Vous êtes sur la page 1sur 21

Mathematics and Origami: The Ancient Arts Unite

Jaema L. Krier
Spring 2007
Abstract
Mathematics and origami are both considered to be ancient arts, but until the 1960s the two were
considered to be as dierent as night and day. Turns out they are not as dierent as everyone thought.
In fact, origami can be used to explain many mathematical concepts in elds such as geometry, calculus,
abstract algebra and others. The relationship between mathematics and origami has yet to be fully
explained. New folds and models are being developed by artists, architects, mathematicians, and other
origami enthusiasts. A set of Postulates, similar to those of Euclidean geometry, have been established
and some origami folds can even solve quadratic and cubic equations. An exploration of various techniques
and models can give a true understanding of how important origami is to the eld of mathematics.
1 Introduction
When the word origami is mentioned, most people probably think of little paper cranes or maybe even
paper airplanes. Although origami is commonly referred to as the art of paper folding, the study of origami
reveals that there are many mathematical characteristics of it as well. Origami can be utilized in the study
of geometry, calculus, and even abstract algebra. For students, origami could be the tangible key to their
mathematical comprehension.
Origami originated in China where it was known as Zhe Zhi. It later became popular in Japan, and is
now considered a Japanese art. Although any kind of paper can be used to construct origami models, the
most common is known as Kami, Japanese for paper. Just slightly thicker than tissue paper, Kami is usually
colored on one side and white on the other. In Japan, the most widely used paper is called Washi, which is
made of wood pulp [16].
There are various styles of origami such as traditional, rigid, and modular, to name a few. Traditional
origami keeps with the belief that models should be made using a single sheet of square paper which cannot
be cut or secured (glued) in any way. Paper cranes and planes would fall into this category. Rigid origami
explores the idea of folding a single sheet of paper in such a way that it collapses easily without bending
the regions between its creases. In other words, it can be folded at with a rigid motion. The solar panel
arrays used for space satellites were designed using the rigid map fold invented by Koryo Miura, a Japanese
astrophysicists [16].
Unfortunately, utilizing a single sheet of paper has its limitations, thus modular origami was invented.
Modular origami models are made with more than one sheet of paper and are formed by constructing units
which then lock together to form a larger model. One dening characteristic of modular origami states that
the larger model must be made up of identical units. Although models made with dierent units can be
constructed and are often referred to as modular, they do not truly meet the requirements of the modular
denition [15].
The rst known example of modular origami can be dated back to 1734. It was a cube called the magic
treasure chest. However, the traditional Kusudama, paper owers strung together into a sphere, is considered
to be the precursor to modern day modular origami. Modular origami did not truly become popular until the
1960s [15]. Since then, mathematicians have discovered its uses in explaining a vast number of mathematical
models. By inventing new folds and models they are continually contributing valuable information to the
eld of mathematics.
2 Relating Origami to Mathematics
Understandably, it would not be odd to compare origami with geometry. Origami can, after all, be used to
construct various geometric shapes. It may come as a surprise, however, to learn that origami has its own set
