Singh KNP Concepts in Coordinate Geometry
Singh KNP Concepts in Coordinate Geometry
Singh KNP Concepts in Coordinate Geometry
Concepts in Cordinate
Geometry
Preface
1. Coordinate System
2. Alternate Viewpoints
3. Polar Coordinate System
4. Homogeneous Coordinates
5. Affine Coordinate System
6. Orthogonal Coordinates
7. Quadratic Equation
8. Notation and Terminology
9. Topological and Analytic Structure
Bibliography
Preface
—Editor
1: Coordinate System
Number Line
In basic mathematics, a number line is a picture of a straight line on
which every point is assumed to correspond to a real number and every real
number to a point. Often the integers are shown as specially-marked points
evenly spaced on the line. Although this image only shows the integers from
-9 to 9, the line includes all real numbers, continuing forever in each
direction, and also numbers not marked that are between the integers. It is
often used as an aid in teaching simple addition and subtraction, especially
involving negative numbers.
It is divided into two symmetric halves by the origin, i.e. the number
zero.
In advanced mathematics, the expressions real number line, or real line
are typically used to indicate the above-mentioned concept that every point
on a straight line corresponds to a single real number, and vice versa.
Drawing the Number Line
The number line is usually represented as being horizontal. Customarily,
positive numbers lie on the right side of zero, and negative numbers lie on the
left side of zero. An arrowhead on either end of the drawing is meant to
suggest that the line continues indefinitely in the positive and negative real
numbers, denoted by ℝ. The real numbers consist of irrational numbers and
rational numbers, as well as the integers, whole numbers, and the natural
numbers (the counting numbers).
A line drawn through the origin at right angles to the real number line
can be used to represent the imaginary numbers. This line, called imaginary
line, extends the number line to a complex number plane, with points
representing complex numbers.
Figure: Illustration of a Cartesian coordinate plane. Four points are marked and labeled with their
coordinates: (2, 3) in green, (-3, 1) in red, (-1.5, - 2.5) in blue, and the origin (0, 0) in purple.
A Cartesian coordinate system specifies each point uniquely in a plane
by a pair of numerical coordinates, which are the signed distances from the
point to two fixed perpendicular directed lines, measured in the same unit of
length. Each reference line is called a coordinate axis or just axis of the
system, and the point where they meet is its origin, usually at ordered pair
(0,0).
The coordinates can also be defined as the positions of the perpendicular
projections of the point onto the two axes, expressed as signed distances from
the origin.
One can use the same principle to specify the position of any point in
three-dimensional space by three Cartesian coordinates, its signed distances
to three mutually perpendicular planes (or, equivalently, by its perpendicular
projection onto three mutually perpendicular lines).
In general, one can specify a point in a space of any dimension n by use
of n Cartesian coordinates, the signed distances from n mutually
perpendicular hyperplanes.
Figure: Cartesian coordinate system with a circle of radius 2 centreed at the origin marked in red. The
equation of a circle is (x – a)2 + (y – b)2 = r2 where a and b are the coordinates of the centre (a, b) and
r is the radius.
The invention of Cartesian coordinates in the 17th century by René
Descartes (Latinized name: Cartesius) revolutionized mathematics by
providing the first systematic link between Euclidean geometry and algebra.
Using the Cartesian coordinate system, geometric shapes (such as curves) can
be described by Cartesian equations: algebraic equations involving the
coordinates of the points lying on the shape. For example, a circle of radius 2
may be described as the set of all points whose coordinates x and y satisfy the
equation x2 + y2 = 4.
Cartesian coordinates are the foundation of analytic geometry, and
provide enlightening geometric interpretations for many other branches of
mathematics, such as linear algebra, complex analysis, differential geometry,
multivariate calculus, group theory, and more.
A familiar example is the concept of the graph of a function. Cartesian
coordinates are also essential tools for most applied disciplines that deal with
geometry, including astronomy, physics, engineering, and many more.
