem is solved. If the edge incident to
a2 is a2 b\ (see Figure 15.17), then consider b2. As before, there is at least one edge
incident to it. If the edge is a\ b2, then the edges a\ b2 and a2 b\ will be a perfect
matching in the graph. Otherwise, if the edge at the b2 is a2 b2, then a\ b\ and a2 b2
will be a perfect matching. Therefore, in all cases, we have a perfect matching.
Figure 15.17
330 THE ART OF PROBLEM SOLVING
Example 8
There are 10 men and 10 women at a dance. Every man knows exactly
2 women and every woman knows exactly 2 men. Prove that after
suitable pairing, every man can dance with a woman he knows.
Proof. Let's randomly pick a\ to represent a man. There are two edges incident to
13^ this vertex. Pick any one, say a\ b\. And from the two edges incident to b\, pick
either one, say az b\. And from az, similarly we find bz,_Since there is only finite
number of vertices in the graph, one of the vertices will be repeated eventually.
Clearly the repeated vertex will be a\. Then this is a cycle of the graph. Let's delete
this cycle from the graph, and the remaining bipartite graph will still have the
property that every vertex has degree 2. Then, using the same method for the
remaining graph, we will get another cycle. Repeating this procedure, we can draw
the following conclusion: The graph can be split into several cycles. (The graph in
Figure 15.19 is an example of this proof.)
It's not hard for us to prove that this bipartite graph has a perfect matching.
First, let's notice a fact that every cycle in the bipartite graph has an even number
of vertices. From the graph in Figure 15.19 we should see this result (we leave the
general proof for the reader as an exercise). Now take any cycle from the bipartite
graph. Starting at any edge, we pick the first, third, fifth,... edges; this is a perfect
matching of that bipartite graph. For example, let's first take look at the above cycle
in the graph in Figure 15.19. Starting at edge ax bv the second edge is b1 a2, the third
Graph,Theory 331
X (men) Y (women)
b2
Figure 15.19
The proof of example 8 not only proves the existence of a perfect matching;
actually it shows the method of finding a perfect matching in a given graph. In
other words, if we have a bipartite graph and all of its vertices have degree 2 (of
course, we don't need the condition of each part having exactly 10 vertices; from
the proof you should see that the number of vertices can be any finite number),
then you just follow this procedure: Pick any vertex and find an edge, then from
that edge find another edge, and so on, until we get a cycle. Then continue the
above procedure until we find all the cycles of that bipartite graph. When we pick
the edges (in odd or even sequences) from each cycle, then we will find the perfect
matching(s) in that bipartite graph.
Construct a bipartite graph with its vertices having degree 2 by yourself.
Then follow the above procedure to find the perfect matching(s). Doing that by
yourself will give you a better understanding of this method.
Example 9
Now let's analyze and translate this problem into a graph theory problem.
First let's construct a graph (the construction is a little bit special). We consider
each row as an vertex. Then we have eight vertices a\, ai,.. •, as. These eight
vertices form the X. Similarly, consider each column as a vertex; then we have
bi, bi,..., bs- These vertices form the Y. If the zth row and )th column have a mark,
then we join a, and bj by an edge. This gives us a bipartite graph (e.g.. Figure 15.20a
is a graph with 16 marks, and the graph in Figure 15.20b is its bipartite graph).
Clearly, all vertices in that bipartite graph have degree 2.
After the construction of that bipartite graph, let's see how to use the graph
theory language to describe the result that we want to prove for example 9. What
we want to prove is that we can put 8 black and 8 white in the 16 marked squares
and make each row and column have exactly 1 black and 1 white. Now assume
that we already put the 8 black and 8 white in the board as required. Let's see if
there is any special property for the 8 black squares. Notice the construction of the
graph, in which every square corresponds to an edge in the graph. That is, the
squares where 8 black are should correspond to 8 edges. And every row has 1
black, which means that for every vertex in air 1 of the 8 edges is joined to it. In
other words, there are no edges with common endpoints. That means these 8 edges
are exactly a matching in that bipartite graph. For the same reasons, the 8 white
squares and their correspondent 8 edges are also a matching in that bipartite graph.
Therefore, what we want to prove in example 9 is the same as the following graph
theory problem: If all vertices in a bipartite graph have degree 2, then we can
separate them into two groups, and each group is a perfect matching in that
bipartite graph.
b2
b3
b4
b5
b6
b7
b8
Proof. In the development of the proof in example 8, we saw that if every vertex
of a bipartite graph has degree 2, then it can be divided into even cycles. In each
cycle, start at any edge, then pick the first, third, fifth, . . . edges that will be a
perfect matching in that graph. Obviously, the remaining second, fourth, sixth,...
edges also yield a perfect matching (see the dashed edges in the graph in Figure
15.19). Since the graph consists entirely of even cycles, this proves that the edges
in the graph can be separated into two perfect matchings.
Readers might like to try to solve the next two examples independently,
using the method we used to solve example 9. After you do so, you will understand
the method of proof much better.
Example 10
There are n pieces of paper. On the front and back of every piece of
paper there has been written one of the numbers 1,2.n, and all
numbers appear exactly twice. Prove that we can arrange to put all the
pieces of paper on the table, such that the numbers on the pieces of
paper show exactly the numbers 1,2, . . . , n.
Example 11
Readers should try to translate examples 10 and 11 into graph theory prob¬
lems. Of course, constructing a correct graph for the problem is very important.
6) Then compare these two problems with example 9 and see if there are any
relationships between them.
Examples 10 and 11 are all related to matchings in a bipartite graph. The
following is an example of a matching in a non-bipartite graph.
Graph Theory 335
Example 12
matching.
To conclude the above discussion, example 12 can be translated into the
Assume there is a graph with six vertices, and every vertex has at least
degree 3. Prove that for this graph there exists a perfect matching.
Proof. Let's consider vertex a. Since the degree of vertex a > 3, there are at least
three edges incident to a. Let's pick one of them, say ab. Then consider vertex c;
since there are at least three edges incident to c, that means that there is at least
one edge whose endpoints are not a and b. Let's say it's cd. (See the edge cd in the
Now let's see vertex e. We know that there are at least three edges incident
to it. Since the endpoint can not be / there are at least three edges from the
following edges: ea, eb, ec, ed. From there, we know e joins at least one of the
endpoints in both edges ab and cd. Let's assume e is joined to the a and c (see Figure
15.21, where we use dashes to represent the edges currently not belonging to the
matching).
Figure 15.21
Now let's see vertex/. Similar to vertex^, we assume there is no edge between
e and/. And also vertex/has at least three edges incident to it, or/has at least three
edges from the following edges: fa,fb,fc,fd. From there, we know that there is at
least one edge from fb and fd. Let's assume /is joined to d (see Figure 15.21; of
course, there are a lot of edges in the graph—since we don't need them for the
proof, we don't draw them). Now, from the edges in the graph in Figure 15.21, it's
not hard for us to find a perfect matching in the graph. The method is to take out
edge cd from the current matching, instead using edges ce and fd. Then the three
edges ab, ce,fd clearly are a perfect matching.
For an undirected graph G, if there is a simple path P that covers all the
vertices in the graph, then we call that path P a Hamilton path. If a simple cycle
covers all the vertices in the graph, then we call that cycle a Hamilton cycle.
The thick edges in the graphs in Figure 15.22 are examples of a Hamilton path
and Hamilton cycle in the graph, respectively.
Not every undirected graph has Hamilton paths or Hamilton cycles. For
instance, the graph in Figure 15.22a does not have a Hamilton cycle. But a lot of
interesting mathematical problems are related to the existence of Hamilton paths
or Hamilton cycles.
Graph Theory 33 7
vx v2
Figure 15.22a
Example 13
Let's first look at the movement of a knight in chess. The knight first goes
vertically or horizontally for one square and then two squares in a perpendicular
direction. Figure 15.23 shows all possible knight moves from square A. (See
Chapter 18, p. 396)
Now let's translate this problem into a graph theory problem. First, we need
to construct a graph. Let's call the squares in the chessboard the vertices in the
graph. There are 64 vertices. Let's take a look at two squares; if the knight can move
from one of the squares to the other, then we put an edge between these two
vertices. For example, in Figure 15.23, the vertices representing the squares A and
B should have an edge joining them. Similarly, there should be edges between A
and C, A and D,_Then we will have a graph. We are not going to draw this
graph.
338 THE ART OF PROBLEM SOLVING
Now we can translate the problem into a graph theory problem as follows:
Does there exist a Hamilton path from the vertex representing the square
S to the vertex representing the square 77
Figure 15.23
Solution. This problem looks very complicated, but since we have the right tools,
we will be able to use a very easy method to solve it. First, let's notice that the graph
in this example is a bipartite graph, because if the knight starts at a white square,
it must end at a black square, and vice versa. Therefore, if there is an edge joining
two vertices, respective squares must be one white and one black. Clearly, we can
separate the vertices into two parts X and Y. They are white and black squares.
And the endpoints of every edge must be such that one of them belongs to X and
the other one belongs to Y. Of course, X and Y each contain 32 vertices.
In a bipartite graph, if there is a path which has 2,4,6,... vertices, of the start
and end vertices of one of the paths, one belongs to X and the other belongs to Y,
respectively. This fact can be easily explained by the graph in Figure 15.24 (this is
a path with eight vertices). If a path's start and end vertices both belong to X or
both to Y, then this path must contain 3, 5, 7,. . . vertices. From that assumption
we see that it is impossible for the knight to start at the upper-left comer and then
go through every square and reach the lower-right comer. Because if there exists
such path, then the bipartite graph will have a Hamilton path with 64 vertices. But
the first and last vertices both belong to X, since both of them represent white
squares.
Graph Theory 339
X r
Figure 15.24
Example 14
There are 10 people at a party. We know that every one of them knows
at least 5 of the others. Prove: After suitable arrangement, we can
make these 10 people sit around a round table and every one will know
the people sitting next to him or her (left and right).
As in several previous examples, let every vertex represent one person, and
there will be an edge joining two vertices if the respective two people know each
other. Then we will have an undirected graph with 10 vertices. If we put these 10
people around a round table and let every person know the people sitting next to (^3
him or her, this is equivalent to finding a Hamilton cycle in the graph.
Therefore, the problem can be translated into a graph theory problem as
follows:
Proof. Let's use the method of contradiction. Assume that there doesn t exist a
Hamilton cycle in graph G. Then we will try to find a contradiction.
Let's consider the two vertices u and v, which are not joined to each other in
the graph. Then we add an edge uv to the graph. If, after adding this edge uv, the
graph still doesn't have a Hamilton cycle, then we add this edge uv to the graph
G. And if, after adding the edge uv, there is a Hamilton cycle, then we stop.
340 THE ART OF PROBLEM SOLVING
Continue the edge-adding procedure until we don't need to add any more edges.
At the end, we will have a graph called G'.
Clearly, the graph G' will have the following properties: (a) G' does not have
a Hamilton cycle; (b) every vertex in G' has at least degree 5 (because G' is obtained
by adding edges to the graph G, the degrees of vertices in G' are at least as great
as in G); (c) after adding an edge in every two unconnected vertices in G', we will
have a Hamilton cycle.
Now let's pick two unconnected vertices, say u and v, and add an edge uv as
described in (c). We have a Hamilton cycle. When we take out the edge uv, we have
a Hamilton path with the endpoints u and v. Let's assume the sequence of the
vertices as follows: vx (= u), v2,v3,, vxo (= v) (see Figure 15.25).
u = V\ Vw = V
*-
V2 V3 Vi V5 ve V7 Vs V9
Figure 15.25
Now let's look at vertex vv We already know that in G', (a) vx and v2 are
connected, (b) vx and v10 are not connected, (c) the degree of vx is at least 5. There¬
fore, among vertices {v3}, {v4}, {v5}, {ve}, [v7\, {y8}, {v9}, there are at least four vertices
incident to vx. We have the same result for vertex vw; there are at least four vertices
in {v2, vy v4, v5, v6, v7, z>8) incident to vw.
Now let's look at the seven pairs of vertices: {v2, v3], {v3, v4}, {v4, v5},
{^5/ iv6' vt\’ (v7' v8}' iv8' v9} • We want to prove that among these seven pairs of
vertices, there is at least one pair {vjr Vj+X) such that vx is joining vj + x and vxo is
joining V:.
Let's use an example to prove this. As previously mentioned, in
[v3, v4,... ,v9) there are at least four vertices incident to vx. For instance, let's
assume that these four vertices are v4, v5, Vg, v9. Now if there is a vertex in
{^3/ ^4- v7, ^8) joining zho we will have the result. There are also at least four vertices
in {v2, v3,... ,v8] joining u10. So, whatever ways we pick the vertices, there will be
at least one vertex from {v3, v7, z;8} joining vx0.
This proves that there exists a pair of vertices [vjf Vj+X) such that vx joins v.+1
and vxo joins Vj. But now there exists a Hamilton cycle in the graph G'. That is
vx,v2,... ,Vj ,vXQ ,v9,... ,Vj + x ,vx (see graph in Figure 15.26, for which this is the
example when j - 4). This contradicts our assumption that there is no Hamilton
cycle in G'.
Figure 15.26
Graph Theory 341
5. Tournaments
This section introduces a set of special directed graphs, in which there is one
arc between every two vertices in the directed graph. We call these graphs
tournaments. The two graphs in Figure 15.27 are two examples of tournaments with
4 and 5 vertices.
(a) (b)
Figure 15.27
Why do we call these tournaments? Because such a graph can represent the
result of a round-robin tournament. For instance, in the graph in Figure 15.27b, we
assume every vertex ax represents a team, and an arc from al to a} means a, beats a}
(assume there is no tie). Therefore, the graph represents the results of a tournament.
Now let's introduce some more terminology related to directed graphs.
First, in a directed graph, we can also consider paths and cycles. If the arcs
in a path or cycle all have the same direction (see graphs in Figure 15.28), then we
call them a directed path or a directed cycle.
Directed Cycle
Directed Path
Figure 15.28
342 THE ART OF PROBLEM SOLVING
Example 15
Prove that in any tournament there exists a directed Hamilton path. Let's
take a look at the examples in Figure 15.27; the two graphs in Figure 15.29 are
Hamilton paths.
Figure 15.29
Graph Theory 343
Proof. The proof procedure actually is a method of finding the required Hamilton
path. Let's first pick any directed path a\, a{,..., af. We will prove that if this
path does not contain all the vertices in the graph, then we can extend this path. _.
For example, say a is not in the path. Now let's look at the following: (a) If the arc
connecting the vertices a and a\ is from a, then a,af ,... ,ap is a longer directed
path; (b) similarly, if the arc connecting vertices a and ap is from ap, then
a\, ai ,..., ap', a is a longer directed path; (c) if both (a) and (b) don't exist, then
there must exist the arcs from af to a and a to ap' (see Figure 15.30).
Figure 15.30
Now let's observe the following sequence a{, a{,..., af, where a;- is the first
vertex such that an arc goes from a to af (see a{ in Figure 15.30). Then clearly the
path af.Uj_ {, a, a'.af is a longer directed path.
Therefore, if a directed path does not contain all the vertices in the graph,
then we can always extend this path. This proves that there must exist a directed
Hamilton path, that is, one that contains all the vertices in the graph.
Example 16
arc from a, to a, or there is a directed path from a, to a, with two arcs (if
this path’s vertices are ahak,aj, then a, beats ak and ak beats ay.
Therefore, a, lost to ak and a, beats ak).
Proof. We want to find a person who wins the most (The answer may be not
unique—why?), that is, the vertex that has the largest outdegree. Let's assume this
vertex is a\ (see Figure 15.31), and assume a\ beats 02,03,, as, that is, a\ wins
s -1 times. But a\ lost to as+1,... ,an. What we want to prove is that for every vertex
dj in {fls +1, as + 2,.. •, an), there must exist an vertex ak in {ai, 03,... ,as} such that ak
beats aj-r that is, there is an arc from ak to aj.
fls +1 «s + 2 an
Figure 15.31
Let fly be any vertex in {fls + 1, as+2,..., an}, so fly beats ak. Now let's consider
the result between fly and [a2, fl3,..., fls}. If fly beat all of these teams, then fly won s
© times. Therefore, fly wins more times then av This contradicts the assumption of
the problem. a1 wins the most. This proves that in {a2, fl3,..., as}, there must exist
an ak such that ak beats fly.
Example 17
Let's use graph theory language to describe the problem we want to prove:
There is a tournament that satisfies the following condition: Every vertex's inde¬
gree is not equal to zero (i.e., every team lost at least once). Then there exists a
directed cycle with three arcs.
«l a2
Figure 15.33
6. Exercises
The following are some exercises for readers to practice using graph theory
as a tool to solve the problems. The starred (*) problems are a little bit difficult, but
readers are encouraged to read the similar examples in the chapter and try to solve
them.
1. Prove: For any undirected graph G with six vertices, either G has K3 as
subgraph, or the complementary graph of G has a subgraph K3.
2. A club has 99 members. Everyone in the club knows at least 66 other
members. Prove that we can pick 4 of them to play bridge such that all 4
know each other.