1
of postulates much like geometry. The process of paper folding can be reduced to seven simple postulates.
The rst six were developed by a nuclear scientist by the name of Humiaki Huzita, and they are considered
to be the most powerful known to date. The seventh was later discovered by Jacques Justin which has since
been conrmed as accurate [3], [4], [9], [11].
POSTULATE 1 Given two points P
1
and P
2
, one can fold a single crease which passes through them.
It is evident this parallels Euclids rst postulate, through any two distinct points, it is possible
to draw (exactly) one straight line, which can be accomplished easily with a straightedge.
POSTULATE 2 Given two points P
1
and P
2
, one can fold a crease placing P
1
onto P
2
.
Straightedge and compass construction states we are able to construct a bisector of a line
segment. In general, this postulate does just that. By placing P
1
onto P
2
we are simply locating
the midpoint of the line segment P
1
P
2
and then folding a crease at that point which then becomes
the bisector.
POSTULATE 3 Given two lines L
1
and L
2
, one can fold a crease placing L
1
onto L
2
.
When L
1
is not parallel to L
2
, this origami move is equivalent to the bisecting of an angle using
straightedge and compass. The resulting crease will go through the intersection of lines L
1
and
L
2
, resulting in the bisection of their vertical angles.
POSTULATE 4 Given a point P
1
and a line L
1
, one can fold a crease which will be to L
1
and pass
through P
1
.
Obviously, this emulates the Euclidean construction allowing us to drop a line from a given
point to a given line.
POSTULATE 5 Given two points P
1
and P
2
, and a line L
1
, one can fold a crease that places P
1
onto L
1
and passes through P
2
This fold can be dicult to accomplish. It may be necessary to fold P
1
onto L
1
at a point which
is o of the sheet of paper in order for the crease to pass through P
2
.
QUESTION 1 What exactly is this postulate accomplishing?
The answer may astound you. It is actually solving a quadratic equation. The crease which is
constructed by this postulate is actually tangent to the parabola with focus P
1
and directrix L
1
[4]. Let us prove it.
Proof 1 Let L
1
be the line forming the bottom edge of our paper. Let P
1
be a point toward the middle,
fairly close to L
1
, and P
2
be a point on the left or right edge of our paper.
Perform Postulate 5, as shown below, leaving the paper folded in this position, call the creased
line L
2
.
Figure 1: Perform Postulate 5.
From the point P
1
construct a line which is to the folded portion of L
1
using Postulate 4.
Let X be the point where this line intersects L
2
.
Figure 2: Completion of Postulate 4.
By opening our paper, we observe that the line segment XP
1
and the line segment from X to L
1
,
call it XA, are equal. You may want to fold and unfold L
2
to convince yourself of this.
Figure 3: The segments from X to P
1
and L
1
are equidistant.
Therefore, X is the point on L
2
which is equidistance to both P
1
and L
1
. By denition, this point
is on the parabola with focus P
1
and directrix L
1
.
By Postulate 3, L
2
is also the bisector of AXP
1
. Therefore, any point on L
2
is equidistant
to A and P
1
Choose a point, Y , on L
2
between P
2
and X. Using Postulate 4, construct the line to L
1
passing through Y , call it Y B. Note, that Y BA is right, thus Y B < Y A = Y P
1
. Since all
points of the parabola must be equidistant to both the directrix, L
1
, and the focus, P
1
, we know
the parabola lies above L
2
at this point.
Figure 4: The point Y lies below the parabola with focus P
1
and directrix L
1
.
Similarly, we can show this is true for any point on L
2
from X to L
1
.
Thus, L
2
is the tangent line to the parabola.