They are the most common coordinate system used in computer
graphics, computer-aided geometric design, and other geometry-related data
processing.
History
The adjective Cartesian refers t o th e Fr en c h math em at ic ian an d
philosopher René Descartes (who used the name Cartesius in Latin). The
idea of this system was developed in 1637 in writings by Descartes and
independently by Pierre de Fermat, although Fermat also worked in three
dimensions, and did not publish the discovery. Both authors used a single
axis in their treatments and have a variable length measured in reference to
this axis.
The concept of using a pair of axes was introduced in later work by
commentators who were trying to clarify the ideas contained in Descartes’ La
Géométrie.
The development of the Cartesian coordinate system would play an
intrinsic role in the development of the calculus by Isaac Newton and
Gottfried Wilhelm Leibniz. Nicole Oresme, a French cleric and friend of the
dauphin (later to become King Charles V) of the 14th Century, used
constructions similar to Cartesian coordinates well before the time of
Descartes and Fermat.
Many other coordinate systems have been developed since Descartes,
such as the polar coordinates for the plane, and the spherical and cylindrical
coordinates for three-dimensional space.
Definitions
Number Line
Choosing a Cartesian coordinate system for a one-dimensional space—
that is, for a straight line—means choosing a point O of the line (the origin), a
unit of length, and an orientation for the line. An orientation chooses which
of the two half-lines determined by O is the positive, and which is negative;
we then say that the line “is oriented” (or “points”) from the negative half
towards the positive half. Then each point p of the line can be specified by its
distance from O, taken with a + or “ sign depending on which half-line
contains p.
A line with a chosen Cartesian system is called a number line. Every real
number, whether integer, rational, or irrational, has a unique location on the
line. Conversely, every point on the line can be interpreted as a number in an
ordered continuum which includes the real numbers.
Cartesian Coordinates in Two Dimensions
The modern Cartesian coordinate system in two dimensions (also called
a rectangular coordinate system) is defined by an ordered pair of
perpendicular lines (axes), a single unit of length for both axes, and an
orientation for each axis. (Early systems allowed “oblique” axes, that is, axes
that did not meet at right angles.) The lines are commonly referred to as the x
and y-axes where the x-axis is taken to be horizontal and the y-axis is taken to
be vertical.
The point where the axes meet is taken as the origin for both, thus
turning each axis into a number line. For a given point P, a line is drawn
through P perpendicular to the x-axis to meet it at X and second line is drawn
through P perpendicular to the y-axis to meet it at Y. The coordinates of P are
then X and Y interpreted as numbers x and y on the corresponding number
lines. The coordinates are written as an ordered pair (x, y).
The point where the axes meet is the common origin of the two number
lines and is simply called the origin. It is often labeled O and if so then the
axes are called Ox and Oy. A plane with x and y-axes defined is often referred
to as the Cartesian plane or xy plane. The value of x is called the x-coordinate
or abscissa and the value of y is called the y-coordinate or ordinate.
The choices of letters come from the original convention, which is to use
the latter part of the alphabet to indicate unknown values. The first part of the
alphabet was used to designate known values.
In the Cartesian plane, reference is sometimes made to a unit circle or a
unit hyperbola.
Cartesian Coordinates in Three Dimensions
Figure: A three dimensional Cartesian coordinate system, with origin O and axis lines X, Y and Z,
oriented as shown by the arrows. The tick marks on the axes are one length unit apart. The black dot
shows the point with coordinates X = 2, Y = 3, and Z = 4, or (2,3,4).
Choosing a Cartesian coordinate system for a three-dimensional space
means choosing an ordered triplet of lines (axes), any two of them being
perpendicular; a single unit of length for all three axes; and an orientation for
each axis. As in the two-dimensional case, each axis becomes a number line.
The coordinates of a point p are obtained by drawing a line through p
perpendicular to each coordinate axis, and reading the points where these
lines meet the axes as three numbers of these number lines.