3. In exercise 2, if every member of that club knows exactly 66 of the 99
club members, prove that we may not be able to pick 4 of them to play
bridge such that all of them know each other.
4. There are 9 people v1( v2 ,. .. , vg. Suppose that v: shook hands with 2
people; v2, v3 shook hands with 4 people; v4, v5, v6, v7 shook hands with
5 people; and v8, v9 shook hands with 6 people. Prove that we can find 3
of these 9 people such that all 3 of them shook hands with each other.
5. The edges in the complete graph /<18 are colored either red or blue. Prove
that whatever way we color the edges, there must exist four vertices such
that all six edges between these four vertices (a K4) are colored by the
same color.
6. Prove: Among 14 people, either 5 of them mutually know each other or
3 of them mutually do not know each other.
*7. Prove: Among 19 people, either 6 of them mutually know each other or 3
of them mutually do not know each other.
8. We color the edges in the complete graph K66 by using four colors: red,
blue, yellow, green. Prove that whatever way we color the edges, there
Graph Theory 347
exists a triangle and all three of its edges (a K3) are colored by the same
color.
*9. Prove: If there are n{n> 2) people in a meeting, there must exist 2 people
who know exactly the same number of people in the meeting.
10. Prove that the cycle(s) in a bipartite graph must contain an even number
of vertices and an even number of edges.
11. There are n men and n women at a dance. We already know that each
man has danced with at least 1 woman but has not danced with all the
women. Similarly, we know that each woman has danced with at least 1
man but has not danced with all the men. Prove: There exist 2 men
x/, x{ and 2 women y/, y{ such that the following pairs have danced
before: x/, y,' and x2', y2'; but the following pairs have not danced before:
x/, y{ and x2', y/.
12. Find a perfect matching in the bipartite graph in Figure 15.18.
13. There are 6 men and 6 women attending a dance. After the dance, they
recall the number of people they have danced with. The numbers are as
follows: 3, 3, 3, 3, 3, 5, 6, 6, 6, 6, 6, 6. Prove that at least one of them
made mistake.
14. Try to put 8 white and 8 black chess pieces in the squares with marks X
in Figure 15.20a, and make every row and column have exactly 1 white
and 1 black.
15. There is a classroom with 25 seats arranged as a square of 5 rows and
5 columns. Assume that at the beginning every seat has a student sitting
in it. Can we change the students’ seats in such a way that every student
can be moved to all adjacent seats?
16. Construct a graph which has 40 vertices and 400 edges but no triangles.
*17. Prove that any graph with 40 vertices and 401 edges must contain
triangle(s).
*18. Prove: If a graph G has 2n + 1 vertices and rt + n+ 1 edges, then the
graph G must contain triangle(s).
*19. There are 2n people at a party. We already know that everyone knows at
least n people. Prove that we can pick 4 people from these 2n people
such that these 4 people sit at a round table and every one of them knows
the people sitting on either side of him or her.
*20 There are 2n people at a party. We already know that each person knows
at least n people. Prove that after suitable arrangement, we can have
these 2n people sit around a round table and every one of them knows
the people sitting on either side of him or her.
21. Decide if there is a Hamilton path or Hamilton cycle in the grid graph in
Figure 15.34. (Hint: Consider the two cases in which n is odd or even.)
348 THE ART OF PROBLEM SOLVING
n vertices
Figure 15.34
22. Prove that there is no Hamilton cycle in the graph in Figure 15.35.
16
A Different Solution for
Problems With Extreme Values
A Didactic Journey Into
the World of Jakob Steiner's Ideas
KARL KIESSWETTER
ROLAND J. K. STOWASSER
LENNI I. HAAPASALO
349
350 THE ART OF PROBLEM SOLVING
In order to show that Steiner (1971) didn't think only geometrically, we offer
another citation:
Although I may now think that the synthetic method is the most
appropriate one for observation and reasoning in these fundamental
theorems and their future development, it is possible that the analytic
approach could be the correct viewpoint in later forthcoming questions
to pursue the subject in appropriate cases.
How is teaching done in school today? Problems with extreme values are not
approached from one-sided viewpoints; that is, they are not restricted to the
best-known analytic models. Rather, it is the aim in mathematics and, accordingly,
in the teaching of mathematics, to develop comprehensive exemplary solutions.
But in the teaching of mathematics, one must not suggest that such solutions are
unique and that it is sufficient merely to find them. Above all, it should not be the
aim to make repeated use of only one method, like a cooking recipe. In addition,
methods other than Steiner's (polynomial transformation, contour lines, recur¬
sions, and so forth) should be a permanent part of the teaching process.
Steiner focuses his endeavors on the main theorem: that, as he observed, the
circle has the greatest area of all plane figures with a given constant boundary. As
is typical for him, he proves this theorem in several different ways. He traces his
procedure of symmetrization to the following "first fundamental theorem": Of all
triangles having the same base and the same perimeter, the isosceles triangle has the largest
area.
Another proof, which he used in later works and which is very simple, is
based on Steiner's "second fundamental theorem": If two sides of a triangle are given,
then the area of the triangle is largest if the two sides include a right angle.
Problems With Extreme Values 351
Problem 1
Given a closed rope with length /. How should the rope be placed so that
it borders the largest possible area? What size is this area?
Figure 16.1
Problem 2
Given a simple closed set of connected edges in a plane, with the lengths
GEOM of the edges svs2,...,sn. Imagine that the edges are bars from a construc¬
tion set that are connected with joints. How should this set of edges be
placed so as to border the largest possible area?
Problem 3
(A contrived story!) About 2,500 years ago, if someone wanted to join the
Pythagoreans, he had to demonstrate his talents. Pythagoras himself is said to have
proposed the following problem for this purpose (for simplification, sides as well
as lengths of sides are designated the same way):
GEOM
Given a convex quadrilateral with sides s1 =7, s2 = 9, s3 = 3, and s4 = 11.
What should be the measure of the angle between and s2 in order for
the quadrilateral to have the largest possible area?
Problems With Extreme Values 353
Figure 16.3
This formulation of the problem will result in direct and strong associations
with the Pythagorean theorem and then to the rectangle. The quite independent
discovery of the second fundamental theorem should now be only a question of
time. One should easily recognize that 72 + 92 = ll2 + 32 and that this is a sufficient
condition for a quadrilateral to have maximum area. Despite this, it is not a
satisfying solution from a didactic viewpoint, because there are almost no possi¬
bilities for generalization; the route to making general connections is blocked.
Problem 4
Given a simple open set of connected edges with the lengths of the edges
Sl> s2.sm, which lies on one side of a straight line g, between the
end points P0 and Pm (on g) (conditions as in problem 2). How should this
set of edges be placed so that, along with the edge P0 Pm, the largest
possible area is enclosed?
Figure 16.4
354 THE ART OF PROBLEM SOLVING
Of course, this isn't a general optimal formulation, but we think it's the best
compromise. Its advantages are that students are able, with comparative ease, to
find their own ideas for the solution and that the entire solution, as well as
individual elements of it, can be used for similar problems in an appropriate way.
Along with the elements of the proof, the student gets an instrument for a relatively
elegant and transparent solution for a whole class of problems with extreme
values. We want to illustrate a detailed approach to the solution of problem 4. The
reader can formulate his or her own strategies.
Step 1
P2
Every student knows this formula for the area. But he or she must also realize
that movement of P2 straight toward PQ causes oppositely oriented changes in g
and h. In addition, it is not self-evident that the student can effect a change to a
simpler form by taking another viewpoint, using a rotation.
Figure 16.6
Problems With Extreme Values 355
Now one need only address one changing parameter (h in Figure 16.6), whose
maximum value is obviously s2 The area F is maximized only when the angle
between s1 and s2 is a right angle.
As we mentioned, this is such an important fact in Steiner s work that he
speaks of it as a fundamental theorem.
Step 2
It is easy to conclude that none of the possible regions could have maximal
area if it is not convex (turn AP2P3P4 outwardly in Figure 16.4).
Step 3
Let's have a look at Figure 16.7. Consider the hatched regions over P0 Pk and
pk Pm as fixed, and shift the final point Pm along g. Only the drawn angle ak at Pk
changes. But in the meantime, the area of AP0PkPm changes as the angle changes.
The area is maximal if ak is a right angle.
Figure 16.7
Po Pm-
356 THE ART OF PROBLEM SOLVING
Step 4
But that is not all. Many students are going to find learning these concepts
difficult. The level of success can be even greater if other thoughts are added to
complete the proof (from another part of mathematics, namely, analysis). One
remaining question is still unanswered: assuming that border lines such as those
in problem 4 are given, is it always possible to inscribe them in a semicircle in the
desired way?
To answer that question, considerations of continuity (intermediate value
concept) must be made. Those could be replaced by the following consideration
of plausibility (whose positive pedagogical value should not be underestimated).
We proceed from the semicircle having a diameter d = + s2 + ■ ■ ■ + sm. We
can lay off the edge drawing on this semicircle, as shown in Figure 16.8, because
the semicircle arc is longer than its diameter. Then we steadily reduce the size of
the semicircle and let the "final point Pm wander into the point E." This is always
possible, because we can find semicircles that are surely too small to lay off the
edge drawing in the manner described (e.g., d = max {sk I k = 1, 2,..., m}).
Figure 16.8
Step 5
To emphasize which final gap must be closed, we'll have a closer look at the
third step. There we chose a procedure that enabled us to specify in each case a
larger surface from among those that were under consideration, aside from the
semicircle case." But perhaps our procedure is useless for precisely this semicircle
case. Perhaps another procedure also gives us an enlargement of surface area. A
good way to get a feeling for this problem is through the following example.
Let's consider the open interval I: = (0, 2) and the "enlargement instruction"
x -^f(x) = x + (2 - x)(x - l)2. Then f(x) e I for all x e I, and x <f(x) for all x e I
with x *1, but 1 =/( 1).
Thus, we have an analogous fact. Obviously from this we cannot conclude
that 1 is the maximum of f(x) in I.
But should it be possible to prove the existence of a maximum, then we must
have a maximum in our semicircle case, because that is the only case for which the
improving procedure (enlarging the area) does not work.
Problems With Extreme Values 357
If it is clear what is missing, the teacher can easily proceed without the
difficult proof. However, he or she can point out that a great mathematician like
Steiner himself left a gap at this point. Finally, we would like to show that the
manner in which we set up problem 4 is so helpful that even this missing proof
could be understood by good students. It is to our advantage that we are able to
describe the course of our border line with a finite number of variables.
Problem 5
Given an open rope, how should it be placed so that, together with the
GEOM
segment AE connecting its endpoints, it encloses a maximal surface?
© To show that the maximal surface occurs with a semicircle, students can work
analogously to the solution of problem 4 (transfer, consolidation). Only the trans¬
© formation of thoughts from step 5 leads to greater difficulties. The teacher should
be satisfied with plausibility considerations, noting that the circle is the limiting
case of a suitably chosen regular polygon of n vertices (equal length).
Problem 1 then easily leads back to problem 5. The given closed rope is now
divided into two equal open ropes. With that, the isoperimetric main clause is
proved. With the help of the method of fixed-edge surfaces and other things, this
can be used to obtain another solution for problem 2. We give still another example
to illustrate this approach.
Problem 6
GEOM
The ends of a rope are fastened also to the endpoints of bar S. How is
the rope to be placed so that the area of the enclosed surface is
maximized?
©
16.11
Problems With Extreme Values 359
Solution
There is exactly one circular arc region smoothly bounded by the rope and
the bar. Its radius is r. On the other side of S the rest of the circular area is to be
completed (hatched in Figure 16.11). The entire area is then maximal only if the
region is circular. Therefore,. . . there are several extreme-value problems of the
kind mentioned above. In approaching these problems, one cannot always follow
the detailed solution elements discussed. However, reflecting upon the simplest
kinds of elementary concepts is normally sufficient (e.g., problem 7).
Problem 7
Given an angle a and a bar S, so that S links the two sides of a. In which
position should S be placed so that the largest possible surface is
enclosed?
Figure 16.12
Another point of view and restructuring of the situation brings (as in the first
step) the realization that the isosceles triangle has the maximum area (Figure
16.12). One realizes that problems 6 and 7 could both be used to solve an additional,
similar problem, as follows.
360 THE ART OF PROBLEM SOLVING
Problem 8
Given the angle a and a rope of length I, whose endpoints A and E can
be shifted along the sides of a. How should one position the rope so that
the largest possible surface is enclosed?
Figure 16.13
(If the reader does not have the solution, he or she should recall the method
of fixed-edge surfaces.)
Epilogue
We hope that the reader had fun with Steiner's "World of Ideas" and that
this world of ideas will bring pleasure to students as well.
Our objective will have been achieved if the reader uses some of the problems
given here in his or her lessons and thereby improves the spectrum of the math¬
ematical handling of extreme-value exercises. Because Steiner's works are good
reading, one can glean some valuable suggestions, such as the statement that of
all n-gons with the same perimeter, the regular one has the largest area, and so on.
For the teacher, the following historical and educational notes may be helpful:
The farming of the fertile field of isoperimetric problems has, of course, a
long and interesting history (Porter, 1933). Euclid had already reserved a tiny
portion of his Elements for talking about the largest areas enclosed by isoperimetric
(equal perimeters) rectangles (Stowasser, 1976a). As Pappus of Alexandria (c. a.d.
320) said, Zenodor (c. 180 b.c.) proved some statements from the lost book about
isoperimetric figures for comparison of polygons having equal perimeters with
regular polygons and the circle (Pappus, 1932).
Problems With Extreme Values 361
Appendix
Historically Relevant
Extreme Value Problems Solved
Elegantly by Means of the Level-Line Method
R. J. K. STOWASSER
The two-dimensional isoperimetric problem was the subject of the preceding chapter.
By specializing in isoperimetric rectangles, and with suitable alterations, elementary and
elegantly solvable problems result, producing, moreover, an important general solution
method (its analytical form is called "Lagrange's multipliers for extrema with constraints").
Of course, the ancient Greeks knew how to solve the following isoperimetric problems.
Problem 1A
GEOM
Among all isoperimetric rectangles, find the one with the greatest area.
Problems With Extreme Values 363
Our "modern" solution uses the Cartesian coordinates system. We put the isoperimetric
rectangles in the coordinates' corner and see that the upper-right vertices lie on a straight
line with slope -1.
The increment of the area with growing w<h is obvious and produces the rigorous proof
that for 10 = h the area of the rectangle becomes maximal.
The strip on the right has less The upper strip has less area
area than the upper strip than the strip on the right
Figure 16.15
The classical case is not very interesting. One can literally "see" the solution because of
the symmetries. Altering problem 1A a little is much more fascinating.
Problem 2A
The "characteristic line" for the upper-right corner is a straight line with slope -1/2. Now
you cannot "see" the maximal area rectangle any longer.
One has to do some calculations, appropriately with a calculator or a computer; for
example, for w + 2h = 360 m:
Consequently, the piece of land at the river should be twice as wide as it is long: w = 2h.
The rigorous proof is analogous to the one in problem 1A.
We quickly solve another extreme value problem for the rectangles on the river.
Problem 3A
GEOM
Which of the “river rectangles” has extreme diagonal length?
Solution
Figure 16.17
Problems With Extreme Values 365
Simple similarity arguments prove: The "river rectangle" with minimal diagonal length
runs away from the river twice as far as it is wide.
With problems 2A and 3A we can enter into the level-line method for geometrically
solving extreme-value problems. For the "river rectangles" with w + 2h = constant we have
as level-line field the segments between the positive x-axis and the positive y-axis with slope
-1/2.
For rectangles with constant diagonal length, the level-line field consists of quadrants of
circles around the origin of the coordinate system which lie in the first quadrant. (For further
extreme-value problems, one should also consider rectangles equal in area. Their level-line
field consists of equilateral hyperbolas.)
The solution of problem 3A now takes the tangential circle to the characteristic line
w + 2h = 360.
The level-line method, shown here in an extremely simple case, is very useful in many
other extreme-value problems.
Problem 4A
(Hint: Find the loci for the points Y, for which the measure of angle
AYB is constant.)
Figure 16.19
366 THE ART OF PROBLEM SOLVING
Problem 5A
Solution
We draw a circle with P at its "highest" point and enter the acceleration g of the falling
body and the acceleration g\ operating along the slope, which is the perpendicular projec¬
tion of g. S or si mark the distances of the free fall or along the slope, respectively.
2 2
Because of s = 0,5gt and Si = 0.5 gj t and because of the right angle in the semicircle,
sl/'s = gi/g is true.
This means that the level lines for the constant time are circles with P as their "highest"
point, and therefore the following drawing shows the solution of the extreme-value prob¬
lem 5A.
Figure 16.21
Problems With Extreme Values 367
Problem 6A
Solution
Here we use the Heron-Fermat principle, according to which the course a billiard ball
takes from A to B with reflections on its way has ar extreme length. We use the level-line
field, which describes constant lengths of a stretch with a "bend." An ellipse defines equally
long stretches with a bend starting from the foci A and B.
Figure 16.23
368 THE ART OF PROBLEM SOLVING
Therefore, in this case, the confocal ellipses with A and B as foci form the level-line field.
The ellipses touching the circle (the round table) produce the solutions for the Alhazen
problem.