(Proof and graphics adapted from [4])


In order to see this more clearly, choose numerous points along the left and right edges of your
paper to represent P
2
. For each of these points, complete Postulate 5 to reveal the outline of the
parabola.
Figure 5: Tangent lines to the parabola with focus P
1
and directrix L
1
.
Construction with strictly straightedge and compass limits our ability to trisect an angle or double the
volume of a cube (constructing a line with length
3

2) [3], [4], [9], [12]. These feats cannot be accomplished


by simple Euclidean construction. Such problems can, however, be addressed with the application of origami
and we will discuss them in Section 3. The following postulate is the key to unlocking this ancient problem.
POSTULATE 6 Given two points P
1
and P
2
and two lines L
1
and L
2
, one can fold a crease that places
P
1
onto L
1
and P
2
onto L
2
.
QUESTION 2 How does this postulate help us in solving cubic equations?
In general, the crease created by this fold is the line which is tangent to two separate parabolas,
one with focus P
1
and directrix L
1
, the other with focus P
2
and directrix L
2
[4]. Lets see how
the slope of this tangent line helps us solve the cubic equation x
3
+ax
2
+bx +c = 0 [3].
Proof 2 Let P
1
be the point (a, 1) and P
2
be (c, b).
Let L
1
be dened by y = 1 and L
2
by x = c.
Perform Postulate 6, placing P
1
and P
2
on L
1
and L
2
respectively, call this crease L
3
. Since L
3
is not parallel to the axis, let it be dened by
y = tx +u
By proof 1, L
3
is tangent to the parabola with focus P
1
and directrix L
1
.
Figure 6: Slope of the tangent line to
1
.
This parabola,
1
, is dened by 4y = (xa)
2
, since its focus is (a, 1) and the directrix is y = 1.
Let (x
1
, y
1
) be a point of L
3
. Thus, L
3
is tangent to
1
at
4y
1
= (x
1
a)
2
[A]
Taking the derivative gives us y

1
=
1
2
(x
1
a), so the slope, t, of the line is
t =
1
2
(x
1
a) [B]
Using the point slope formula we write the equation of the line as y y
1
=
1
2
(x
1
a)(x x
1
)
which we will rewrite as
y =
1
2
[(x
1
a)x (x
1
a)x
1
] +y
1
Notice that this is equivalent to y = tx
1
2
(x
1
a)x
1
+y
1
and results in
u =
1
2
(x
1
a)x
1
+y
1
By substituting in for y
1
from [A], and t from [B], we see u = tx
1
+
1
4
(x
1
a)
2
. Notice this
can be reduced further to u = tx
1
+t
2
. We insert x
1
from our slope formula, [B], above to get
u = t(2t +a) +t
2
, and nally
u = t
2
at
Similarly, the equation for the parabola,
2
, with focus (c, b) and directrix x = c can be written
as 4cx = (y b)
2
which intersects the parabola at
4cx
2
= (y
2
b)
2
By implicit dierentiation we obtain y

2
=
2c
y
2
b
. Therefore, the slope of our line is
t =
2c
y
2
b
The point slope formula gives us the equation of the line as y y
2
=
2c
y
2
b
(x x
2
). This can be
rewritten to show
y =
2c
y
2
b
x
2cx
2
y
2
b
+y
2
Thus, we can say
u = y
2

2cx
2
y
2
b
After substitutions of x
2
and t, this reduces to
u =
2c
t
+b
c
t
= b +
c
t
Therefore, u = t
2
at AND u = b +
c
t
, therefore t
2
at = b +
c
t
. When c = 0,
t
3
+at
2
+bt +c = 0
However, when c = 0, this means that P
2
is on L
2
so either t = 0 or u = b. In this case, we see
t
2
+at +b = 0

(Proof based on [3])


Since the trisecting of angles and doubling of cubes reduces to the solving of a cubic equation,
one can begin to realize how this postulate can be utilized for such problems.
POSTULATE 7 Given a point P and two lines L
1
and L
2
, one can fold a crease placing P onto L
1
which
is to L
2
[3].
3 Useful Origami Techniques
There are several origami folds which prove to be very useful both for construction and for the solving of
various mathematical problems. One of the most common techniques is the folding of a length into n
ths
. This
is necessary for many origami models and should be considered a vital operation in origami construction.
Dividing paper in
1
2
,
1
4
,
1
8
,...
1
2
n
is not considered dicult. It may not be obvious, however, how one would
divide a paper into
1
3
,
1
5
,
1
7
,...
1
2n+1
. Lets start with learning the case for
1
3
and then generalize it for
1
2n+1
.
3.1 Dividing a Length Into Thirds
1. Start with a square piece of paper.
2. Fold paper in half. Open.
3. Fold the paper on the diagonal. Open.
4. Now fold one half of the paper on the diagonal such that it intersects the main diagonal.
Open.
Your paper should resemble Figure 7.
Figure 7: After completion of Step 4.
QUESTION 3 What has this series of folds done for us?
The point where the two diagonals intersect, call it P, is exactly
2
3
from the left of our paper.
Thus, the remainder is equal to
1
3
. The series of folds has allowed us to identify this point [7].
Proof 3 Let P = (x, x) and the squares sides equal one unit.
Figure 8: Dividing a paper into 3
rds
.
This gives us AB = 1, FD = x, DP = x, and DC = 1 x.
Note ABC is similar to PDC. So,
|AB|
|PD|
=
|CB|
|CD|
=
1
x
=
1
2
1 x
=x = 2 2x
=3x = 2
x = 2/3