Alternatively, the coordinates of a point p can also be taken as the
(signed) distances from p to the three planes defined by the three axes. If the
axes are named x, y, and z, then the x coordinate is the distance from the
plane defined by the y and z axes. The distance is to be taken with the + or “
sign, depending on which of the two half-spaces separated by that plane
contains p. The y and z coordinates can be obtained in the same way from the
(x,z) and (x,y) planes, respectively.
Figure: The coordinate surfaces of the Cartesian coordinates (x, y, z). The z-axis is vertical and the x-
axis is highlighted in green. Thus, the red plane shows the points with x=1, the blue plane shows the
points with z=1, and the yellow plane shows the points with y=-1. The three surfaces intersect at the
point P (shown as a black sphere) with the Cartesian coordinates (1, -1, 1).
Generalizations
One can generalize the concept of Cartesian coordinates to allow axes
that are not perpendicular to each other, and/or different units along each
axis. In that case, each coordinate is obtained by projecting the point onto one
axis along a direction that is parallel to the other axis (or, in general, to the
hyperplane defined by all the other axes). In those oblique coordinate systems
the computations of distances and angles is more complicated than in
standard Cartesian systems, and many standard formulas (such as the
Pythagorean formula for the distance) do not hold.
Notations and Conventions
The Cartesian coordinates of a point are usually written in parentheses
and separated by commas, as in (10,5) or (3,5,7). The origin is often labelled
with the capital letter O. In analytic geometry, unknown or generic
coordinates are often denoted by the letters x and y on the plane, and x, y, and
z in three-dimensional space. w is often used for four-dimensional space, but
the rarity of such usage precludes concrete convention here. This custom
comes from an old convention of algebra, to use letters near the end of the
alphabet for unknown values (such as were the coordinates of points in many
geometric problems), and letters near the beginning for given quantities.
These conventional names are often used in other domains, such as
physics and engineering. However, other letters may be used too. For
example, in a graph showing how a pressure varies with time, the graph
coordinates may be denoted t and P. Each axis is usually named after the
coordinate which is measured along it; so one says the x-axis, the y-axis, the
t-axis, etc.
Another common convention for coordinate naming is to use subsc rip ts,
as in x1, x2, ... xn for th e n coordinates in an n-dimensional sp ac e; espec
ially wh en n is greater than 3, or variab le. Some authors (and many
programmers) prefer the numbering x0, x1, ... xn-1. These notations are
especially advantageous in computer programming: by storing the
coordinates of a point as an array, instead of a record, one can use iterative
commands or procedure parametres instead of repeating the same commands
for each coordinate.
In mathematical illustrations of two-dimensional Cartesian systems, the
first coordinate (traditionally called the abscissa) is measured along a
horizontal axis, oriented from left to right. The second coordinate (the
ordinate) is then measured along a vertical axis, usually oriented from bottom
to top.
However, in computer graphics and image processing one often uses a
coordinate system with the y axis pointing down (as displayed on the
computer’s screen). This convention developed in the 1960s (or earlier) from
the way that images were originally stored in display buffers.
For three-dimensional systems, the z axis is often shown vertical and
pointing up (positive up), so that the x and y axes lie on a horizontal plane. If
a diagram (3D projection or 2D perspective drawing) shows the x and y axis
horizontally and vertically, respectively, then the z axis should be shown
pointing “out of the page” towards the viewer or camera. In such a 2D
diagram of a 3D coordinate system, the z axis would appear as a line or ray
pointing down and to the left or down and to the right, depending on the
presumed viewer or camera perspective. In any diagram or display, the
orientation of the three axes, as a whole, is arbitrary. However, the orientation
of the axes relative to each other should always comply with the right-hand
rule, unless specifically stated otherwise. All laws of physics and math
assume this right-handedness, which ensures consistency. For 3D diagrams,
the names “abscissa” and “ordinate” are rarely used for x and y, respectively.