Remark
In case A and B are equidistant from the center of the circle, there is a very simple solution
without using the Heron-Fermat principle:
Figure 16.24
Figure 16.24 shows the three possible solutions. The proof is given by the circumferential
angle theorem. (Think about it yourself!)
References
Rademacher, H., & Toeplitz, O. (1957). The enjoyment of mathematics. Princeton, NJ:
Princeton University Press.
Steiner, J. (1971). Gesammelte Werke (2nd ed., Ed. K. Weierstrass, 2 vols.). Bronx,
NY: Chelsea.
Stowasser, R. J. K. (1976a). Extremale Rechtecke: Eine Problemsequenz mit Kurz-
filmen. Der Mathematikunterricht, 3.
Stowasser, R. J. K. (1976b). Kiistenschiffahrt, Landmessen, Billard-Drei Problem-
felder der Geometrie. Der Mathematikunterricht, 3.
Yaglom, J. M., & Boltjanski, W. G. (1961). Convex figures. New York: Holt, Rinehart,
and Winston.
17
The Problem of the
Duplication of a Cube
HANS K. KAISER
hroblem solving is an art. This art has a lot to do with skills such as creativity of
mind, flexibility of thinking, experience in the application of mathematical tech¬
niques, and the like. How can we learn this art? One way of mastering this art is
"learning by doing," but much insight can also be gained by analyzing the way
particular problems have been solved by famous mathematicians throughout the
history of mathematics. From this point of view, one of the most instructive
examples is the history of the famous problem of the duplication of a cube. In the
following, we sketch a few important episodes from the history of this problem.
The problem of the duplication of a cube is the following:
Given a cube. Construct one side of a cube which has a volume twice GEOM
This problem is one of the three classical problems of antiquity (the other two
problems are the trisection of any angle and the quadrature of the circle).
371
372 THE ART OF PROBLEM SOLVING
It is said that one of the ancient tragic poets brought Minos on the scene,
who had a tomb build for Glaucus. When he heard that the tomb was
a hundred feet long in every direction, he said:
"You have made the royal residence too small, it should be twice as
great. Quickly double each side of the tomb, without spoiling the
beautiful shape."
He seems to have made a mistake. For when the sides are doubled,
the area is enlarged fourfold and the volume eightfold. The geometers
then started to investigate how to double a given body, without
changing its shape, and this problem was called the duplication of the
cube, since they started with a cube and tried to double it. After they
had looked for a solution in vain for a long time, Hippocrates of Chios
observed that, if only one could find two mean proportionals between
two line segments, of which the larger one is double the smaller, then
the cube would be duplicated. This transformed the difficulty into
another one, not less great.
It is further reported that, after some time, certain Delians, whom an
oracle had given the task of doubling an altar, met the same difficulty.
They sent emissaries to the geometers in Plato's academy to ask them
for a solution. These took hold with great diligence of the problem of
constructing two mean proportionals between two given lines. It is said
that Archytas solved it with half cylinders, Eudoxus with so-called
curved lines, (van der Waerden, 1963, p. 160)
It is most likely that this letter is a fake. But it tells us something about
problem solving. If we cannot solve a given problem we may try to transform it
into an equivalent problem and hope to be able to give a solution of the reformu¬
lated problem. Let us take a look at the contribution of Hippocrates. He claims the
following:
Let a and b be two line segments. If it is possible for us to solve the continued
ALG
proportion a : x = x : y = y : b (e.g., if we are able to construct x and y such that the
continued proportion holds), we can solve the duplication problem. Indeed, if we
solve the above proportion, we obtain x2 = ay and y2 = bx. Hence, we have x4 =
fl2 y2 = a2bx, which yields x3 = a2b. Now we only have to choose for a the side of the
original cube and then we set b = 2a. Then x3 = 2a3, which solves our problem.
How would we solve the continued proportion today? The proportion yields
two equations: x2 = ay and xy = ab. We solve this system of two equations
geometrically: We interpret x2 = ay as a parabola and xy = ab as a hyperbola. The
solutions for x and y we are looking for are the coordinates of the point of
intersection of the two curves (see Figure 17.1).
Duplication of a Cube Problem 373
Figure 17.1
Figure 17.2
Figure 17.3
We draw a coordinate system and fix the point A on the x-axis and B on the
y-axis. Then we position our instrument on the plane and move the ruler in such
a way that A meets KL and B meets GH, with G remaining on the line AO and K
on the line OB. This solves our problem.
Another solution was proposed by Eratosthenes, a younger contemporary of
Archimedes. He was so proud of this solution that he built a bronze model of his
duplication instrument, which he put on top of a stone. And he gave orders to
engrave the instruction for using it in the stone.
The instrument of Eratosthenes consists of three congruent rectangles that
can be moved in a frame. On this frame we find a rod that can be rotated around
point O. On each of the rectangles we draw the diagonals as shown in Figure 17.4.
Figure 17.4
Duplication of a Cube Problem 375
The mathematicians of ancient Greece did not accept the obtained solutions
of the duplication problem as being mathematically correct. They claimed that the
acceptance of such a solution would "destroy the beauty in mathematics. To
understand this attitude we have to make an excursion into the history of mathe¬
matics in antiquity.
The birth of mathematics as a science took place in ancient Greece around
600 b.c. The Greeks used mathematical methods, as did the Egyptians and the
Babylonians before them. But while in Egypt and in Mesopotamia mathematics
was performed by using (sometimes incorrect) recipes without explanations (as
far as we know today), the Greeks tried to find arguments for the correctness of
the various mathematical methods. In other words, ancient Greece introduced the
concept of proof into mathematics. The first proofs of mathematical assertions are
attributed to Thales of Milete. Unfortunately, we do not know how he proved
mathematical results; we only know what he proved. Proclus reports that Thales
was the first to prove (a) that a circle is divided into two equal parts by its diameter
and (b) that when two straight lines intersect, angles are equal.
In the generation after Thales, the most eminent mathematician was
Pythagoras of Samos. He used mathematics as his philosophical credo and tried
to explain the world by mathematics. The doctrine of Pythagoras and his followers
_the so-called Pythagoreans—proclaimed that God had ordered the universe by
means of numbers. Harmony is divine and consists of numerical ratios. If someone
achieves full understanding of this number harmony, he becomes divine and
immortal.
376 THE ART OF PROBLEM SOLVING
If there be two straight lines, and one of them be cut into any number
of segments whatever, the rectangle contained by the two straight lines
is equal to the rectangle contained by the uncut straight line and each
of the segments. (Bk. II, Prop. 1)
On the other hand, Plato made the following proposal for making mathemat¬
ics immune to criticism in the fashion of Zenon. He suggested fixing a number of
postulates that are generally accepted to be true. From these basic postulates other
propositions should be derived by using logical principles.
We find this idea, which we call the axiomatic method today, magnificently
manifested in Euclid's Elements. Euclid formulates the following postulates:
Given two unequal straight lines, to cut off from the greater a straight
line equal to the less.
Figure 17.5
One may use this curve to find two mean proportionals between given line
segments.
We draw EK perpendicular to CD. The similarity of the triangles ADXH
and ADEK yields DH : HX = DK : KE. Since DK = CH and KE = HZ, we have
DH : HX = CH : HZ.
Observe that ADZC is rectangular, hence we obtain HZ2 = CH x DH, or
HZ : DH = CH : HZ. Putting these two proportions together gives
CH:HZ = HZ:DH = DH: HX.
Thus, HZ and DH are two mean proportionals between CH and HX. In order
to solve our duplication problem, we have only to find X on a given cissoid such
that CH : HX = 2:1.
This can be achieved by locating the intersection of the cissoid on the line CP
that determines on the two mutually perpendicular radii OC and OB segments OC
and OP in the given ratio 2 : 1 (again, we make use of the similarity of triangles:
AOPC is similar to AHXC). Again, this solution was unacceptable from Euclid's
point of view, since we only know a pointwise construction of the cissoid under
the sole use of Euclidean tools.
Today we know that the Greek geometers were bound to fail in their efforts
to solve the duplication problem on the basis of Euclid's postulates. It is not
possible to find a construction for doubling a cube by using Euclidean tools only.
We are going to outline the proof of this proposition. We regard the problem from
a different point of view; namely, we shall look at the geometrical problem from
the algebraic point of view and employ a little bit of modem algebra.
Suppose we are given a line segment which we define to be one unit length.
A real number a is called constructible if we are able to construct a line segment
I a I in a finite number of steps from this unit length by using only ruler and
compasses.
If we are given constructible numbers a and (5, then a + (5 is constructible as
well. For a, (3 > 0 the proof is given geometrically:
Duplication of a Cube Problem 379
r
a
A
V
5 A
>
NT
—V-
a+P
Figure 17.6
The other possible cases, a > 0, p < 0; a < 0, P > 0; a < 0, P < 0, are shown in
a similar way.
Similarly, we proceed if we want to show that (X — P is also constructible. The
next two figures indicate the following result: If a and P are constructible, then
a x p and a x 1/ p (p * 0) are constructible.
Figure 17.7
The first is just the task of finding the solution of the system of equations that
represent the two lines
ax + by + c = 0 (a,b,ce Q)
ai x + bx y + c\ - 0 (a\, b\, c\ e Q).
x2 + y2 + dx + ey+f= 0 (d,e,fe Q)
ax + by + c = 0 (a,b,ce Q).
The third is of the same type as the second, since it may be considered as the
intersection of one circle with the common chord of the two circles, algebraically:
the solution of the system consisting of the equations of the two circles
is the same as
Figure 17.8
Duplication of a Cube Problem 381
Hence, we have shown the following important result: The field F of con-
structible numbers consists precisely of all real numbers which can be obtained
from Q by taking square roots of positive numbers a finite number of times and
applying a finite number of field operations.
In modem algebraic terminology, we can formalize this result as follows: If
a is constructible and at! Q, then there exists a finite sequence of real numbers
a0 e Q, cq,..., a„ = a such that Q (cq,. .a,) is an extension of Q (cq,..., a,_ a) of
degree 2, i = 1, 2,..., n. In particular, the degree of Q(a) over Q: [Q(a): Q] = 2r for
some integer r > 0. (Degree 2 means that oc, is the zero of a polynomial of degree 2,
which is irreducible over Q (oq,..., o.i_1).)
With this machinery at hand, we can easily show that the problem of the
duplication of a cube cannot be solved with ruler and compasses alone. Let the
given cube have a side of length 1. Hence it has volume 1. The cube we look for
has volume 2, hence a side of length 3a/2~. But 3a/2~ is a zero of the irreducible
polynomial x3 - 2 over Q. Hence the degree of Q(3V2) over Q is 3. In view of the
above result, we would need to have 3 = 2r for some integer r. Clearly, such an
integer r does not exist.
Our perspective on problem solving is certainly enriched by our little excur¬
sion into the history of the duplication of the cube problem. Most people consider
mathematics as a natural tool for solving problems. But there is not always an
immediate solution at hand, and it is not always immediately obvious which
method we should apply. One way of coping with such a situation is to look at the
problem from different viewpoints or, to put it in a more mathematical language,
try to transform the problem into an equivalent one. We then try to solve the
problem in its new form. We find the approach not only in the history of the
duplication of the cube problem but also, for example, in modem times in various
attempts to solve the famous four-color map problem. The mechanical solutions
of our problem are nice examples of recursive thinking, which is another technique
often used in problem solving. One of the main features of the story told above is
the question of accepting a solution as correct or not. In our case, the attempts at
finding solutions with Euclidean tools only brought forth most of the higher
geometry of antiquity. The proof of the impossibility of a solution on the basis of
Euclid's axioms again is an example of changing from one point of view (geomet¬
rical modeling of the problem) to another one (algebraic modeling of the problem).
So, we see, problem solving asks for a number of skills, of which flexibility of
thinking and creativity are not the least.
Reference
van der Waerden, B. L. (1963). Science awakening. New York: John Wiley.
382 THE ART OF PROBLEM SOLVING
Suggested Reading
MARION KAUKE
SABINE ZILLER
383
384 THE ART OF PROBLEM SOLVING
This chapter will deal with the topic of how even younger children, aged 7
to 12, can be optimally motivated by successful classroom experiences to enjoy
mathematical tasks, promote their skills by puzzling out solutions, and cultivate
a permanent interest in mathematics. This kind of teaching will challenge, and
perhaps inspire, all students, even those who are undisciplined, disheartened, or
not so good. It is not aimed chiefly at the promotion of special talents.
Teachers should place high value on the indirect influences that help develop
students' confidence in their ability to find a solution. The creation of a stimulating
atmosphere for learning is still often neglected. The teacher sometimes begins
immediately with abstract subject matter and urges students to proceed with the
set task. No wonder the majority of children, who are distracted from their joy of
playing by everyday school life, hardly want to follow the teacher's explanations.
Moreover, mathematics has such a formidable reputation that it can intimi¬
date overmotivated, anxious children (symptoms of "math phobia" are described
in specialized literature). It is necessary to alleviate this fear so students at least try
to solve the tasks presented and do not lose the thread too fast or get bored because
the abstract matter taxes their abilities too heavily.
Therefore, the first principle for a mathematics teacher who wants to moti¬
vate is to create a relaxing atmosphere.
student, but this does not preclude increasing step by step one's expectations for
the mathematical abilities of the children.
A teacher with a good sense of humor has the talent to recognize when the
trickiness of a problem has reached its limit and immediately interjects a humorous
comment to defuse the students' frustration. The mental energy gained by this can
be channeled into finding a successful solution. A clever teacher approaches the
situation from the start by assuming the least common knowledge. In this respect,
it makes more instructional sense to dampen motivation than to say, Try a bit
harder," since too much strain can cramp the cognitive aspect of learning. With a
reliance on self-organization, one should not interfere with the natural process,
rather, let it develop by itself.
So the pattern is changed from fear of failure to hope for success. The relative
disassociation from immediate responsibility encourages the type of alienation
that allows the problem to be seen in a different light, thus releasing students from
preconceptions that can hinder creative problem solving.
This clearly is not a mystical luxury but a conditio sine qua non for improving
students' "brain power" and mental energy. The capacity of healthy children
increases, thus lessening boredom and sleepiness and ensuring students optimal
concentration on the task.
After students feel pressures diminish, the teacher should bring them into
the realm of mathematical imagination. This starts with students convincing
themselves that although they may not be mathematics fanatics, they can act as if
they were.
There are a lot of possibilities for stimulating such an attitude, mostly having
to do with the symbol-substance of numbers, their vivid representation of numeri¬
cal facts. The teacher could encourage students to use objects to represent numbers,
such as baseball cards, jet planes, jewels, coins, video games, and so on. The idea
of using Morse code also rouses motivational energies. The teacher can discuss
how one can encode a language so that it is accessible only to "insiders." These
methods of personalizing subject matter and boosting motivation can make stu¬
dents more enthusiastic and give classwork more personal importance.
It is always necessary to find out what occupies the students' minds in each
age group. Armed with this knowledge, the teacher can find various motives and
suitable wordings for the tasks, which would otherwise appear quite dry. Students
should by no means be hindered if they want to put themselves in the position of
an unbeatable detective or a commander whose power depends on the solution o
exciting navigation problems. Students could also put a stop to a criminal's
activities by cracking difficult codes and presenting evidence m a mock court
Mathematical puzzles during a treasure hunt (perhaps set in an imaginary world)
can provide exciting adventures for children of different age groups. Humamtar-
386 THE ART OF PROBLEM SOLVING
ian relief actions with mathematical calculations can inspire the altruistic engage¬
ment of the students. And using e-mail to solve problems with children in other
countries can be an intriguing situation for students.
Psychologists have found that putting problems in context enables stu¬
dents—even the less talented—to solve geometric tasks that usually prove too
difficult when presented as just abstract patterns. The teacher's imagination can
put students in a mindset of building the Egyptian pyramids as the origin of
civilization or of animating a "virtual" future world of computers. A dramatic
biography of a well-known mathematician can promote affection for the subject.
A teacher could use the history of mathematics or a related field to demonstrate
how similar problems have been solved, before students try to "reinvent the
wheel."
Mathematical fantasy does not have to be based in a more or less real past,
present, or future; it could, for example, take the form of imagining that, with the
help of a code word, students could call a mathematical genius who tells them the
solution. At first, these games may seem nonsensical, but psychologically they are
very effective. They can reduce the seriousness of the problem and relegate it to a
higher power, an unconscious intuition, that sometimes becomes blocked or
repressed.
Visualization is a method used by many top mathematicians to simplify the
search for the thought that breaks through the barrier of a challenge. Something
like this imagination now becomes accessible to many people, to some extent,
through computers. For younger students, computer-generated games that trans¬
late abstract mathematical matters into vivid fantasies would be recommended.
The prospects of success can also be unlocked with the magic question. How
would fairy-tale figures—a dwarf, a giant, or a princess—solve these tasks?
Embedding problems in contexts remote from reality can bridge the difficulty of
given problems and demanded solutions by translating problems into psychologi¬
cally simpler terms.
Figure 18.1
that the students come to value not only a successful solution but also the ability
to articulate how they achieved it.