(Proof and graphics adapted from [7])


3.2 Dividing a Length Into n
ths
Proving this method in general is very similar to the previous proof. Since we are wanting to divide our
paper into segments equal to
1
n
the original length, we must rst fold our paper by
1
n1
its length [7]. See
Figure 9.
Proof 4 Let P = (x, x) and the squares sides equal one unit.
Figure 9: Dividing a paper into n
ths
. In this case n = 5.
As shown in proof 3 and by use of induction we can see,
|AB|
|PD|
=
|CB|
|CD|
=
1
x
=
1
n1
1 x
=x = (n 1)(1 x)
=x = n nx 1 +x
=nx = n 1
x =
n 1
n

(Proof and graphics adapted from [7])


This technique will allow us to expand our construction abilities and help to solve complicated tasks such
as the doubling of a cube.
3.3 Hagas Theorem
Hagas Theorem was named for Kazuo Haga, a retired professor of biology from the University of Tsukuba,
Japan [7]. His theorem allows for the proof of more complicated origami constructions.
THEOREM 1 By choosing a point, P, on top of a square and setting the bottom corner onto P, we can
observe three similar triangles, A, B, and C.
Figure 10: Three similar triangles formed by placing a corner onto a point, P.
Proof 5 Observe from Figure 10 that
2 +3 = 90

1 +2 = 90

3 +4 = 90

This results in
1 = 3
2 = 4
By denition of similarity we have
A B
Similarly, one can show
A B C

(Theorem, proof, and graphics adapted from [7])


3.4 Doubling the Cube
As stated previously, the doubling of a cube cannot be accomplished with straightedge and compass alone.
The problem began in Greece when Eratosthenes wrote of the gods demanding an alter twice the size of the
existing one in order to rid the people of a plague. Plato exclaimed that the gods only wanted to shame
the Greeks for their neglect of mathematics and their contempt of geometry. Although the Greeks were
unable to accomplish this task through straightedge and compass construction, other methods of doubling
the cube were developed [12].
By utilizing the 6
th
postulate, however, we are able to complete the task of doubling the cube with a few
folds of a paper.
3.4.1 Peter Messers Solution
1. Begin with a square sheet of paper.
2. Fold the paper into 3
rds
from top to bottom. Open.
3. Let one of the creases be L
1
and let P
1
be the point where the other crease meets the papers
edge.
4. Let L
2
be the edge of the paper opposite P
1
, while P
2
is the corner closest to P
1
. See Figure
11 for proper labeling.
5. Perform Postulate 6 without re-opening the paper.
Figure 11: Steps 1 - 4.
QUESTION 4 What has this fold done?
The point where P
2
falls on L
2
is equal to
X
Y
=
3

2 where X is the length of L


2
from P
2
to the
corner, which includes the intersection with L
1
and Y is the remaining length of L
2
[7].
Figure 12:
X
Y
=
3