When they are, the z-coordinate is sometimes called the applicate.
The words abscissa, ordinate and applicate are sometimes used to refer
to coordinate axes rather than values.
Octant (Solid Geometry)
An octant in solid geometry is one of the eight divisions of a Euclidean
three-dimensional coordinate system defined by the signs of the coordinates.
It is similar to the two-dimensional quadrant and the one-dimensional ray.
The generalization of an octant is called orthant.
Numbering]
Figure: For z > 0, the octants have the same numbers as the corresponding quadrants in the plane.
Usually, the octant with all three positive coordinates is referred to as the
first octant. There is no generally used naming convention for the other seven
octants.
Number Name x y z Octal Octal
(+=0) (+=1)
I top-front-right + + + 0 7
II top-back-right – + + 4 3
III top-back-left – – + 6 1
IV top-front-left + – + 2 5
V bottom-front- + + – 1 6
right
VI bottom-back- – + – 5 2
right
VII bottom-back- – – – 7 0
left
VIII bottom-front- + – – 3 4
left
[Note the use of row vectors for point coordinates and that the matrix is
written on the right.]
To be orthogonal, the matrix A must have orthogonal rows with same
Euclidean length of one, that is,
This is equivalent to saying that A times its transpose must be the identity
matrix. If these conditions do not hold, the formula describes a more general
affine transformation of the plane provided that the determinant of A is not
zero.
The formula defines a translation if and only if A is the identity matrix.
The transformation is a rotation around some point if and only if A is a
rotation matrix, meaning that
A11 A22 - A21A12 = 1.
A reflection or glide reflection is obtained when,
A11 A22 - A21A12 = -1.
Assuming that translation is not used transformations can be combined
by simply multiplying the associated transformation matrices.
Affine Transformation
Another way to represent coordinate transformations in Cartesian
coordinates is through affine transformations.
In affine transformations an extra dimension is added and all points are
given a value of 1 for this extra dimension.
The advantage of doing this is that then all of the euclidean
transformations become linear transformations and can be represented using
matrix multiplication.
The affine transformation is given by:
[Note the A matrix from above was transposed. The matrix is on the left
and column vectors for point coordinates are used.]
Using affine transformations multiple different euclidean transformations
including translation can be combined by simply multiplying the
corresponding matrices.
Scaling
An example of an affine transformation which is not a Euclidean motion
is given by scaling.
To make a figure larger or smaller is equivalent to multiplying the
Cartesian coordinates of every point by the same positive number m.
If (x,y) are the coordinates of a point on the original figure, the
corresponding point on the scaled figure has coordinates
(x′,y′) = (mx,my).
If m is greater than 1, the figure becomes larger; if m is between 0 and 1,
it becomes smaller.
Shearing
A shearing transformation will push the top of a square sideways to form
a parallelogram.
Horizontal shearing is defined by:
(x′,y′) = (x + ys,y)
Shearing can also be applied vertically:
(x′,y′) = (x,xs + y)
Figure: The left-handed orientation is shown on the left, and the right-handed on the right.
In mathematics, orientation is a geometric notion that in two dimensions
allows one to say when a cycle goes around clockwise or counterclockwise,
and in three dimensions when a figure is left-handed or right-handed. In
linear algebra, the notion of orientation makes sense in arbitrary dimensions.
In this setting, the orientation of an ordered basis is a kind of asymmetry
that makes a reflection impossible to replicate by means of a simple rotation.
Thus, in three dimensions, it is impossible to make the left hand of a human
figure into the right hand of the figure by applying a rotation alone, but it is
possible to do so by reflecting the figure in a mirror.
As a result, in the three-dimensional Euclidean space, the two possible
basis orientations are called right-handed and left-handed (or right-chiral and
left-chiral).
The orientation on a real vector space is the arbitrary choice of which
ordered bases are “positively” oriented and which are “negatively” oriented.
In the three-dimensional Euclidean space, right-handed bases are
typically declared to be positively oriented, but the choice is arbitrary, as they
may also be assigned a negative orientation.