The mathematics teacher should also show students the advantage of ex¬
panding the range of means to the right solution. A narrow path to solving a (^3
problem can lead to psychological barriers, which hinder productive solutions.
An important way of eliminating psychological barriers is to tolerate even
seemingly inadmissible things: Just saying "That's not possible" only testifies to
rigid thinking and destroys the train of thought. It is more constructive to ask the
student to follow through on that train of thought.
It is important to show how a problem can be transformed into equivalent
terms, or without losing its essential features, until one comes to a familiar path to
the solution.
Highly individual thinkers owe their originality and flexibility to free ways
of thinking and to their courage in thinking. Teachers should vividly demonstrate
this to students and allow them to jump from one idea to another in problem
solving.
The following tasks illustrate creative heuristics as part of teaching strategies.
Fairy tales in particular are adapted to children's mentality and are suitable for the
introduction, abstraction, and exercise of creative principles.
These examples are such that every student can find at least one solution or
one statement of the solution. The students' different approaches should then
serve as the starting point for becoming aware of the intuitive procedure, the
finding of strategies, and recognizing patterns. These help to systemize and
facilitate the search for solutions. Everybody should get a chance to try the tasks
and test ideas.
Example 1
In the fairy tale “The Devil With the Three Golden Hairs,” the hero is ALG
given a task by the king: In order to be allowed to marry the king’s
daughter, he must find the three golden hairs of the devil. Arriving in
hell, he asks the devil’s grandmother to draw out the three golden hairs
for him, which she does with cunning.
Now imagine the grandmother is blind, and to make the task easier, COMB
the hero has to bring only one golden hair. Since the devil has
altogether five hairs on his head, three of which are the gold ones, how
many hairs does the blind grandmother have to pull out to be sure she
got at least one golden hair?
The first principle for understanding the real problem is to consider the
extreme case: that the grandmother first pulls out all nongolden hairs.
390 THE ART OF PROBLEM SOLVING
Solution
The task then becomes: How many hairs does the blind grandmother have
to pull out of the devil's head to make sure she got at least z golden hairs?
First of all, for the task to be valid, y < x and z < y, and all values have to be
positive integers, where upon x,y,z> 0.
The extreme case is that the blind grandmother first pulls out all nongolden
hairs; that means she does not get one golden hair in the beginning.
From assumptions 1 and 2 it follows that the devil has x-y nongolden hairs
on his head. So if the grandmother pulls out x-y hairs, the worst case is that these
could be all nongolden. In this case, the {x-y + l)th hair is definitely a golden one,
and therefore so are all succeeding hairs.
To get at least z golden hairs, the grandmother has to pull out x-y-vz devil's
hairs.
{Note: The task can also be made more difficult by leaving out the number
of hairs on the devil's head. The students then have to find out themselves that
this value is necessary in order to find a solution.)
A large variety of similar tasks can be formulated and solved.
Example 2
ALG
We are in the realm of natural numbers, and we now want to have a
closer look at the 2-digit numbers in that realm. Outwardly regarded,
COMB
each 2 digits, pairs of the digits from 0 through 9, constitute a pair. How
many 2-digit numbers are there?
Let us suppose only the digits 0, 1, 2, and 3 are available. What
2-digit pairs can be built from them?
Using Game Strategies 391
The first principle for finding the solution is to account systematically for all
possibilities.
Another principle is to find the solution graphically, by making a picture or
table.
Again, deductive and inductive reasoning play a part in generalizing the
solution. What does the problem look like for additional fours and so forth? What
solution do we get if not pairs but groups of three are wanted? What further
possibilities of generalization can be created?
Solution 1
Solution 2
The first digits can be 1,..., 9; the second can be 0,..., 9. Every possible first
digit can be combined with every possible second digit; that means 1 as the first
digit can be combined with 0,..., 9.
There are exactly 9 possibilities for the first digit and 10 possibilities for the
second digit. Thus, there are altogether 9 x 10 = 90 combinations of digits to form
a 2-digit number.
We could also make this a more general problem:
Solution 1
Solution 2
(^) The first figure can be 1,..., 9, the second, third, and so forth can be 0,..., 9.
Every possible first digit can be combined with every possible second digit
and every possible third digit; that means 1 as the first digit can be combined with
0, . . ., 9 as the second digit (10 possibilities each) and with 0, . . ., 9 as the third
digit. That makes already 9x10x10 possibilities.
There are exactly 9 possibilities for the first digit and 10 possibilities for each
subsequent digit. There are exactly n - 1 subsequent digits. Thus, there are alto¬
gether 9 x 10” 1 combinations to form an n-digit number.
There is a large variety of possibilities for formulating similar analogous
tasks, such as using letters of the alphabet instead of numbers. The adoption of a
different point of view—that is, finding the quantity of multidigit numbers by
various ways—may increase the understanding of the solution.
Example 3
Figure 18.2
□ co oo irm mini :
, but not
Solution
Part 1: From the preconditions it follows that every piece must have a size
between 1 and 5 squares (that means each summand must be between 1 and 5). It
follows as well that there must be at least 3 pieces.
The only solution to separate the figure into 3 pieces is the one with 3 "fivers."
For separation into 4 pieces there are the following possibilities:
5 + 5 + 4 + 1
5 + 5 + 3 + 2
5 + 4 + 4 + 2
4 + 4 + 4 + 3.
□
Figure 18.4
a. all in a row
□ [
Figure 18.5
we get the same solutions as we got for 4 rectangles, but the strip of 5 is added.
For 6 rectangles with 1 through 6 units:
□ m
Figure 18.6
Using Game Strategies 395
a. all in a row
Example 4
The hero is sitting in the dungeon and is given the task of inventing a
new flag, which does not already exist, forthe ruler’s country. If he does
not offer at least x proposals, he shall lose his life. Can we help him?
The first principles for determining the scope of the solution are:
• organizing data
• approximating
• determining necessary and sufficient conditions
• determining characteristics of objects
the solution.
396 THE ART OF PROBLEM SOLVING
Solution
How many possibilities are there for combining 3 colors? There are exactly
6 possibilities for combining 3 different colors (A, B, and C) in the way required:
ABC, ACB, BAC, BCA, CAB, and CBA.
How many possibilities are there for selecting 3 different colors out of 6
different colors (A, B, C, D, E, and F)? ABC, ABD, ABE, ABF, ACD, ACE, ACF, ADE,
ADF, AEF, BCD, BCE, BCF, BDE, BDF, BEF, CDE, CDF, CEF, and DEF—20 possi¬
bilities. (In general, we can deduce that there are |n possibilities to choose k ele¬
ments, k<n, from n elements.) w
A different order creates a new possibility. Therefore, our hero can make 6 x
20 = 120 suggestions for a new flag. Of course, 5 of these are existing flags, leaving
115 possibilities. So, if the ruler demands not more than 115 suggestions for a flag
with the above conditions, our hero is saved. If he demands more, either the hero
will have to die or we will have to outsmart the ruler.
There are innumerable possibilities for the creative student and teacher, by
changing or extending the story.
Example 5
GAME
Let us imagine a chessboard, where there is a knight (N) on the bottom
COMB left square. How can this knight reach the top row from there without
NUMB
moving backwards? What are the possibilities?
Using Game Strategies 397
Figure 18.7
Again, the principles for finding the solution are to account systematically
for all possibilities and to solve it graphically, by making a picture or a table.
Solution
or 0.
The moves are classified as "short" or "long" only in the upward direction-
moves to the right or left are not treated separately. Therefore, there are only the
following possibilities for the knight in the left bottom comer to reach the top row:
(g + &) I
a\b\
All 21 different possibilities, according to the order and number of long and
short moves of the knight, lead right to the aim. The knight comes from the bottom
left comer to the top row of the chessboard.
The task can be variously extended or generalized, or analogous problems
can be created. For instance, all concrete possibilities of the positions of the knight's
moves could be found. One could also abandon the condition of not going
backwards. Then, with the restriction that no square may be touched twice, there
would be at most 63 possible moves for the knight.
It is really a demanding brain teaser for students interested in chess mathe¬
matics to exhaust this maximal number of moves and touch all squares at most
once.
To solve this task in general, one needs some knowledge of grid-point theory
(number theory—divisibility, discrete geometry—reflection). This task is also
called Euler's knight's-jump problem.
The interested teacher can find further chess-related mathematical problems
in Gik's (1986).
Outlook
creatively and playfully? Their joy in playing turned out to be the secret of their
success.
Thanks to innovative teaching and learning strategies, students can recog¬
nize that levity helps destroy the usual mental barriers. Playing and academic
performance are not exclusive of one another, as it is unfortunately often sup¬
posed. Students' joy in playful thought can be shaped in diversified and produc¬
tive ways. The playful wording of mathematical problems and the discovery of
mathematical principles in games can promote better concentration in students,
as well as long-lasting interests.
Children who are allowed to become absorbed in their playing are capable
of a better intellectual performance than children who are drilled and pressured.
Motivation by playing does not lessen the incentive and the desire to finish even
difficult tasks, but rather promotes these qualities.
The overlapping teaching strategies recommended in this chapter fall under
the pedagogical term "ludification." Ludification aims at the development, crea¬
tion, and improvement of the motivation for mathematical tasks by using game
elements and game strategies or, conversely, the discovery of mathematical prob¬
lems through games.
Only by succeeding in this can teachers lead their students into creative
mathematical problem solving.
References
HOPE J. HARTMAN
This chapter gives an overview of rationales for and methods of using coopera¬
tive learning to teach mathematical problem solving. Cooperative learning in¬
volves students working together in a teaching-learning situation. There are at
least three basic forms of cooperative learning: tutoring (peer or cross-age), in
which one student teaches another; pairs, who work and learn with each other;
and small groups of students teaching and learning together. First this chapter
focuses on teachers' and researchers' conclusions about cooperative learning as an
effective method of improving mathematical problem solving. Next it describes
several methods of using cooperative learning to teach mathematical problem
solving. Then the chapter concentrates on the teacher's role in cooperative learning
and concludes with some pitfalls involved in cooperative learning.
Sometimes the best way to learn something is to teach it. Teaching requires
considerable depth of knowledge, memory of important concepts, understanding.
401
402 THE ART OF PROBLEM SOLVING
the teacher can demystify the problem-solving process and reduce math anxiety.
Anxiety and misconceptions about mathematics can contribute to poor mathe¬
matics self-concepts and can inhibit success in mathematical problem solving
(Gourgey, 1992). Too often, students feel dependent on the teacher's expertise for
learning mathematics. In order to think mathematically—one component of which
is problem solving—students also need to develop their metacognitive skills of
planning, monitoring, and evaluating so they relate these skills to their problem
solving (Schoenfeld, 1989). Cooperative learning is an especially effective method
of activating metacognitive aspects of problem solving (Artzt & Armour-Thomas,
1992).
Studies comparing cooperative learning with competitive and individualis¬
tic learning have demonstrated that cooperative learning promotes higher
achievement in mathematics than the other two methods. Not only do students
solve math problems more successfully and learn and retain mathematical con¬
cepts, but cooperative learning also results in more use of higher-level thinking,
more frequent discovery, more new ideas and solution strategies, and more trans¬
fer of what is learned in groups about problem solving to individual problem¬
solving situations. These benefits are a result of student internalization of math
concepts and problem solving through their discussions and explanations of
problem-solving strategies and approaches with their peers (Johnson & Johnson,
1990). Giving mathematical explanations requires deeper understanding than just
putting an answer on a worksheet (Artzt & Newman, 1990).
There are numerous effective strategies students can learn (such as working
backwards and trying extreme values) to help them become effective problem
solvers. A major benefit of cooperative learning is that it allows students to
compare the strategies they use and discuss the advantages and disadvantages of
using them for different problems. When structuring groupwork around instruc¬
tion in such problem-solving techniques, make sure groups of students ask and
answer key questions about their strategy use: What is the strategy? Why is it a
good strategy to use in this situation? How will it be applied? What other strategies
could be used? What are the advantages and disadvantages of each? Through
dialogue with other students about the use of problem-solving strategies, miscon¬
ceptions can be clarified and strategy use improved.
Students can work cooperatively on mathematically significant, interesting,
and complex tasks. Through working with others they can enhance their ability to
communicate about mathematics, to understand it, and to think critically about it
(Good, Reys, Grouws, & Mulryan, 1989/1990). Discussing problems helps stu¬
dents become aware of what they know and do not know and what they under¬
stand and do not understand. This awareness leads, in turn, to control over the
problem-solving process as students begin to see the need for specific problem¬
solving strategies and the limitations of other approaches. When students find that
their approaches and outcomes for the same problem differ, this discrepancy can
stimulate reorganization and development of their thinking to a new and higher
level.
404 THE ART OF PROBLEM SOLVING
Scaffolding
Problem solving is usually much more fun and exciting to students when
they work together. The social context and active involvement make it more
intrinsically motivating to learn mathematics. Research has shown that coopera¬
tive learning increases confidence in students' mathematical abilities and im¬
proves self-esteem and feelings of self-efficacy in approaching math problems.
These effects increase the likelihood of students—especially for females and
minorities—choosing careers in math and science (Jackson, 1989). Other research
has documented the effects of cooperative learning on improving interpersonal
Cooperative Learning Approaches 405
write questions for a mathematics class. They learned how to use questioning to
help them plan, monitor, and evaluate problem solving. The teacher modeled
examples of questions by thinking aloud. Then students generated and used their
own questions. The three groups generated questions for this problem as shown
here:
Group 1
Student 1: What should the diagram look like? Where do the values belong?
What do I want to find?
Student 2: Where do I start? How do I find the desired answer? Where do the
numbers belong in the formula? Which number goes to which part?
Student 3: What does the diagram look like? What variables should I use?
Where does the 2 feet/second go? What derivatives do I have to find?
Student 4: How do I draw a picture to represent what the problem says? What
parts of the diagram get labeled? What is the unknown? What equation
do I use to get the derivative?
This group discussed their questions and made the following list to use when
solving the problem:
Group 2
Student 1: What does this look like? What formula do I use? How do I
approach it? What do I want to find?
Student 2: Are they on the water? How long is the rope? Is this a controlled
area with no waves and no current? How heavy is the boat?
Student 3: What do I have to find? What speed, velocity, and rate will help
me solve the problem? Why can't I figure this out? How do I differen¬
tiate the problem with respect to time?
Student 4: What am I given? What must I find?
After discussing their questions, this group made the following list:
Group 3
Student 1: What do I find? What does the diagram look like? How do I draw
this? Is it similar to something we've done? What is the equation? Is my
algebra correct? Am I using the right formula? Is the diagram correct?
Student 2: What type of diagram will this be? What are the dimensions of the
diagram? What is the rate of the rope being pulled in? What is the
problem looking for? What is given? What formula will be needed?
Student 3: Is the rate 2 feet/second horizontal or vertical? Where does the
equation come from? Is 10 feet used right now or after the derivative?
How is the Pythagorean theorem used if the hypotenuse isn't there?
1. What is given?
2. What should the diagram look like?
3. Is the diagram correct?
4. Is this similar to something we've done before?
5. What formula will be needed?
6. Is this the right formula?
7. What do we do with the speed and distance of the rope?
8. Is the algebra right?
While the groups worked on their questions and used them to solve the
problem, the teacher walked around to watch and listen to each group to make
sure they were on track and making reasonable progress. As she checked up on
each group she saw that some students still could not solve the problem. She
checked the individual and group lists of questions and saw that they were
incomplete, so she decided to have the groups share their questions, evaluate them
as a class, and come up with a composite list. She guided the discussion to make
sure the class generated questions for all three phases of the problem-solving
process (planning, monitoring, and evaluating). The following composite list
emerged:
Planning
Monitoring
1. Is my algebra correct?
2. Am I using the correct formula?
3. Is my diagram labeled correctly?
Evaluating
Students then returned to solving the problem with the new set of questions.
Individuals within each group shared their answers with each other, decided on
the correct answer, and raised their hands to let the teacher know when they were
finished so she could check their solutions. She randomly asked students in the
groups to explain their solutions to make sure everyone in the group understood
the problem and solution process. Then she had students who had solved the
problem help those who had difficulty. At the end of the lesson the class looked
at how the questions related to each part of the problem solution. Constructing,
comparing, discussing, and evaluating problem-solving questions individually, in
small groups, and with the entire class enriched students' understanding of what
questions and strategies were best suited for the particular problem. Some of the
students said that in the past, they had been so concerned with getting the right
answer that they had never given as much thought to the thinking process.
When each student has just his or her own knowledge, thoughts, and ques¬
tions, the perspective on problem solving is much more narrow and shallow.
Mathematicians frequently discuss their solution strategies and outcomes with
others. They know that others can sometimes detect limitations and suggest
alternative approaches to, and applications of, problem solutions. By discussing
problem solving with others, students learn to think more like mathematicians.
Encouragement, practice, and discussion are needed for students to learn to work
together successfully.
Three stages of instruction make up a Math Solution lesson. Lessons may
vary in length, lasting from a class period to a week or longer. The stages are
introduction, exploration, and summarizing. Whole-class instruction occurs dur¬
ing the introduction stage, when teachers (a) review or present concepts that are
needed, (b) pose a similar or smaller problem or part of the problem for students
to try, (c) present the problem to be solved, and (d) discuss to make sure students
understand what to do. Cooperative learning begins with the exploration stage.