2.
Proof 6 Let Y = 1. This results in the sides of the square, s, being equal to X + 1.
Let A be the point where L
1
intersects X, B be the lower left corner, and C be the point where
the crease meets the bottom edge.
Figure 13: Solving for X, using Hagas Theorem.
Let BC = d. Since the bottom edge equals X + 1, this results in P
2
C = X + 1 d. Rewriting d
via the Pythagorean Theorem we get,
d =
(X
2
+ 2X)
(2X + 2)
Also, notice that P
1
P
2
=
1
3
s, which in terms of X is
(X+1)
3
.
We can also derive the value of AP
2
by taking X and subtracting
1
3
s, giving us a value of
(2X1)
3
.
Now, by Hagas Theorem, we know that P
2
AP
1
is similar to CBP
2
. Therefore we can say,
d
X + 1 d
=
2X 1
X + 1
=
X
2
+ 2X
X
2
+ 2X + 2
=
2X 1
X + 1
=X
3
+ 3X
2
+ 2X = 2X
3
+ 3X
2
+ 2X 2
=X
3
= 2
=X =
3

(Theorem, proof, and graphics adapted from [7])


This is not the only method for doubling a cube. Koshiro Hatori has another method that proves to be
quite eloquent.
3.4.2 Koshiro Hatoris Rendition
1. Begin with a square sheet of paper and let the sides represent 8 units.
2. Divide the paper into 4
ths
on the x-axis and 8
ths
on the y-axis. Open.
3. Label your orgin as shown in Figure 14.
Figure 14: Where to label the orgin.
4. Let P
1
= (0, 1) and P
2
= (2, 0).
5. Let L
1
be the line y = 1 and L
2
be x = 2.
6. Perform Postulate 6. Open. See Figure 15.
Figure 15: Steps 1 - 6 of Hatoris method.
QUESTION 5 How does this method solve the dilemma of doubling a cube?
Recall proof 2. There we showed that Postulate 6 helps us solve cubic equations. The steps listed
above solved the equation x
3
2 = 0 where a and b were both equal to zero [3]. Therefore, the
slope of the crease is our solution. Notice that the rise, Y , appears to be equal to 5 and the run,
X, is approximately 4. Therefore,
Y
X

5
4
1.25
3

2
You can see that you would derive the same result by simply setting the rise, Y , to approximately
5
8
and the
run, X, approximately
1
2
. However, the former set-up allows for easier reference to proof 2.
3.5 Trisecting the Angle
Euclidean constructions enable us to easily bisect an angle, but the trisecting of an angle has proved to be
an ancient problem. It simply cannot be done with straightedge and compass alone. By utilizing origami,
however, we can trisect the angle.
3.5.1 Abes Method
1. Begin with a square piece of paper.
2. Fold a crease, call it L
1
, which passes through the bottom, left corner, P
2
. This will form an
angle, A, between the bottom edge of your paper and L
1
. Open.
3. Fold the paper into 4
ths
from top to bottom. Open.
4. Let L
2
be the crease
1
4
from the bottom edge of the paper.
5. Label P
1
as the point where the left edge and the
1
2
way mark intersect.
Figure 16: Set-up for trisecting the angle, A.
6. Perform Postulate 6. Do not re-open.
7. Extend L
2
by folding a new crease, L
3
. See Figure 17. Open.
8. Continue L
3
so that it goes through the corner, P
2
.
Figure 17: Extend L
2
to form L
3
.
QUESTION 6 Is L
3
the line which trisects A?
This answer is yes! The angle formed by L
3
and the bottom edge of our paper is equal to
2
3
that of A [4]. We can use Postulate 3 to fold the bottom edge of paper onto L
3
, thus forming
another crease, L
4
, which gives us all three angles. Lets prove that these three angles are in fact
equal to one another, thus each equal to
1
3
A.
Proof 7 Beginning with Figure 17, let A be the point where P
1
lies on L
1
, B be the point where L
2
intersects L
3
, and C be the point where P
2
lies on L
2
.
Using Postulate 1, fold the crease which goes through A, B, and C. Notice that this is the line
formed by the edge of your paper in Figure 17, therefore giving us P
2
BA = P
2
BC = 90