A vector space with an orientation is called an oriented vector space,
while one without a choice of orientation is called unoriented.
Definition
Let V be a finite-dimensional real vector space and let b 1 and b 2 be two
ordered bases for V. It is a standard result in linear algebra that there exists a
unique linear transformation A : V → V that takes b 1 to b2. The bases b 1 and
b 2 are said to have the same orientation (or be consistently oriented) if A has
positive determinant; otherwise they have opposite orientations.
The property of having the same orientation defines an equivalence
relation on the set of all ordered bases for V. If V is non-zero, there are
precisely two equivalence classes determined by this relation. An orientation
on V is an assignment of +1 to one equivalence class and -1 to the other.
Every ordered basis lives in one equivalence class or another. Thus any
choice of a privileged ordered basis for V determines an orientation: the
orientation class of the privileged basis is declared to be positive.
For example, the standard basis on Rn provides a standard orientation on
Rn (in turn, the orientation of the standard basis depends on the orientation of
the Cartesian coordinate system on which it is built). Any choice of a linear
isomorphism between V and Rn will then provide an orientation on V.
The ordering of elements in a basis is crucial. Two bases with a different
ordering will differ by some permutation. They will have the same/opposite
orientations according to whether the signature of this permutation is ±1.
This is because the determinant of a permutation matrix is equal to the
signature of the associated permutation.
Similarly, let A be a nonsingular linear mapping of vector space Rn to
Rn. This mapping is orientation-preserving if its determinant is positive. For
instance, in R3 a rotation around the Z Cartesian axis by an angle α is
orientation-preserving:
Zero-dimensional Case
The concept of orientation defined above did not quite apply to zero-
dimensional vector spaces (as the only empty matrix is the identity (with
determinant 1), so there will be only one equivalence class). However, it is
useful to be able to assign different orientations to a point (e.g. orienting the
boundary of a 1-dimensional manifold).
A more general definition of orientation that works regardless of
dimension is the following:
An orientation on V is a map from the set of ordered bases of V to the set
{±1} that is invariant under base changes with positive determinant and
changes sign under base changes with negative determinant (it is equivarient
with respect to the homomorphism GLn → ±1 ). The set of ordered bases of
the zero-dimensional vector space has one element (the empty set), and so
there are two maps from this set to {±1} .
A subtle point is that a zero-dimensional vector space is naturally
(canonically) oriented, so we can talk about an orientation being positive
(agreeing with the canonical orientation) or negative (disagreeing). An
application is interpreting the Fundamental theorem of calculus as a special
case of Stokes’ theorem.
Two ways of seeing this are:
• A zero-dimensional vector space is a point, and there is a unique map
from a point to a point, so every zero-dimensional vector space is
naturally identified with R0, and thus is oriented.
• The 0th exterior power of a vector space is the ground field , which
here is R1, which has an orientation (given by the standard basis).
2: Alternate Viewpoints
Multilinear Algebra
For any n-dimensional real vector space V we can form the k th-exterior
power of V, denoted ΛkV. This is a real vector space of dimension . The
vector space ΛnV (called the top exterior power) therefore has dimension 1.
That is, Λ nV is just a real line. There is no a priori choice of which direction
on this line is positive. An orientation is just such a choice. Any nonzero
linear form ω on ËnV determines an orientation of V by declaring that x is in
the positive direction when ω(x) > 0. To connect with the basis point of view
we say that the positively oriented bases are those on which ω evaluates to a
positive number (since ω is an n-form we can evaluate it on an ordered set of
n vectors, giving an element of R). The form ω is called an orientation form.
If {ei} is a privileged basis for V and {e*} is the dual basis, then the
orientation form giving the standard orientation is
The connection of this with the determinant point of view is: the
determinant of an endomorphism T:V^V can be interpreted as the induced
action on the top exterior power.