As students work on the problem, the teacher (a) observes student interaction,
listening to each group's strategies, procedures, and ideas; (b) assists only when
needed (if hands are raised or the group is not working); and (c) provides an
extension activity for groups that finish early. The teacher may pose several prob¬
lems at once so groups can continue working at their own pace. The third stage,
summarizing, has three goals. First, groups share their problem-solving processes,
procedures, and strategies. Students critically evaluate their approaches and con¬
sider alternatives for future application. Second, groups present solutions, show¬
ing their work whenever possible. During this stage, it is recommended to ask,
How did you decide if your findings made sense? How can you check the solution
(Bums, 1990, p. 32)? Third, generalize from the solutions. During this stage,
students abstract what they did in solving a particular problem to think about how
they would approach related problems. Questions to ask include, Are there
patterns or relationships you can see from your solution? Does the problem remind
you of another problem you have solved? How are they alike or different (Bums,
1990, p. 32)?
Finding Out/Descubrimiento
Real Maths
The Real Maths curriculum (Terwel, 1990) for students ages 12 to 16 is based
on the theory of Van Hiele, an internationally known Dutch mathematics teacher
and researcher. Van Hiele's level theory identifies three levels:
Zero or perceptive level: Students look at the whole problem without ana¬
lyzing the parts.
First or descriptive level: Students describe the parts and their charac¬
teristics and have an intuitive understanding, but there is no reflection
on fundamental ideas.
Second or theoretical level: Students' intuitive concepts are formulated
more explicitly, and students reflect on concepts and on the relation¬
ships between problem parts and the whole.
1. Introduction: Working with the whole class, the teacher introduces the
problem, explores aspects of it, and may give hints about the solution and
place it in an everyday-life context.
2. Group work: Students work in groups as the teacher observes and man¬
ages their cooperative problem solving. When necessary, the teacher deals
with individual problems.
3. Reflection and evaluation: Students discuss several topics related to their
group process and results. The discussion involves identifying all the
different solutions and strategies groups used to solve the problem,
questioning (by the teacher) to explore other possible solution strate¬
gies, reformulating and summarizing solutions, and generalizing about
solutions.
Sample lessons involve using newspaper ads, deciding which video rental
shop to become a member of, and selling badges as a small business enterprise.
The Real Maths curriculum has accompanying teacher's manuals as resources for
using real-life situations for mathematical problems and for teachmg with small,
heterogeneous groups. It also has student materials and videotapes.
students process a lot of the paperwork that ties a teacher down. Students check
each other's work as they progress through the units,... [providing the] immedi¬
ate feedback students need [as] it identifies problems that often can be handled in
the group or answered by the teacher if further help is needed" (Slavin, 1990, p. 81).
A fifth-grade teacher reported that individual accountability and group rewards
are especially effective for students who dislike mathematics. TAI enables students
to work at their own level and achieve success, thereby eliminating frustration and
boredom. Student success is translated into team points, which motivates students
to do well. The individual student gets self-satisfaction; the group gets a reward
and positive reinforcement from the teacher. A high school special-education
teacher who used TAI noted that working toward a team score led students to
encourage each other to complete class assignments and resulted in a faster rate
of mastery than did individualized mathematics instruction. Mastery of 80% is
required to move to the next level, so the quality of work remains high.
and between steps. This method can be used as a tutoring technique in a peer or
cross-age tutoring program or for classwide peer tutoring in the regular classroom.
Pairs of students can take turns solving problems, either alternating roles as tutor
and student or with stronger students tutoring weaker students. The entire class
gets trained in this learning method so that all understand what happens and why.
This method not only can help students become active learners and independent
problem solvers, it also develops students' questioning, communication, and
leadership skills while simultaneously building their self-esteem and mathematics
self-concepts. The six steps to use for problem solving include
4. Prod: If the student still has not solved the problem, ask him or her to make
an educated guess about how to solve it, or ask how the student feels
(rather than thinks) it might be done. If the students solves the problem,
go to step 6.
5. Prompt: If the student is stuck and all prior attempts have failed, give a
hint or suggestion, or show the student how to start the process. If this
similar problem.
The tutor should paraphrase frequently but not continuously. At the end of
questioning, ask the student to summarize what was learned. See the dialogue
T = tutor, S = student
1. Preface
problems.
T: So you want me to help you with a specific problem? [paraphrase]
S: Yeah, I figured out some of it, but now I'm stuck.
T: What problem are you working on?
S: A factoring problem, number 4: Factor 6x3 + I2x1 + 6x.
T: So you only need help solving part of the problem? [paraphrase]
2. Pace
3. Probe
T: Can you tell me more about the trinomial you have left, x2 + 2x + 1?
S: I remember something about two sets of parentheses, but I can't remember
what numbers to put in them.
T: Do you know the name for the trinomial?
S: What do you mean by a name?
T: There's a specific equation the trinomial is associated with.
S: Is that the quadratic equation?
T: That's right. What else can you tell me?
S: Nothing. Nothing at all.
T: OK, so you know that you arrange numbers in two sets of parentheses
and that somehow this is connected with the quadratic equation, but
that's it. [paraphrase]
S: Yes.
T: How are the parentheses set up?
S: They are next to each other.
T: And what does that mean?
S: That you multiply what's inside each of the parentheses. Oh! So I can put
an x inside each parenthesis and multiply them to get x2. Is that right?
T: Yes, good! When you multiply the xs in the parentheses, it results in the
first term of the quadratic equation.
S: I guess I knew more than I thought!
4. Prod
T: OK, now look at the problem again and make an educated guess about
how to fill in the rest of the quantities.
Cooperative Learning Approaches 415
S: Hum ... I'm getting a little clearer about it. I've got to put the numbers
and xs into the parentheses so that when they are multiplied together I
get the trinomial x2 + 2x + 1. But I don't know what to do to get it what
else to put where.
5. Prompt
T: I'll give you a little hint. Remember what function you used to get the
x2 from the xs in the parentheses?
S: I multiplied.
T: Right, so what function will you use to get the 2x + 1?
S: I guess I have to multiply to get the last term—oh yeah, and then I add
to get the one in the middle! I have to find numbers that when multi¬
plied you get one and when added you get 2x. So what would that be?
Let's see ... one times one equals one, and one plus one equals two.
That must be it! Is it?
T: Sounds like you got it! So what does it all look like?
S: It must be 6x(x + l)(x + 1).
T: Great, that's it! But can you take it one step more? Isn't there a simpler
way of representing the information in the two sets of parentheses?
S: Oh, I see, (x + 1) (x + 1) can be changed to (x + l)2. So that means it
ends up as 6x(x + l)2. Is that it?
T: Alright, now you got it!
6. Process
T: How can you use what you learned to solve other problems?
S: Well I was stuck on homework problems 7, 24, and 30, because I think
they are similar problems that I got stuck on the same general way.
T: Yes, they're similar. So you think that you can solve problems 7, 24, and
30 now based on what you have learned? [paraphrase]
S: I think so. Thanks for the help!
T: Will you summarize what you learned before you try those other
problems?
Summary
S: I started with the problem: Factor 6x3 + 12x2 + 6x. I saw that I could pull
out the 6 as a common factor, and that I could also pull out the x, but
then I got stuck. Now I know to arrange the numbers in parentheses
side by side so when they're multiplied they give me the end terms m
the trinomial and when added they give me the middle term.
In this method, telling or showing how to solve the problem is only used as
a last resort. Through 6PQ, students learn that they know more than they think
they do, which helps build their confidence and willingness to persist.
416 THE ART OF PROBLEM SOLVING
Thinker and listener pairs working on problems and rotating roles has
become a popular way of helping students think about their own problem solving
(Whimbey & Lochhead, 1982). Students take turns serving as thinkers (problem
solvers) who externalize their thought processes by thinking aloud, while analyti¬
cal listeners track and guide the problem-solving process as needed. It is a higher-
level thinking (metacognitive), self-monitoring strategy that gives students feed¬
back on what is understood and what is still unclear. It helps students identify
what parts of a problem they understand and where they get stuck. This method
makes problems more engaging, teaches communication skills, and fosters coop¬
eration. It encourages skills of reflecting on beginning and later thoughts. Pair
problem solving facilitates self-checking and evaluating. It encourages the forma¬
tion of study and support groups and exposes teachers and students to various so¬
lution approaches (Heiman, Narode, Slomianko, & Lochhead, 1987a). By listening
to his or her own thoughts, the student gains awareness and control over problem
solving. Externalizing thoughts enables them to be seen from a fresh perspective.
1. Translate your thoughts (ideas, images, etc.) into words and recite them
aloud.
2. Verbalize aloud all the steps you go through when solving problems. No
thought or step is too small, easy, obvious, or unimportant to verbalize.
3. Verbalize all the thinking you do before you start to solve the problem (e.g.,
what you are going to do, when, why, and how). Even second-guessing
yourself is important to verbalize aloud—for example, "l think I should
use that long, complicated formula we were using a couple weeks ago.
What was it called, the quadratic equation? No, maybe not. Maybe I'm
supposed to use the formula we did in class yesterday."
1. Think along with the problem solver. Follow every step and make sure
you understand every step. If not, ask a question. Have the problem solver
identify and define important terms, variables, rules, procedures, and so
forth. Make sure the problem solver vocalizes all the steps and does all
the work. If the problem solver skips over a step without thinking aloud,
ask her or him to explain the missing thought.
2. Do not work on the problem independently. Listen to and work along with
the problem solver.
Cooperative Learning Approaches 417
3. Never let the problem solver get ahead of you. Whenever necessary, ask
the problem solver to wait so you can check a procedure or computation
and catch up. If the problem solver is working too fast, slow her or him
down so you can follow carefully, analytically, and accurately.
4. Check the problem solver at every step. Don't wait for the answer. Check
everything—each computation, diagram, procedure, and so forth. In the
back of your mind, constantly ask yourself, "Is that right? Did I check
that?"
5. If you find an error, avoid correcting it. Point it out and try to get the
problem solver to self-correct. If he or she gets stuck, ask questions to
guide thinking in the right direction. If necessary, give suggestions, hints,
or partial answers. Give the answer as a last resort. Let the problem solver
know that you are not trying to be difficult but are trying to help him or
her become an independent problem solver (Whimbey & Lochhead, 1982).
To make sure listeners really do their job, teachers should periodically ask
listeners to summarize the steps the problem solvers used. The following activities
were suggested by Larcombe to help student pairs in elementary grades or
remedial high school students learn to externalize their mathematical thought
processes:
I DREAM of A
having the problem solver think aloud periodically. The problem solver answers
questions and thinks aloud, as prompted by the questioner. Questions are asked
for each of seven components I DREAM of A. Although most questions focus on
knowledge and strategies needed to solve a problem, the questioner occasionally
asks about the problem solver's feelings to establish and maintain a positive
attitude. The questioner decides what questions to ask, when to ask the problem
solver to think aloud, and when to ask about the problem solver's attitudes. An
illustration of the model follows.
I • Identify
What is given/to be found? How could you state the problem in your own
words? What do you know about this type of problem? Does anything seem
confusing? What information is relevant? What isn't? How do you feel about your
ability to solve this problem?
D • Diagram
How would you sketch or draw a picture of that problem? What symbols or
notation should be in your diagram? Are you sure you have all the important
parts? Are you sure they are in accurate relationship to each other? How does your
diagram help you identify and define the problem?
R • Recall
Have you solved problems like this before? How? Do you have an example
of how to solve this type of problem in your text or notes? What definitions, rules,
concepts, procedures, equations, and such might you need to solve this problem?
Why? When would you use them? Is there anything you should review? Are there
any mistakes you tend to make on these problems?
E • Explore
Think out loud about how you might solve this problem. What approaches
could you use? Are there any general strategies, such as working backwards or
trying extreme values, that might be useful? Which is best? Why? What would
you estimate the answer to be? What difficulties might arise while solving the
problem? How could you break the problem into parts? What would you do first,
second,... ?
A • Apply
What is your plan for solving this problem? Think out loud as you apply your
problem-solving plan.
Cooperative Learning Approaches 419
M • Monitor
How are you doing so far? Is your approach leading you where you need to
go? How can you tell? Should you try another approach? How have you been
checking to make sure you haven't forgotten anything and haven't made any
careless mistakes? Are you keeping a positive outlook? Does everything make
sense?
A • Assess
Are you sure you have answered the problem completely? Looking back at
the problem statement, does you answer make sense? How can you verify your
answer? What did you learn from solving this problem that could help you in the
future?
In a peer or cross-age tutoring situation, the tutor is the questioner and the
person being tutored is the problem solver. In pair learning or cooperative learn¬
ing, students take turns performing roles of questioner and problem solver. To
become independent problem solvers, students can learn to ask themselves such
questions when working on their own and internalize the components of the
problem-solving process. The sample questions are presented to stimulate think¬
ing about appropriate types of questions to ask. Each situation will require some¬
what different questions. Research shows that student-generated self-questions
are more effective than teacher-imposed questions. Teacher-imposed questions
are, however, good building blocks or scaffolds which temporarily guide students
through problem solving until students can independently self-question, maintain
a positive attitude, and guide their own problem solving.
One of the key features of cooperative learning is that the teacher delegates
considerable responsibility to the students. Both teachers and students must adopt
roles that are quite different from whole-class instruction. "The essence of good
problem solving is self-correction. . . . Teachers should become facilitators of
learning, not sole dispensers of truth, [and] must relinquish the safe seat of
authority and step into the classroom ..(Heiman et al., 1987b, p. 8).
An important task in making cooperative learning successful is attitude
change. The teacher's own attitude has to shift from teacher as transmitter of
knowledge, center of attention, and authority to teacher as manager and facilitator
of learning Attitudes toward noise in the classroom sometimes must be changed.
Many people equate a noisy classroom with disruption and chaos. In cooperative
learning noise in the classroom can be a sign of high-level thinking and learning.
Several student attitudes must change so students show interest in finding solu-
420 THE ART OF PROBLEM SOLVING
Erickson (1989) devised a list of nine steps for teachers to follow in imple¬
menting cooperative learning:
Johnson and Johnson (1983) are well known for their contributions to coop¬
erative learning. They emphasizes face-to-face interaction as students work in
heterogeneous groups with individual accountability on tasks that require posi¬
tive interdependence. Their guidelines for training students for cooperation are:
Cooperative Learning Approaches 421
Teacher modeling of social skills and role playing are effective strategies for
training students for cooperative learning. Johnson and Johnson (1983) identify
four categories of skills to be developed for cooperative learning:
Forming Skills
Functioning Skills
• using names
• looking at the speaker
• being respectful of others
• expressing support
• asking for help or clarification
Formulating Skills
• seeking elaboration
• seeking clever ways of remembering information
• demanding vocalization
• asking other members to plan out loud
422 THE ART OF PROBLEM SOLVING
Fermenting Skills
Structuring Groups
group is large enough for generating ideas and discussing solutions of challenging
problems, is small enough for all students to participate; can be conducted without
a leader, and can be split up into pairs for occasional practice (Davidson, 1990).
Groups may stay together for a single class period, a week, a school term, or the
entire school year. Many proponents of cooperative learning recommend chang¬
ing groups often enough that eventually all students in the class have the oppor¬
tunity to work together. Some teachers allow students to choose their own groups;
many assign students to groups to ensure heterogeneity. Assignment can be done
on the basis of students' achievement records, through random grouping, or
according to the regular seating chart. Substantial research supports the value of
gender, racial, and ethnic diversity in cooperative-learning groups.
Deciding how to arrange groups in the classroom is another decision teachers
have to make. Movable desks or tables and chairs are important but not absolutely
necessary; with them, groups can be arranged so as to facilitate communication
without disturbing other groups. Pairs can easily be used even in large lecture halls
with no movable furniture or room to sit on the floor.
There are numerous ways to use cooperative learning for teaching problem
solving. Some methods are relatively informal in structure but involve students
working together cooperatively to achieve a problem-solving goal. Other methods
are highly structured in terms of student roles, instructional techniques, materials,
and assessment strategies. Well-structured cooperative-learning groups can in¬
crease the chances that all students will examine alternative solution strategies,
will observe peers engaged in problem solving, and will formulate, analyze, and
interpret problems and solutions. Positive interdependence and face-to-face inter¬
action are guiding features behind some approaches to cooperative leammg m
mathematics. These are facilitated through assigning distinct roles to perform
Teachers sometimes have each group member fulfill a distinct role to ensure
that all students participate in the process and to achieve a variety of goals.
However, specific roles are not necessary. For example, the Small-Group Discov¬
ery Method of using cooperative learning in mathematics does not involve student
roles. Some teachers prefer to assign students to their roles; other teachers prefer
students to choose their own roles in the groups. Students should understand that
their participation in the group is not limited to the roles they perform. Their roles
are just one aspect of their participation in the group's problem-solving activities.
Students need to see and be part of the "big picture" of problem solvmg. Many
teachers have students change roles after one class period or after one group
project. Regardless of which method is used, teachers should make sure that
eventually all students get a chance to function in each of the group roles. There
are several models of roles that can be use for problem-solving groups.