. This
can be seen in Figure 18.
Perform Postulate 4, constructing a line which is to the bottom edge of your paper, at say D,
that goes through C.
Figure 18: Angle trisection proof.
Now observe,
AB = BC = CD
This is true, since each of them are
1
4
the width of the paper.
Since P
2
BA = P
2
BC and P
2
B = P
2
B we have that P
2
BA

= P
2
BC by SAS. Therefore,
1 = 2
We can use the same logic to show P
2
BC

= P
2
DC. This results in
2 = 3
=1 = 2 = 3

(Theorem, proof, and graphics adapted from [4])


3.6 Constructing a 60

Angle
It seems intuitive to say, that in the process of constructing various models, the need will arise to construct
an angle of a given degree. The formation of a 60

angle is one which will prove handy in the next section.


Lets learn Francis Ows 60

Unit which was published in the December 1986 issue of British Origami
Magazine (No. 121 p 32) [5].
1. Fold your paper in 4
ths
. Open.
2. Fold inward on the
1
4
and
3
4
creases so that the edges of your paper touch the
1
2
line.
Figure 19: Step 2.
3. Using only the top layer of the right hand side, fold the paper in half again, call this crease
L
1
. You will only need to do this for the top quarter of your paper. Unfold this step to return
to the previous step.
4. Using Postulate 5, we place the upper left corner, P
1
, onto L
1
so that it passes through P
2
,
where P
2
is the point at the top of the
1
2
line. Call this L
2
.
Figure 20: Steps 3 and 4.
5. Fold the upper right corner onto L
2
so that the crease passes through P
2
.
Figure 21: Step 5.
The resulting angle equals 60

.
Figure 22: Francis Ows 60

Unit.
(Theorem and graphics adapted from [5])
Proof 8 Recall that we folded the paper into 4
ths
.
Figure 23: Ows 60

Unit proof.
Therefore,
AB = AC =
1
4
= DE = EF
However, B and C lie midway between the latter two segments, by construction. Thus,
BM =
1
8
= MC
=BC =
1
4
This results in an equilateral ABC.
BAC = 60