Orientation on Manifolds
One can also discuss orientation on manifolds. Each point p on an n-
dimensional differentiable manifold has a tangent space T M which is an n-
dimensional real vector space. One can assign to each of these vector spaces
an orientation. However, one would like to know whether it is possible to
choose the orientations so that they “vary smoothly” from point to point. Due
to certain topological restrictions, there are situations when this is impossible.
A manifold which admits a smooth choice of orientations for its tangents
spaces is said to be orientable.
Axes Conventions
Figure: Heading, elevation and bank angles (Z-Y’-X’’) for an aircraft. The aircraft’s pitch and yaw
axes Y and Z are not shown, and its fixed reference frame xyz has been shifted backwards from its
centre of gravity (preserving angles) for clarity. Axes named according to the air norm DIN 9300
Mobile objects are normally tracked from an external frame considered
fixed. Other frames can be defined on those mobile objects to deal with
relative positions for other objects. Finally, attitudes or orientations can be
described by a relationship between the external frame and the one defined
over the mobile object.
The orientation of a vehicle is normally referred to as attitude. It is
described normally by the orientation of a frame fixed in the body relative to
a fixed reference frame. The attitude is described by attitude coordinates, and
consists of at least three coordinates.
While from a geometrical point of view the different methods to describe
orientations are defined using only some reference frames, in engineering
applications it is important also to describe how these frames are attached to
the lab and the body in motion.
Due to the special importance of international conventions in air
vehicles, several organizations have published standards to be followed. For
example, German DIN has published the DIN 9300 norm for aircraft
(adopted by ISO as ISO 1151–2:1985).
Ground Reference Frames: ENU and NED
Basically, as lab frame or reference frame, there are two kinds of
conventions for the frames (sometimes named LVLH, local vertical, local
horizontal):
• East, North, Up, referred as ENU
• North, East, Down, referred as NED, used specially in aerospace
These frames are location dependent. For movements around the globe,
like air or sea navigation, the frames are defined as tangent to the lines of
coordinates.
• East-West tangent to parallels,
• North-South tangent to meridians, and
• Up-Down in the direction to the centre of the earth (when using a
spherical Earth simplification), or in the direction normal to the local
tangent plane (using an oblate spheroidal or geodetic ellipsoidal
model of the earth) which does not generally pass through the centre
of the Earth.
• a With the thumb, index, and middle fingers at right angles to each
other (with the index finger pointed straight), the middle finger points
in the direction of c when the thumb represents a and the index finger
represents b.
Other (equivalent) finger assignments are possible. For example, the first
(index) finger can represent a, the first vector in the product; the second
(middle) finger, b, the second vector; and the thumb, c, the product.
Direction Associated with a Rotation
Figure: Prediction of direction of field (B), given that the current I flows in the direction of the thumb
Figure: The right-hand rule as applied to motion produced with screw threads
This version of the rule is used in two complementary applications of
Ampère’s circuital law:
1. An electric current passes through a solenoid, resulting in a magnetic
field. When you wrap your right hand around the solenoid with your
fingers in the direction of the conventional current, your thumb points
in the direction of the magnetic north pole.
2. An electric current passes through a straight wire. Here, the thumb
points in the direction of the conventional current (from positive to
negative), and the fingers point in the direction of the magnetic lines
of flux.
The principle is also used to determine the direction of the torque vector.
If you grip the imaginary axis of rotation of the rotational force so that your
fingers point in the direction of the force, then the extended thumb points in
the direction of the torque vector. The right-hand grip rule is a convention
derived from the right-hand rule convention for vectors. When applying the
rule to current in a straight wire for example, the direction of the magnetic
field (counterclockwise instead of clockwise when viewed from the tip of the
thumb) is a result of this convention and not an underlying physical
phenomenon.
Applications
The first form of the rule is used to determine the direction of the cross
product of two vectors. This leads to widespread use in physics, wherever the
cross product occurs. A list of physical quantities whose directions are related
by the right-hand rule is given below. (Some of these are related only
indirectly to cross products, and use the second form.)