424 THE ART OF PROBLEM SOLVING
Model 1
• Approximator: Gets the group to estimate what the answer will be before
they begin actually solving the problem.
• Reviewer/mistake manager: Has the group figure out how they can learn
from whatever mistakes may have been made and, when the group is
successful, has the group determine how their solution could be even
better next time.
• Confidence builder: Encourages the group to keep going because they will
succeed if they persist and work together effectively. (Johnson & Johnson,
1990)
Model 2
• Idea person: Gives mathematical ideas to the group, such as how to solve
the problem.
• Ego builder: Builds pride in the group by praising its members. (Crabill
1990)
Model 3
Model 4
• Checker for accuracy and direction: Checks for mistakes, makes sure the
approach they are using is leading in the right direction.
• Summarizer: Reviews progress the group has made and what still needs
to be done.
Model 5
There at least two different types of group roles that can be assigned:
management and instructional roles. Examples of management roles include
facilitator and recorder; examples of instructional roles include chief investigator
and adviser. Some controversy exists about the value and effectiveness of different
types of roles. Management roles have been found to be especially important for
younger children because they help prevent confusion, wasted time, and conflicts,
while instructional roles sometimes create problems for both older and younger
children (Good et al., 1989/1990).
Although some teachers find they can adapt their existing mathematics
curriculum for cooperative learning, others find that more appropriate materials
are needed. There are several resources specially developed for implementing
cooperative learning in mathematics. One especially good resource is Get It To¬
gether: Mathematics Problems for Groups, Grades 4-12 (Erickson, 1989), developed by
Project Equals at Lawrence Hall of Science, University of California, Berkeley.
The NCTM handbook on cooperative learning is a clear and concise guide
with carefully constructed problems, clear instructional objectives, problem
sheets, activities, and teaching notes (Artzt & Newman, 1990). Others include the
Real Maths curricular materials (Terwel, 1990) in the Netherlands, the Finding
Out/Descubrimeinto curriculum (Lotan & Benton, 1990) from the Stanford Uni¬
versity, California, and Developmental Mathematics (Hackworth, Howland, & Al-
win, 1989). The final chapter of the outstanding book by Davidson (1990), Coopera¬
tive Learning in Mathematics, identifies additional resource materials. Designing
Groupwork (Cohen, 1984) contains excellent descriptions of several cooperative-
learning methods and especially useful materials and procedures for training
426 THE ART OF PROBLEM SOLVING
students for cooperation. There are other good sources, too numerous to mention,
on using cooperative learning to teach mathematical problem solving.
than as a requirement.
Another problem is that one or two students can do all the work solving
problems while the others do not. To prevent this, individual accountability is
essential. Groups must be structured to foster cooperation among students. As¬
signing roles, sharing materials, requiring a group product, and using group
incentives can help structure effective cooperation.
Time can be a problem when implementing cooperative learning, and lessons
sometimes end without summarizing what was learned and assessing the group
process. One way of handling this problem is to assign roles of summarizer and
leader of group processing. Otherwise the teacher can lead the summarizing and
group processing at the end of each lesson (or set of lessons).
Cooperative learning in mathematics can lead to incoherent presentation and
interruptions while working so that students need to spend more time reviewing
and practicing. If initial training of students to work cooperatively is not adequate
for some students or groups, follow-up training may be needed. Sometimes
problems arise if teachers set only academic goals rather than specifying both
academic and social goals in advance (Good et al., 1989/1990).
Conclusion
Note
References
Aronson, E., Blaney, N. Stephan, C., Sikes, J., & Snapp, M. (1978). The jigsaw
classroom. Beverly Hills, CA: Sage.
Artzt, A., & Armour-Thomas, E. (1992). Development of a cognitive-metacognitive
framework for protocol analysis of group problem solving in mathematics.
Cognition and Instruction, 9(2), 137-175.
Artzt, A., & Newman, C. (1992). How to use cooperative learning in the mathematics
class. Reston, VA: National Council of Teachers of Mathematics.
Bransford, J., & Stein, B. (1984). The IDEAL problem solver. New York: Freeman.
Bums, M. (1990). The math solution: Using groups of four. In N. Davidson (Ed.),
Cooperative learning in mathematics. Tucson, AZ: Zephyr.
Chemeketa Community College. (1986). The 6PQ method of discovery learning.
Salem, OR: Author.
Erickson, T. (1989). Get it together: Mathematics problems for groups grades 4-12.
Tucson, AZ: Zephyr.
Fantuzzo, J., Alperin King, J., Rio Heller, L. (1992). Effects of reciprocal peer
tutoring on mathematics and school adjustment. Journal of Educational Psychol¬
ogy, 84(3), 331-339.
Frankenstein, M. (1986, Spring/Summer). Ideas for teaching a non-rote college
arithmetic course. Mathematics in College, pp. 22-31.
Codperative Learning Approaches 429
Good, T., Reys, B., Grouws, D., & Mulryan, C. (1989/1990). Using work-groups in
mathematics instruction. Educational Leadership, 47(4), 56-62.
Gourgey, A. (1992). Tutoring developmental mathematics: Overcoming anxiety
and fostering independent learning. Journal of Developmental Education, 15(3),
10-14.
Hackworth, R., Howland, J., & Alwin, R. (1989). Developmental mathematics. Clear¬
water, FL: H & H Publishing.
Hartman, H. (1992). Intelligent tutoring. Clearwater, FL: H & H Publishing.
Heiman, M., Narode, R., Slomianko, J., & Lochhead, J. (1987a). Teaching thinking
skills: Mathematics. Washington, DC: National Education Association.
Heiman, M., Narode, R., Slomianko, J., & Lochhead, J. (1987b). Teaching thinking
skills: Science. Washington, DC: National Education Association.
Johnson, D., & Johnson, R. (1975). Learning together and alone. Englewood Cliffs, NJ:
Prentice Hall.
Johnson, D., & Johnson, R. (1983). Overview of cooperative learning. Minneapolis:
University of Minnesota.
Johnson, D., & Johnson, R. (1990). Using cooperative learning in mathematics. In
N. Davidson (Ed.), Cooperative learning in mathematics. Tucson, AZ: Zephyr.
Lotan, R., & Benton, J. (1990). Finding out about complex instruction: Teaching
math and science in heterogeneous classrooms. In N. Davidson (Ed.), Coopera¬
tive learning in mathematics. Tucson, AZ: Zephyr.
Pressley, M., Wood, E., Woloshuyn, V., Martin, V., King, A., & Menke, D. (1992).
Encouraging mindful use of prior knowledge: Attempting to construct ex¬
planatory answers facilitates learning. Educational Psychologist, 27(1), 91-109.
Rosenshine, B„ & Meister, C. (1992, April). The use of scaffolds for teaching
higher-level cognitive strategies. Educational Leadership, pp. 26-33.
Schoenfeld, A. (1989). Teaching mathematical thinking and problem solving. In
Toward a thinking curriculum: Current cognitive research. Alexandria, VA: Asso¬
ciation for Supervision and Curriculum Development Yearbook.
Slavin, R. (1990). Cooperative learning research, theory and practice. Englewood Cliffs,
Plenum.
Whimbey, A., & Lochhead, J. (1982). Problem solving and comprehension. Philadel¬
Recommended Reading
Baron, J., & Sternberg. (Eds.). (1987). Teaching thinking skills. New York: Freeman.
Chipman, S., Segal, J., & Glaser, R. (Eds.). (1985). Thinking and learning skills.
Hillsdale, NJ: Lawrence Erlbaum.
Wertheimer, M. (1959). Productive thinking (Enlarged ed.). New York: Harper &
Row.
20
Problem Solving and the
Mathematically Gifted Student
A Psychological Perspective
BRIGITTE A. ROLLETT
In spite of the fact that the development of higher mathematical ability has now
been studied intensively for almost a century, there is still a broad range of
intriguing questions that have yet to be investigated. Foremost amongst these are
the causes for the highly gifted individuals' exceptionally efficient ways of pro¬
cessing information. Of special importance is the neurophysiological basis of
cognitive and personality development and the interaction of the related factors.
Some insight into problem solving can be obtained from an understanding of these
phenomena.
Defining Giftedness
431
432 THE ART OF PROBLEM SOLVING
Children who are later identified as gifted often show interest in their
environment at a very early age. Even as newborn infants, they are "efficient
orienters, as Stapf and Stapf (1988) established in an investigation of 51 gifted
children. The orienting reaction is an indicator of the child's readiness to learn from
his or her environment. This active information-seeking behavior is optimal in
furthering the development of the functional units of the brain. One of the basic
observations made when comparing efficient learners with their less able peers is
Mathematically Gifted Students 433
that the former seem to be better able to identify relevant information from
unimportant data and to concentrate on processing it.
We therefore want to define giftedness as the ability expertly to identify,
construct, process, and handle relevant information in one or more culturally
defined fields of action. Renzulli's classic model of giftedness further elaborates
this by drawing attention to the role creativity (the ability to produce relevant new
ideas) and task commitment (willingness to work toward high accomplishment)
play in attaining superior results.
It is not surprising that novelty preference is a better predictor of cognitive
development than the parents' educational level, as Rose and Wallace (1985)
showed in a longitudinal study of the cognitive functioning of preschool children.
Highly gifted children need less time to react to and process information. As
most functions of the cerebral cortex only develop after birth in response to
stimulation by the environment, this interest in new information brings about a
significant leap forward in the child's development, one which can, of course, be
increased if the child's persons of reference know how to play with him or her in
a stimulating way and to encourage it. It is typical of mathematical giftedness to
appear, together with high verbal proficiency, at a very early age (3 to 4 years).
Early mastery of the symbolic nature of language seems to be a milestone in intel¬
lectual giftedness. Another is the early specialization of the functions of the brain.
Degree of Lateralization of
Cerebral Functions and Giftedness
The general interpretive area, or "Wernicke's area," in the brain plays the
most important role in the development of language, including dealing with
numbers and the handling of meaningful information in general. Consciousness
and self-awareness are mediated through this area. Thus, it represents the core of
the personality and its development.
As one hemisphere—usually the left one—is principally used when process¬
ing new information, it gains dominance over the other hemisphere, becoming
more and more expert in dealing with new learning experiences and their results.
This process of lateralization leads to an efficient division of labor between the two
hemispheres, as Nobel laureate Sperry (1974) showed: With most people, the left
hemisphere becomes devoted to the processing of language and numbers and,
later, the development of the writing and reading centers, while the right hemi¬
sphere is dedicated to spatial orientation and construction, the processing of
pictorial and color information, music, and the understanding and interpreting of
higher-order emotions. From this description it is clear that both hemispheres have
to work together when solving mathematical problems, as Levin and Spiers (1985)
in their study of acalculia showed. While the left hemisphere handles the numeri¬
cal or digital information to be computed, the right one processes the schematic,
Creativity, the ability to make unusual associations of ideas and find uncom¬
mon solutions, presents a special problem (see Urban, 1992). It is central to
extraordinary achievement in many fields and to success in later life but is rarely
stimulated by parents and educators. These people have to be creative themselves
in order to be able to recognize creative attainments as such. Parents and educators
also need to react in a suitable way, that is, by helping to develop the idea in a
playful manner or simply by acknowledging its correctness, as Gauss's teacher did
when the young Gauss presented him with the ingenious solution for the problem
of adding all the numbers from 1 to 100 by multiplying 101 by 50, thus intuitively
deducing the principle of the arithmetical progression.
At the age of about 4, most children make a "creativity jump." At this age,
language, especially, develops; but also the imagination begins, and with it, joy in
new, surprising associations of thoughts and words, new word creations, and cre¬
ative, imaginative games, including games with numbers. Depending on whether
this development is fostered by the environment or dismissed as "nonsense," it
can grow into a stable feature of character or merely atrophy.
Recently, there has been much discussion whether or not to accelerate the
gifted. Renzulli and McGreevy (1986) investigated 23 pairs of twins, of whom only
one twin in each case had been diagnosed as being gifted and admitted to a special
training program. It was seen that the other twin later indeed displayed a poorer
academic performance but was much more creative and independent in thought.
In the course of their study of mathematically precocious youths, Swiatek
and Benbow (1991) compared accelerated gifted mathematics students with others
who were not accelerated, following their academic and psychosocial develop¬
ment during a 10-year period. As was to be expected, the accelerated group
finished college 1 year earlier. Both groups showed outstanding academic and
personal satisfaction at age 23. The authors concluded that the accelerates had only
a slight advantage over the unaccelerated. However, an analysis of their data in
terms of effect sizes shows a clear shift of the accelerates toward academic prefer¬
ences, while the unaccelerated group had more diverse interests and developed
higher self-esteem, presumably because they were not continually exposed to
academic peer pressure. Whether one advocates acceleration of the gifted or
simply provides them with an enriched curriculum depends on the educational
Mathematically Gifted Students 435
goals one prefers. One such form of enrichment would be the integrated inclusion
of a problem-solving strand in the syllabus.
Developing Metacognition
Metacognition has been a buzzword now for some time. As a general com¬
petency it begins to appear at about 8 years of age, accelerating its development
at about 11, when, according to Piaget, formal logic and abstract thinking develop.
Gifted children show precursors of metacognitive thinking at a very early age.
Shore and Dover (1987) have been able to demonstrate that there are charac¬
teristic differences between gifted and normal children with regard to the use of
metacognitive strategies. Metacognition refers to thinking about thinking, that is,
rational deliberations about the best strategic employment of one's own compe¬
tences and strategies to solve problems. Gifted children manage to guide and thus
optimize their handling of problems through metacognition earlier and more
systematically.
The way parents and teachers treat the child when solving problems together
has proved to be of special significance to the development of cognitive compe¬
tences. Parental interaction with the child when dealing with problems is the
scaffolding for the development of intelligent strategies to solve problems. In the
course of dealing with problems in conjunction with parents, the child experiences
more and more success in constructing his or her own solutions, thus developing
his or her intellectual and metacognitive competences.
The difference in information processing among gifted, normal, and slow
students is well illustrated by a study conducted by Lehtinen (1992, p. 127), who
asked students to summarize a mathematical lesson dealing with changing frac¬
tions into decimals, writing down what was most important.
A gifted girl wrote: "During this lesson we were taught many important
things about turning a fraction into decimals, where the denominator is 10,100 or
2 21
1000, e.g., — — 0.2 and = .21, and the multiples of a fraction, whose denominator
is not 10,100, or 1000, and their transformation into a decimal ee — = — = n 16-
25 100 '
l = Ji = 0A- 3I = 3^ = 3-4-
Mathematically Gifted Students 437
This girl knew how to identify the relevant core information of the lesson and
construct a well-ordered information module to be remembered, thus augmenting
her own internal knowledge base, by illustrating the procedures by a few well-
chosen examples.
On the other hand, one average student, wrote: "During this lesson that
5 25
[which was important] was taking multiples, e.g., - = — = 2.5. This transforming
into a decimal was the most important, e.g., = 0.50, and telling about it was the
most important."
While this description does not include any false information, it is certainly
not complete and it lacks structure.
The third example was produced by a willing but very slow learner "This
lesson. During this lesson one was supposed to listen and ask questions. — — 0.2.
Without instruction I would not have known this. (On the blackboard, the teacher
had written = O.2.")
During the lesson, this student had willingly played the "question-and-
answer game," but he was unable to correctly reconstruct the content of the lesson
for himself. He would have needed far more individual encouragement and
information feedback to arrive at a workable knowledge base enabling him to keep
pace with his peers.
mathematics.
When working on a problem, gifted students resort to a procedure closely
3. Carry it out.
4. Most important, look back and check the results not only for corrections,
but also for determining possible further use (Polya, 1988, pp. xvi-xvii).
Ill
1+2+4+8+
It is understood that this sum is the limit of the series of partial sums.
We then get the formula for the limit = lim S„ = a/(l - r), where a is
the first term and r the common ratio.
It is not always easy for students to see the connections and recognize
the fact that the sum of this infinite series can be set equal to a particular
value. In the above case, the result isl+- + - + ... = 2.
2 4
One 16-year-old student's argument at the time with the help of a
number line
Mathematically Gifted Students 439
S1 S2 S3 s4
H-1-1—I—I-
0 1 2
Figure 20.1
It is important to note that this was a spontaneous creative insight into the
problem by this student and not a prepared part of the teacher's lesson.
Here is another instance of the way gifted students reformulate a problem struc¬
ture. Eva Turner gave the following problem to her lOth-grade mathematics class:
u\ - 3ai = 24
/n = fV3
1 fli
n=3hl=2' 3
A2 : fli = n : (hi - n)
a\ n
ai = -.-
hi - n
Figure 20.2
U2 = 3fl2 = 12
In this case, the student realized that he could use the tangent function he had
previously learned to arrive at a more elegant solution:
a\ = 8 cm => u\ = 24 cm
3ri
a2 = a2 = 4cm => u2 = 12 cm
As the teacher explained: "Normally, students at age 15 have not learned the
tangent function and therefore do not have it at their disposal. However, this year
I chose to begin with trigonometry and then did the required work in the proper
order" (Turner, personal communication, 1993).