4 Various Modular Models & Their Characteristics


Origami can be used to construct various mathematical models from 2-space, 3-space, and even fractional
space. In 2-space, obviously origami can construct various polygons such as squares, rectangles, triangles,
pentagons, hexagons, decagons, and dodecagons to name a few. In 3-space, it is possible to construct
regular polyhedra models such as tetrahedrons, cubes, octahedrons, and dodecahedrons, as well as semi-
regular polyhedra like the truncated tetrahedrons, truncated octahedrons, lesser rhombicosidodecahedrons,
and others. It is even possible to construct models which represent fractional dimensions, obviously, referring
to fractals. Here we will explore a few of the more popular models.
4.1 Buckyballs
Buckyballs are named after the American artist and architect, Richard Buckminster Fuller. Born in 1895,
Fuller was considered a visionary who excelled in architectural design and inventions. He designed the
geodesic dome structure, which is known for its self supporting nature. It is the only man-made structure
which gets proportionally stronger as it grows in size. Buckminsters contribution revolutionized the eld of
engineering. The rst geodesic structure was constructed in 1949 [14].
DEFINITION 1 A Buckyball is a polyhedron which has the following two properties:
i.) Every face consists of either a pentagon or a hexagon.
ii.) Every vertex is of degree 3.
It can be proven, that EVERY buckyball contains 12 pentagons no matter how large or small of
a model we construct. However, we can construct a model with numerous hexagons.
If those 12 pentagons are evenly distributed across the surface, it is then called a Spherical
Buckyball [7].
These models are constructed utilizing a unit designed by Thomas Hull of Merrimack College. He calls
the unit a Pentagon-Hexagon-Zig-Zag, or PHiZZ unit. The smallest structure which can be constructed with
this unit is a dodecahedron made with 30 PHiZZ units. However, by utilizing 90 PHiZZ units we see the
formation of a truncated icosahedron, also known as a Buckminster fullerene or a C60 molecule [7].
Figure 24: Example of a Buckyball constructed with 30 PHiZZ units.
4.2 Sonobe Units
The Sonobe Unit was created by Mitsubobu Sonobe and Kunihiko Kasahara. The unit itself consists of two
tabs and two pockets. This unit is interesting in that it allows for us to cap any polyhedron with triangular
faces. They are capped with a pyramid constructed from three interlocking Sonobe units [6].
Each unit contributes to the construction of two pyramids. Therefore, for every three units used, two
complete pyramids can be formed. So, a model made with 12 units, constructs 8 pyramids which cap a
octahedron.
QUESTION 7 What does a model constructed with 30 units cap?
Since every three units construct two pyramids, we rst divide 30 by 3 to get 10 and then
multiply this by 2 to get 20. Therefore, the model constructed by 30 Sonobe units is a capped
icosahedron.
Figure 25: Capped icosahedron (left) and a capped octahedron (right).
QUESTION 8 What does a model made with 6 units look like?
Using the same method as above, we take 6 and divide it by 3 to get 2. Multiplying by 2 gives us
4, which is a tetrahedron. Notice, however, that a capped tetrahedron results in the formation
of a Cube.
Figure 26: A capped tetrahedron, aka a cube.
4.3 Fractals
Fractals are complex in nature and can be dicult to model. By utilizing modular origami in the study
of fractals, mathematicians are able to create 2 and 3-Dimensional representations of various fractals. The
Menger Sponge, for example, can be easily represented by constructing cubes and interlocking them to form
each iteration of the model. One can see that it would be dicult (but not impossible), not to mention
time consuming, to construct anything larger than a level three Menger Sponge. By using business cards, a
moderate size level one model is constructed measuring approximately 6
1
4
inches cubed. Such a model takes
192 business cards.
Figure 27: Level one Menger Sponge constructed with 192 business cards.
It is also possible to construct a Sierpinski tetrahedron. An example of a 2-Dimensional representation
of a fractal model would be the Koch Snowake or the Sierpinski triangle [10]. Modular origami allows us
to represent numerous other fractal models which may be dicult or even impossible in other mediums.
4.4 Buttery Bombs
Building a bomb is probably not your idea of fun; at least I hope not. Building Buttery Bombs on the
other hand can be fun and is not dangerous OR illegal! A buttery bomb is a cuboctahedron (6 square faces
and 8 triangular faces) constructed in such a way that the units do not lock. In other words, the model is
unstable. Most modular origami models are constructed with the use of a locking mechanism; holding the
units in place. The Buttery Bombs triangular faces are actually triangular cavities [7].
The unstable nature of the Buttery Bomb makes it a fun puzzle to put together and even more fun to
take apart. Construction of the model can prove to be dicult, especially if you have small or uncoordinated
hands. Each piece of the model must be gently weaved into place. The last piece must be in place, before
the model will stay together. Construction of the model is only half the fun. By throwing the model into
the air and hitting it with your palm, a beautiful explosion occurs, similar to that of your local rework
show.
Figure 28: Small and large versions of the Buttery Bomb.
It is also possible to construct a bomb from the Buttery Bombs Dual. The capped octahedron is
similar to that constructed by the Sonobe unit in Section 4.2. This model is constructed in the same weaving
fashion as the Buttery Bomb giving it the same unstable quality. Another interesting quality possessed by
this model is its ability to tessellate 3-dimensional space. The pyramid of the Dual ts perfectly into the
triangular cavity of the Buttery Bomb [7].
Figure 29: A Buttery Bomb (bottom) with a Buttery Bomb Dual (top).
4.5 Five Intersecting Tetrahedrons (FIT)
One of the most fascinating models is that of the Five Intersecting Tetrahedrons, or FIT. The symmetry is
what gives this model its complex and unique characteristics.
We begin with a dodecahedron. By connecting four equidistant corners we nd the formation of an
inscribed tetrahedron. Since the dodecahedron has twenty corners, it is then possible to inscribe a total of
ve tetrahedrons within it [5].
Figure 30: Formation of the FIT model. (Graphics used with permission from [5] )
In order to represent the intersection of the pyramids, we must construct the framework of each one
and then intertwine them symmetrically. This will require the use of Francis Ows 60