It is typical for gifted students to integrate new information into their existing
knowledge base, forming a new gestalt that makes it possible to see how different
modules of information relate to each other and can be used to advantage. Average
students tend to "compartmentalize" what they know. It is therefore a good
teaching strategy to work against this tendency, to let students deduce relations
between different modules of information, and to show them how they can make
use of them.
Intelligent students often try to solve problems according to a "least-effort
principle, finding shortcuts and producing elegant solutions. Another example
Eva Turner relates begins with this problem:
a2 - d, a2, a2 + d
so that
and
3 fl2 = 24
fl2 = 8
^2 — 2^2^ + d + #2 ± ^2 ± ^ — 642
2d2=450
d = ±15,
find the fraction midway between two fractions when, to his conster¬
nation, McKay, one of the students, interrupted, "Sir, you don't have to
go to all that trouble to find a fraction between two fractions, all you
have to do is add the tops and bottoms."
Of course, McKay was right:
If 0 < b<~d' t^len ^ < < ~d’ is McKay's theorem, (pp. 294-295)
The ease with which gifted students can find new and labor-saving proce¬
dures may, however, sometimes turn against them, when they shirk the trouble
of actually working through the problem. The following story by high school
mathematics teacher Richard Henner elucidates this problem:
After being given the problem to develop the equation and use appropriate
methods to answer the various questions, all the students very diligently com¬
pleted the work. All students saved the most time-consuming work, drawing the
equation, for the end—except for Johanna, a very gifted and ambitious, though not
always very diligent, girl. She began to solve the problem by using the graph,
which in large measure she did incorrectly. Because she did not have the equation
of the function at her disposal, she had only the half-life values to use in graphing
the curve. In the process, she noticed something I had overlooked in making the
assignment coincidentally a triple half-life had passed since the date of the
accident. Therefore, a good part of the assignment was, for her, simply mental
arithmatic. One eighth of a substance still exists, and therefore 3/8 still needs to
be disposed of.
In this case, an "effort avoidance motive" (Rollett, 1987) led the student to
find this elegant solution. This motive, however, can prove to be a double-edged
sword, as her teacher comments:
Mathematically Gifted Students 443
As the story of Fermat's (1601-1665) famous "last theorem shows, this reluctance
of brilliant thinkers to work out their solutions in detail may have delayed
discovery of the proof about 350 years.
The conclusion can only be that giftedness per se does not guarantee success,
but needs skillful mentoring, including praise not only for brilliant ideas but also
for diligently realizing them. Also crucial is a supporting environment that ensures
that mathematical problem solving remains a rewarding and intellectually stimu¬
lating experience.
Since enhancing individuals' mathematical knowledge base plays such an
important role in ensuring progress, students should be encouraged to learn not
simply toward the goal of passing exams, but to strive to augment their own
repertoire of strategies and insights and, to use Schoenfeld's terms, resources and
heuristics—in short, learn to learn for future use.
The teacher should not insist on rote learning and standard procedures but
allow the students to find new solutions, if this is feasible, and teach them how to
control what they have achieved, actively using metacognitive techniques.
Another important characteristic of expert mathematicians that may lead to
a useful strategy for efficient math instruction is their not being afraid of making
mistakes. As Sawyer (1943) admonishes his readers, "The important thing is to
learn how to strike out for yourself. Any mistakes you make can be corrected later.
If you start by trying to be perfect, you will get nowhere. The road to perfection is
by way of making mistakes" (p. 24). Only those who are willing to risk trymg a
seemingly useful procedure, but are aware of the possible need to change their
approach, can become active, self-reliant explorers in the fascinating field of
mathematics.
References
Polya, G. (1988). How to solve it: A new aspect of mathematical method. Princeton, NJ:
Princeton University Press.
Renzulli, J. S. (1980). What we don't know about programming for the gifted and
talented. PhiDelta Kappan, 61, 601-602.
Renzulli, J. S., & McGreevy, A. M. (1986). Twins included and not included in
special programs for the gifted. Roeper Review, 9(2), 120-127.
Rollett, B. (1987). Diagnosis and intervention in education and therapy. In K. Hur-
relmann, F.-X. Kaufmann, & F. Losel (Eds.), Social intervention: Potential and
constraints (pp. 241-251). New York: de Gruyter.
Rose, S. A., & Wallace, I. F. (1985). Visual recognition memory: A predictor of later
cognitive functioning in preterms. Child Development, 56, 843-852.
Sawyer, W. W. (1943). Mathematician's delight. Baltimore, MD: Penguin.
Schneider, E. (1992). Analyse mathematischer Problemlosungsprozesse. Unpublished
master's thesis. University of Vienna, Institute of Psychology.
Shore, B., & Dover, A. (1987). Metacognition, intelligence and giftedness. Gifted
Child Quarterly, 32(1), 37-39.
Shotwell, J. M., Wolf, D., & Gardner, H. (1979). Exploring early symbolization:
Styles of achievement. In B. Sutton-Smith (Ed.), Play and learning (pp. 127-156).
New York: Gardner.
1980,1985)
Barr, S. (1965). A miscellany of puzzles. New York: Crowell.
Barr, S. (1969). Second miscellany of puzzles. New York: Macmillan. (Reissued as:
Mathematical brain benders. New York: Dover, 1982.)
Barry, D. T., & Lux, J. R. (1984). The Philips Academy Prize Examination in mathemat¬
ics. Palo Alto, CA: Dale Seymour Publications.
445
446 THE ART OF PROBLEM SOLVING
Bates, N. B., & Smith, S. M. (1980). 101 puzzle problems. Concord, MA: Bates
Publishing.
Berloquin, P. (1976). 100 geometric games. New York: Scribner's. (Retitled as: Geo¬
metric games. London: Unwin, 1980).
Berloquin, P. (1976). 100 numerical games. New York: Scribner's.
Berloquin, P. (1977). 100 games of logic. New York: Scribner's. (Retitled as: Games of
logic. London: Unwin, 1980.)
Berloquin, P. (1985). The garden of the sphinx. New York: Scribner's.
Birtwistle, C. (1971). Mathematical puzzles and perplexities. London: Allen & Unwin.
Brandes, L. G. (1975). The math wizard (Rev. ed.). Portland, ME: J. Weston Walch.
Bridgman, G. (1981). Lake Wobegon math problems (Rev. and enlarged ed.). Minnea¬
polis: Author.
Brousseau, A. (1972). Saint Mary's college mathematics contest problems. Palo Alto,
CA: Creative Publications.
Bryant, S. J., Graham, G. E., & Wiley, K. G. (1965). Nonroutine problems in algebra,
geometry, and trigonometry. New York: McGraw-Hill.
Bryant, V., & Postill, R. (1980). The Sunday Times book of brain teasers—Book 1.
London: Unwin. (Also published as The Sunday Times book of brain teasers. New
York: St. Martin's Press, 1982.)
Bryant, V., & Postill, R. (1983). The Sunday Times book of brain teasers—Book 2.
Englewood Cliffs, NJ: Prentice Hall.
Burkill, J. C., & Kundy, H. M. (1961). Mathematical scholarship problems. London:
Cambridge University Press.
Butts, T. (1973). Problem solving in mathematics. Glenview, IL: Scott, Foresman.
CEMREL. (1975). Elements of mathematics book, problem book (2 vols.). St. Louis, MO:
Author.
Clarke, B. R. (1994). Puzzles for pleasure. New York: Cambridge University Press.
Clarke, B. R., Gooch, R., Newing, A., & Singmaster, D. (1993). The Daily Telegraph
book of brain twisters No. 1. London: Pan.
Conrad, S. R., & Flegler, D. (1992). Math contests for high school (2 vols.). Tenafly,
NJ: Math League Press.
Conrad, S. R., & Flegler, D. (1992). Math contests grades 7 and 8 (2 vols.). Tenafly,
NJ: Math League Press.
Conrad, S. R., & Flegler, D. (1994). Math contests grades 4, 5, and 6 (2 vols.). Tenafly,
NJ: Math League Press.
Dorofeev, G., Potapov, M., & Rozov, N. (1973). Elementary mathematics: Selected
topics and problem solving. Moscow: Mir Publishers.
Dorrie, H. (1965). 100 great problems of elementary mathematics. New York: Dover.
Dowlen, N., Powers, S., & Florence, H. (1987). College of Charleston mathematics
contest books. Palo Alto, CA: Dale Seymour Publications.
Dudney, H. E. (1926). Modern puzzles. New York: Pearson. (New edition, nd [1936])
Dudney, H. E. (1932). Puzzles and curious problems. London: Nelson.
Sources for Problems
447
well.
Filipiak, A. S. (1942). Mathematical puzzles. New York: Bell.
Fisher L. & Kennedy, B. (1984). Brother Alfred Brousseau Problem Solving and
Mathematics Competition, Introductory Division. Palo Alto, CA: Dale Seymour
Publications. .
Fisher L„ & Medigovich, W. (1984). Brother Alfred Brousseau Problem Solving and
Mathematics Competition, Senior Division. Palo Alto, CA: Dale Seymour Pu 1-
cations.
Friedland, A. J. (1970). Puzzles in math and logic. New York: Dover.
Frohlichstein, J. (1962). Mathematical fun, games and puzzles. New York: Dover.
Fujimura, K. (1978). The Tokyo puzzles (Ed. by Martin Gardner). New York: Scnb-
ner s.
Gamow, G., & Stem, M. (1958). Puzzle-math. London: Macmillan.
Gardner, M. (1959). Arrow book of brain teasers. New York: Scholastic Press
Gardner M (1959). The Scientific American book of mathematical puzzles and diver-
“ s New York: Simon & Schuster. (Revised, with new afterword and refer¬
ences as: Hexaflexagons and other mathematical diversions. Chicago: University
Gardner, M. (1971). Martin Gardner's sixth book of mathematical games from Scientific
American. San Francisco: Freeman. (Reprinted in 1983 by the University of
Chicago Press.)
Gardner, M. (1975). Mathematical carnival. New York: Knopf. (Revised ed. in 1989
by the Mathematical Association of America, Washington, DC.)
Gardner, M. (1976). The incredible Dr. Matrix. New York: Scribner's. (Contains all
of The Numerology of Dr. Matrix)
Gardner, M. (1977). Mathematical magic show. New York: Knopf. (Revised ed. in
1990 by the Mathematical Association of America, Washington, DC.)
Gardner, M. (1977). More perplexing puzzles and tantalizing teasers. New York:
Archway (Pocket Books).
Gardner, M. (1978). Aha! Insight. New York: Scientific American & Freeman.
Gardner, M. (1979). Mathematical circus. New York: Knopf. (Revised ed. in 1992 by
the Mathematical Association of America, Washington, DC.)
Gardner, M. (1981). Science fiction puzzle tales. New York: C. N. Potter.
Gardner, M. (1982). Aha! Gotcha. New York: Freeman.
Gardner, M. (1983). Wheels, life and other mathematical amusements. New York:
Freeman.
Gardner, M. (1985). The magic numbers of Dr. Matrix. Buffalo, NY: Prometheus, 1985.
(Contains all of The Incredible Dr. Matrix)
Gardner, M. (1986). Entertaining mathematical puzzles. New York: Dover.
Gardner, M. (1986). Knotted doughnuts and other mathematical entertainments. New
York: Freeman.
Gardner, M. (1986). Puzzles from other worlds. New York: Vintage (Random House).
Gardner, M. (1987). Riddles of the sphinx. Washington, DC: New Mathematical
Library, Mathematical Association of America.
Gardner, M. (1988). Time travel and other mathematical bewilderments. New York:
Freeman.
Gardner, M. (1989). Penrose tiles to trapdoor ciphers. New York: Freeman.
Gardner, M. (1992). Fractal music, hypercards and more. New York: Freeman.
Gardner, M. (1994). My best mathematical and logical puzzles. New York: Dover.
Garvin, A. D. (1975). Discovery problems for better students. Portland, ME- Weston
Walch.
Sources for Problems 449
Gleason, A. M., Greenwood, R. E., & Kelly, L. M. (1980). The William Lowell Putnam
Mathematical Competitions. Problems and solutions: 1938-1964. Washington, DC:
Mathematical Association of America.
Gould, P. (1992). Senior Challenge '85-'91: Mathematical education on Merseyside.
Liverpool, UK: University of Liverpool.
Gould, P., & Porteous, I. (1984). Senior Challenge '80-'84: Mathematical education on
Merseyside. Liverpool, UK: University of Liverpool.
Graham, L. A. (1959). Ingenious mathematical problems and methods. New York:
Dover.
Graham, L. A. (1968). The surprise attack in mathematical problems. New York: Dover.
Grietzer, S. L. (1978). International Mathematical Olympiads, 1959-1977. Washington,
DC: Mathematical Association of America.
Haber, P. (1957). Mathematical puzzles and pastimes. Mount Vernon, NY: Peter
Pauper.
Halmos, P. R. (1991). Problems for mathematicians young and old (Dolciani Mathe¬
matical Expositions #12). Washington, DC: Mathematical Association of
America.
Higgins, A. M. (1971). Geometry problems. Portland, ME: J. Weston Walch.
Hill, T. J. (1974). Mathematical Challenges II—Plus Six. Washington, DC: National
Council of Teachers of Mathematics.
Holton, D. (1988-1990). Problem solving series. Leicester, UK: Mathematical Asso¬
ciation.
Honsberger, R. (1978). Mathematical morsels. Washington, DC: Mathematical Asso¬
ciation of America.
Hunter, J. A. H. (1965). Fun with figures. New York: Dover.
Hunter, J. A. H. (1966). More fun with figures. New York: Dover.
Hunter, J. A. H. (1972). Figures for fun. London: Phoenix House.
Hunter, J. A. H. (1976). Mathematical brain teasers (Corrected and enlarged). New
York: Dover.
Hunter, J. A. H. (1979). Challenging mathematical teasers. New York: Dover.
Hunter, J. A. H. (1983). Entertaining mathematical teasers and how to solve them. New
York: Dover.
Kendall, P. M. H., & Thomas, G. M. (1962). Mathematical puzzles for the connoisseur.
London: Griffin.
King, T. (c. 1927). The best 100 puzzles solved and answered. London: Foulsham.
Kinnaird, W. C. (Ed.). (1946). Encyclopedia of puzzles and pastimes. New York:
Grosset & Dunlap.
Klamkin, M. S. (1986). International Mathematical Olympiads, 1979-1985. Washing¬
ton, DC: Mathematical Association of America.
Kordemsky, B. A. The Moscow puzzles (Ed. Martin Gardner). New York: Scribner's.
Krechmer, V. A. (1974). A problem book in algebra (V. Shiffer, Trans.). Moscow: Mir
Publishers.
450 THE ART OF PROBLEM SOLVING
Krulik, S., & Rudnick, J. A. (1980). Problem solving: A handbook for teachers. Boston:
Allyn and Bacon.
Kutepov, A., & Rubanov, A. (1975). Problems in geometry (O. Meshkov, Trans.).
Moscow: Mir Publisher.
Kutepov, A., & Rubanov, A. (1978). Problem book: Algebra and elementary function
(L. Levant, Trans.). Moscow: Mir Publisher.
Larson, L. C. (1983). Problem solving through problems. New York: Springer-Verlag.
Lenchner, G. (1983). Creative problem solving in school mathematics. Boston:
Houghton Mifflin.
Loyd, S. (1914). Sam Loyd's cyclopedia of 5,000 puzzles, tricks and conundrums. New
York: Bigelow.
Loyd, S. (1927). Sam Loyd's tricks and puzzles (Vol. 1). New York: Experimenter
Publishing.
Loyd, S. (1928). Sam Loyd and his puzzles. New York: Barse & Co.
Loyd, S. (1959). Mathematical puzzles of Sam Loyd (Vol. 1). New York: Dover.
Loyd, S. (1960). Mathematical puzzles of Sam Loyd (Vol. 2). New York: Dover.
Moser, W., & Barbeau, E. (1976). The Canadian Mathematics Olympiads, 1969,1975.
Montreal: Canadian Mathematical Congress.
Morris, I. (1969). The Riverside puzzles. New York: Walker & Co.
Morris, I. (1970). The lonely monk and other puzzles. Boston: Little, Brown.
Morris, I. (1972). Foul play and other puzzles of all kinds. New York: Vintage (Random
House).
Moscovich, I. (1984). Super-games. London: Hutchinson.
Moscovich, I. (1991). Fiendishly difficult math puzzles. New York: Sterling.
Moscovich, I. (1991). Fiendishly difficult visual perception puzzles. New York: Sterling.
Moser, W. O. J., & Barbeau, E. J. (1978). The first ten Canadian Mathematics Olympiads
(1969-1978). Montreal: Canadian Mathematical Society.
Mosteller, F. (1965). Fifty challenging problems in probability. New York: Dover.
Mott-Smith, G. (1954). Mathematical puzzles for beginners and enthusiasts. New York:
Dover.
Newton, D. E. (1972). One hundred quickies for math classes. Portland ME: J. Weston
Walch.
Phillips, H. (1932). The week-end problems book. London: Nonesuch.
Phillips, H. (1933). The playtime omnibus. London: Faber & Faber.
Phillips, H. (1934). The sphinx problem book. London: Faber.
Phillips, H. (1936). Brush up your wits. London: Dent.
Phillips, H. (1937). Question time. London: Dent.
Phillips, H. (1945). Ask me another. London: Ptarmigan.
Phillips, H. (1945). Hubert Phillips's heptameron. London: Eyre & Spottiswoode.
Phillips, H. (1945). Something to think about. London: Ptarmigan. (With additional
Foreword, one problem omitted, and 11 problems added, London: Max Par¬
rish, 1958.)
Sources for Problems 451
Shortz, W. (1991). Will Shortz's best brain twisters. New York: Times Books (Random
House).
Shortz, W. (1993). Brain twisters from the First World Puzzle Championships. New
York: Times Books (Random House).
Sierpinski, W. (1964). A selection of problems in the theory of numbers. London:
Pergamon / Macmillan.
Sierpinski, W. (1970). 250 problems in elementary number theory. New York: Ameri¬
can Elsevier.
Snape, C., & Scott, H. (1991). How puzzling. New York: Cambridge University
Press.
Sole, T. (1988). The ticket to heaven and other superior puzzles. London: Penguin.
Sitomer, H. (1974). The new mathlete problems book. Valley Stream, NY: Nassau
County Interscholastic Mathematics League.
Soifer, A. (1987). Mathematics as problem solving. Colorado Springs: Center for
Excellence in Mathematics Education.
Stemhaus, H. (1963). One hundred problems in elementary mathematics. New York:
Pergamon.
Straszewicz, S. (1965). Mathematical problems and puzzles from the Polish Mathematical
Olympiads (J. Smsliska, Trans.). New York: Pergamon.
Tngg, C. W. (1967). Mathematical quickies. New York: McGraw-Hill.
Ulam, S. M. (1960). Problems in modern mathematics. New York: John Wiley.
Vout, C., & Gray, G. (1993). Challenging puzzles. New York: Cambridge University
Press.
Wall, H. S. (1963). Creative mathematics. Austin: University of Texas Press.
Wells, D. (1982). Can you solve these? Norfolk, UK: Stradbroke.
Wells, D. G. (1979). Recreations in logic. New York: Dover.
Williams, W. T., & Savage, G. H. (1940). The Penguin problems book. London:
Penguin.
Williams, W. T., & Savage, G. H. (c. 1940). The Strand problems book. London:
Newnes.
Williams, W. T., & Savage, G. H. (1944). The second Penguin problems book. London:
Penguin.
Williams, W. T., & Savage, G. H. (1946). The third Penguin problems book. London:
Penguin.
Yaglom, A. M., & Yaglom, I. M. (1964). Challenging mathematical problems with
elementary solutions (2 vols.). San Francisco: Holden-Day.
Readings on Problem Solving
Ackoff, R. L. (1978). The art of problem solving. New York: John Wiley.
Adams, J. L. (1974). Conceptual blockbusting. San Francisco: Freeman.
Adler, I. (1970). Mathematics and mental growth. London: Dobson.
Andre, T. (1986). Problem solving and education. In G. Phye & T. Andre (Eds.),
Cognitive classroom learning. Orlando, FL: Academic Press.
Arnold, W. R. (1971). Students can pose and solve original problems. The Mathe¬
matics Teacher, 64, 325.
Averbach, B., & Chein, O. (1980). Mathematics: Problem solving through recreational
mathematics. San Francisco: Freeman.
Bransford, J. D., & Stein, B. S. (1984). The ideal problem solver. New York: W. H.
Freeman.
Brown, S. I., & Walter, M. I. (1983). The art of problem posing. Hillsdale, NJ: Lawrence
Erlbaum.
Butts, T. (1985). In praise of trial and error. The Mathematics Teacher, 78,167.
Charles, R., & Lester, F. (1982). Teaching problem solving: What, why, and how. Palo
Alto, CA: Dale Seymour Publications.
Chipman, S., Segal, J., & Glaser, R. (1985). Thinking and learning skills: Vol. 2. Research
and open questions. Hillsdale, NJ: Lawrence Erlbaum.
Cofman, J. (1990). What to solve? Problems and suggestions for young mathematicians.
New York: Oxford University Press.
Costa, A. (1984, November). Mediating the metacognitive. Educational Leadership,
pp. 57-62.
Curcio, F. (Ed.). (1987). Teaching and learning: A problem solving focus. Reston, VA:
NCTM.
Davis, R., Jockusch, E., & McKnight, C. (1978). Cognitive processes in learning
algebra. Journal of Children's Mathematical Behavior, 2(1).
453
454 THE ART OF PROBLEM SOLVING
Derry, S. J., & Murphy, D. A. (1986). Designing systems that train learning ability:
From theory to practice. Review of Educational Research, 56(1), 1-39.
Emmet, E. R. (1981). Learning to think. Verplanck, NY: Emerson.
Fisher, R. B. (1981). Brain games. London: Fontana.
Fixx, J. F. (1978). Solve it! New York: Doubleday.
Frederiksen, N. (1984). Implications of cognitive theory for instruction on problem
solving. Review of Educational Research, 54(3), 363-407.
Gardner, M. (1978). Aha! Insight. New York: Scientific American & Freeman.
Gardner, M. (1982). Aha! Gotcha. San Francisco: Freeman.
Gordon, W. J. J. (1961). Synectics—The development of creative capacity. New York:
Flarper & Row.
Hadamard, J. (1954). The psychology of invention in the mathematical field. New York:
Dover.
Heiman, M., Narode, R., Slomianko, J., & Lochhead, J. (1987). Thinking skills:
Mathematics, teaching. Washington, DC: National Education Association.
Honsberger, R. (1970). Ingenuity in mathematics. Washington, DC: Mathematical
Association of America.
Honsberger, R. (1973). Mathematical gems (Vol. 1, Dolciani Mathematical Exposi¬
tions #1). Washington, DC: Mathematical Association of America.
Honsberger, R. (1976). Mathematical gems (Vol. 2, Dolciani Mathematical Exposi¬
tions #2). Washington, DC: Mathematical Association of America.
Honsberger, R. (1978). Mathematical morsels (Dolciani Mathematical Expositions
#3). Washington, DC: Mathematical Association of America.
Honsberger, R. (1979). Mathematical plums (Dolciani Mathematical Expositions #4).
Washington, DC: Mathematical Association of America.
Honsberger, R. (1985). Mathematical gems III (Dolciani Mathematical Expositions
#9). Washington, DC: Mathematical Association of America.
Honsberger, R. (1991). More mathematical morsels (Dolciani Mathematical Exposi¬
tions #10). Washington, DC: Mathematical Association of America.
Hough, J. S. (Ed.). Problem solving: Newsletter (Vol. 1-5). Philadelphia: Franklin
Institute Press.
Hughes, B. (1975). Thinking through problems. Palo Alto, CA: Creative Publications.
Jensen, R. J. (1987). Stuck? Don't give up! Subgoal-generation strategies in rpoblem
solving. The Mathematics Teacher, 80, 614.
Karmos, J., & Karmos, A. (1987). Strategies for active involvement in problem
solving. In M. Heiman & J. Slomianko (Eds.), Thinking skills instruction: Con¬
cepts and techniques (pp. 99-110). National Educational Association.
Kluwe, R. (1987). Executive decisions and regulation of problem solving behavior.
In F. Weinert & R. Kluwe (Eds.), Metcognition, motivation and understanding.
Hillsdale, NJ: Lawrence Erlbaum.
Krulik, S. (Ed.). (1980). Problem solving in school mathematics, 1980 yearbook. Reston,
VA: National Council of Teachers of Mathematics.
Readings on Problem Solving 455
Krulik, S., & Rudnick, J. (1987). Problem solving: A handbook for teachers (2nd ed.).
Boston: Allyn and Bacon.
Krulik, S., & Rudnick, J. (1989). Problem solving: A handbook for senior high school
teachers. Boston: Allyn and Bacon.
Krulik, S., & Rudnick, J. (1993). Reasoning and problem solving: A handbook for
elementary school teachers. Boston: Allyn and Bacon.
Krulik, S., & Rudnick, J. (1995). The new sourcebook for teaching reasoning and problem
solving in elementary schools. Boston: Allyn and Bacon.
Krulik, S., & Rudnick, J. (1996). The new sourcebook for teaching reasoning and problem
solving in secondary schools. Boston: Allyn and Bacon.
Mason, J. (1978). Learning and doing mathematics. Milton Keynes, UK: Open Univer¬
sity Press.
Mason, J., with Burton, L., & Stacey, K. (1985). Thinking mathematically. Reading,
MA: Addison-Wesley.
Mayer, R. (1986). Mathematics. In R. Dillon & R. Sternberg (Eds.), Cognition and
instruction. Orlando, FL: Academic Press.
McKim, R. H. (1980). Thinking visually: A strategy manual for problem solving. Palo
Alto, CA: Dale Seymour Publications.
Moses, S. (1974). The art of problem-solving. London: Transworld.
Mottershead, L. (1978). Sources of mathematical discovery. Oxford, UK: Blackwell.
Mottershead, L. (1985). Investigations in mathematics. Oxford, UK: Blackwell.
Nickerson, R. (1981, October). Thoughts on teaching thinking. Educational Leader¬
ship, pp. 21-24.
Nickerson, R., Perkins, D., & Smith, E. (1985). The teaching of thinking. Hillsdale, NJ:
Lawrence Erlbaum.
Polya, G. (1945). How to solve it. Princeton, NJ: Princeton University Press.
Polya, G. (1954). Introduction and analogy in mathematics. Princeton, NJ: Princeton
University Press.
Polya, G. (1954). Patterns of plausible inference. Princeton, NJ: Princeton University
Press.
Polya, G. (1962). Mathematical discovery (2 vols.). New York: John Wiley. (Combined
ed. with foreword by Peter Hilton, bibliography extended by Gerald Alexan-
derson, and index extended by Jean Pedersen, New York: John Wiley, 1981.)
Posamentier, A. S„ & Schulz, W. (Eds.). (1996). The art of problem solving: A resource
for the mathematics teacher. Thousand Oaks, CA: Corwin.
Reeves, C. A. (1987). Problem solving techniques helpful in mathematics and science.
Reston, VA: National Council of Teachers of Mathematics.
Schoenfeld, A. H. (1983). Problem solving in the mathematics curriculum. Washington,
DC: Mathematical Association of America.
Schoenfeld, A. H. (1985). Mathematical problem solving. Orlando, FL: Academic
Press.
Segal, J., Chipman, S„ & Glaser, R. (Eds.). (1985). Thinking and learning skills. Vol.
2: Relating instruction to research. Hillsdale, NJ: Lawrence Erlbaum.
456 THE ART OF PROBLEM SOLVING
457
458 THE ART OF PROBLEM SOLVING
Professor Schulz has been concentrating his efforts and interests in the theoretical
and practical aspects of the instruction of mathematics as well as the preparation
of mathematics teachers. Professor Schulz is an author/editor of numerous Ger¬
man mathematics textbooks for secondary and upper-elementary school students.
He also is coauthor of the textbook Gruppoide und Funktionalgleichungen designed
to train mathematics teachers. As an author of numerous journal articles, he
frequently lectures at conferences throughout Europe. Throughout his career. Dr.
Schulz has worked with highly motivated students in the area of problem solving,
and he has had numerous enriching experiences in which students stumble onto
uniquely clever solutions to relatively common problems.
About the Contributors
Ethan Akin received his B.S. degree in 1965 from the City College of New York,
where he is now Professor of Mathematics. His many research papers and exposi¬
tory articles include such titles as "Recurrence of the Unfit," "Bulgarian Solitaire,"
"Ovulation Regulation: The Right Number or Nothing," and "The Spiteful Com¬
puter," as well as monographs: The Metric Theory of Banach Manifolds and The
Geometry of Population Genetics. His comprehensive survey. The General Topology of
Dynamical Systems was published by the American Mathematical Society in 1993.
Professor Akin has helped design an alternative problem-solving calculus course
and taught one of its initial two sections. The psychology of teaching and learning
remains one of his strong interests.
Steven R. Conrad has taught since 1980 at Roslyn High School, Roslyn Heights,
New York. Prior to that, he taught at Benjamin Cardozo High School in Bayside
and Francis Lewis High School, in Flushing—both in New York City. He received
a B.S. from Queen's College, an M.S. from Yeshiva University, and a certificate in
School Administration from St. John's University. He has done additional gradu¬
ate work at the University of San Francisco and Fordham University. A speaker at
numerous international, national and regional conferences, Mr. Conrad has also
been a contributor to the problem sections of many mathematics journals. He has
served as Editor of Summation and Associate Editor of The New York State Mathe¬
matics' Teachers' Journal and The Mathematics Student Journal. He has had articles
published in Mathematics Magazine and The Mathematics Teacher. Mr. Conrad has
coauthored six books of mathematics contests.
Ira Ewen has taught mathematics at the secondary and college levels ever since
he earned his A.B. and MA degrees in pure mathematics in 1952 and 1953 from
Harvard University. He has served as a teacher of mathematics, an assistant
principal supervising a high school mathematics department, and a high school
principal. He has been a teaching fellow in mathematics at Harvard University, a
459
460 THE ART OF PROBLEM SOLVING
Hans Humenberger is a high school teacher in Vienna and holds a master's degree
in Mathematics from the University of Vienna. His interest in the use of applica¬
tions to teach mathematics brought him to work as an assistant in the Institute of
Mathematics at the University of Vienna under the auspices of Professor Reichel.
There he was involved in the training of student teachers and pursued the study
of the relationship between language and mathematics. His main interest is
problem solving in mathematics. In 1992, he was awarded a doctorate from the
Department of Mathematics at the University of Vienna.
sional career, he has tried to combine all these interests. His fields of research are
universal algebra, applications of algebra in computer science, and the history and
didactics of mathematics. He has published a number of research articles in these
areas and is coauthor of several books (Algebra fur Informatiker [Algebra for Computer
Scientists] for university students and Geschichte der Mathematik im Schulunterricht
[History of Mathematics for the Schools], which is intended for teachers of secondary
schools). At the University of Technology of Vienna, he teaches mathematics to
students who want to become engineers or teachers in mathematics and computer
science.
Karl Kiesswetter was bom in Sudetendland and has studied mathematics at the
University of Koln. He was a high school teacher and then held positions at the
Universities of Munster, Bielefeld, and Hamburg. He has been interested in
mathematics education and fostering creativity among gifted school children
("Hamburg model").
Wei Lee currently teaches mathematics at the City College of New York. He
received his B.A. from the Shanghai University of Science and Technology. Upon
graduation, he worked at the Computer Research Institute, when, after 4 years of
problem-solving experiences of a somewhat different nature, he emigrated to the
United States, where he is currently pursuing his doctoral studies while continu¬
ously looking for novel mathematical problem solving experiences. His contribu¬
tions to this book exhibit his interest in using combinatorics and graph theory to
solve routine and nonroutine mathematical problems.
Fred Paul was with the Bureau of Mathematics of the New York State Education
Department for 27 years until his retirement in October 1992. For the last 17 of
these years he served as Bureau Chief, and in that capacity he directed all aspects
About the Contributors 463
Jan Trojak was bom in Prague, Bohemia, where he also received his university
education. In 1961, he graduated from the Faculty of Science of the Pedagogical
University with a diploma of a teacher of mathematics and descriptive geometry.
Since that time he worked and studied at the Faculty of Mathematics and Physics
of the Charles University in Prague. Between 1968-1971 and 1973-1975, he worked
as a lecturer at the College of Science of the University of Basrah and at the
University of Baghdad in Iraq, giving lectures on projective, non-Euclidean and
differential geometries, graph theory, calculus and analytical geometry, and linear
algebra. Dr. Trojak has cowritten a textbook (published in three volumes) on
About the Contributors 465
analytical geometry of the Euclidean, affine spaces, and affine and congruence
transformations.
MAY U
WAT n
8T997
FEB 2 *:
CO
CO
0©
FFR1 2
LU t 6 199ft
0 C1 1 2 lWf~
OCT1 t10ft
wv' iQqn
iJJJ
APR 1 7 2000
0 1 64 0419797 6
The Art of Problem Solving
A Resource for the Mathematics Teacher
Editor:
Alfred S. Posamentier
The City College of the City University of New York
Associate Editor:
Wolfgang Schulz
Humboldt University, Berlin
As a mathematics teacher, you know how hard it can be for your students to understand and solve math
problems. The old ways of problem solving don’t always work. Even the most innovative teachers need
some fresh ideas to make mathematics something students comprehend and enjoy.
Problem solving is the main theme of this new, idea-filled handbook. The chapter authors look at the
subject in a completely new light, and the result is an enticing, entertaining, and useful resource.
The editors of this guidebook present a host of interesting ideas that range from practical to theoretical,
from common to glitzy, that you can adapt for use in your classroom. With margin notes identifying sub¬
ject matter and strategy type, this book is as easy to read as it is valuable.
Mathematics professionals from all over the world bring you their personal favorite strategies for prob¬
lem solving. The chapters are independent of each other, and the editors encourage you to use the chap¬
ters in whatever order suits your needs.
If you have ever despaired of making mathematics into a favorite subject for yourself and your class, this
is the book you need. These passionate professionals give you new insight into the art of problem solving.