Unit discussed in
Section 3.6, with a slight modication.
After completion of the 60

unit using a 1 x 3 inch paper (forming angles on the short end), open
the two aps. The ap on the left, must be inverted to form a pocket. The right hand ap
must then be folded again, this time to the existing crease. You will then repeat all these steps
on the opposite end of your paper, being sure to preserve the right-handedness of the unit [5].
Figure 31: A modied 60

unit. (Adapted from [5] )


By connecting three units, we form a 180

vertex of one tetrahedron. Another three units will complete


the rst tetrahedron. The remaining pyramids must be constructed within the original to achieve the desired
result.
The nished product is a mind boggling display of beauty.
Figure 32: A completed FIT model constructed with 2 x 6 inch paper.
5 Conclusion
Mathematics and origami can both be considered beautiful forms of artwork in their own unique way. More
powerful and mysterious, however, is the beauty hidden within their unique relationship. By exploring these
relationships further, mathematicians stand to learn more about the world we live in, while the artists will
create new ways to represent lifes beauty. There is much to be learned from mathematical origami and much
more to be discovered. They may be considered ancient arts, but it has taken modern day mathematicians
(and artists of course) to unite them for eternity.
References
[1] R. Alperin, A mathematical theory of origami constructions and numbers, New York Journal of
Mathematics 6 (2000), 119-133 pp.
[2] K. Burczyk, Polyhedra classication, Krystyna Burczyks Origami Gallery - Regular Polyhedra,
http://www1.zetosa.com.pl/burczyk/index-en.html
[3] K. Hatori, Origami construction, Ks Origami , http://origami.ousaan.com/library/conste.html
[4] T. Hull, A comparison between straight edge and compass constructions and origami, Origami and
Geometric Constructions, http://kahuna.merrimack.edu/thull/omles/geoconst.html
[5] , Five Intersecting Tetrahedra, Origami and Geometric Constructions,
http://www.merrimack.edu/thll/t.html
[6] , Handouts for using origami in undergraduate math classes; MAA minicourse, Joint Meetings,
New Orleans, LA , 2007.
[7] , Project Origami; Activities for Exploring Mathematics, A K Peters, Wellesley, 2006.
[8] T. Ishida, Kusudama, http://www.ask.ne.jp/kanzasi/en/e-kusu.html
[9] L. Kinsey and T. Moore, Symmetry, Shape, and Space; An Introduction to Mathematics Through
Geometry, Key College Publishing, Emeryville, 2002.
[10] M. Kosmulski, Fractals; IFS, Modular Origami , http://hektor.umcs.lublin.pl/mikosmul/origami/fractals.html
[11] D. Lister, Humiaki Huzita, The History of Origami ,
http://www.britishorigami.info/academic/lister/humiaki huzita.htm
[12] J. OConnor and E. Robertson, Doubling the cube,
http://www-groups.dcs.st-and.ac.uk/history/HistTopics/Doubling the cube.html
[13] T. Row, Geometric Exercises in Paper Folding, Dover Publications, New York, 1966.
[14] Wikipedia, Buckminster Fuller, Wikipedia, The Free Encyclopedia,
http://en.wikipedia.org/w/index.php?title=Buckminster Fuller&oldid= 125716544
[15] , Modular origami, Wikipedia, The Free Encyclopedia,
http://en.wikipedia.org/w/index.php?title=Modular origami&oldid= 122835057
[16] , Origami, Wikipedia, The Free Encyclopedia,
http://en.wikipedia.org/w/index.php?title=Origami&oldid=126184321

Vous aimerez peut-être aussi