• The angular velocity of a rotating object and the rotational velocity of
any point on the object
• A torque, the force that causes it, and the position of the point of
application of the force
• A magnetic field, the position of the point where it is determined, and
the electric current (or change in electric flux) that causes it
• A magnetic field in a coil of wire and the electric current in the wire
• The force of a magnetic field on a charged particle, the magnetic field
itself, and the velocity of the object
• The vorticity at any point in the field of flow of a fluid
• The induced current from motion in a magnetic field (known as
Fleming’s right-hand rule)
• The x, y and z unit vectors in a Cartesian coordinate system can be
chosen to follow the right-hand rule. Right-handed coordinate
systems are often used in rigid body physics and kinematics.
Figure: Fleming’s left-hand rule
Fleming’s left-hand rule is a rule for finding the direction of the thrust on
a conductor carrying a current in a magnetic field.
Left-hand Rule
In certain situations, it may be useful to use the opposite convention,
where one of the vectors is reversed and so creates a left-handed triad instead
of a right-handed triad. An example of this situation is for left-handed
materials.
Normally, for an electromagnetic wave, the electric and magnetic fields,
and the direction of propagation of the wave obey the right-hand rule.
However, left-handed materials have special properties, notably the negative
refractive index. It makes the direction of propagation point in the opposite
direction.
De Graaf’s translation of Fleming’s left-hand rule - which uses thrust,
field and current - and the right-hand rule, is the FBI rule. The FBI rule
changes thrust into F (Lorentz force), B (direction of the magnetic field) and I
(current).
The FBI rule is easily remembered by US citizens because of the
commonly known abbreviation for the Federal Bureau of Investigation.
Symmetry
In Two Dimensions
Fixing or choosing the x-axis determines the y-axis up to direction.
Namely, the y-axis is necessarily the perpendicular to the x-axis through the
point marked 0 on the x-axis. But there is a choice of which of the two half
lines on the perpendicular to designate as positive and which as negative.
Each of these two choices determines a different orientation (also called
handedness) of the Cartesian plane.
The usual way of orienting the axes, with the positive x-axis pointing
right and the positive y-axis pointing up (and the x-axis being the “first” and
the y-axis the “second” axis) is considered the positive or standard
orientation, also called the right-handed orientation.
A commonly used mnemonic for defining the positive orientation is the
right hand rule. Placing a somewhat closed right hand on the plane with the
thumb pointing up, the fingers point from the x-axis to the y-axis, in a
positively oriented coordinate system. The other way of orienting the axes is
following the left hand rule, placing the left hand on the plane with the thumb
pointing up.
When pointing the thumb away from the origin along an axis, the
curvature of the fingers indicates a positive rotation along that axis.
Regardless of the rule used to orient the axes, rotating the coordinate system
will preserve the orientation. Switching any two axes will reverse the
orientation.
In Three Dimensions
Once the x- and y-axes are specified, they determine the line along which
the z-axis should lie, but there are two possible directions on this line. The
two possible coordinate systems which result are called ‘right-handed’ and
‘left-handed’. The standard orientation, where the xy-plane is horizontal and
the z-axis points up (and the x- and the y-axis form a positively oriented two-
dimensional coordinate system in the xy-plane if observed from above the xy-
plane) is called right-handed or positive.
The name derives from the right-hand rule. If the index finger of the
right hand is pointed forward, the middle finger bent inward at a right angle
to it, and the thumb placed at a right angle to both, the three fingers indicate
the relative directions of the x-, y-, and z-axes in a right-handed system. The
thumb indicates the x-axis, the index finger the y-axis and the middle finger
the z-axis. Conversely, if the same is done with the left hand, a left-handed
system results.
Figure 7 depicts a left and a right-handed coordinate system. Because a
three-dimensional object is represented on the two-dimensional screen,
distortion and ambiguity result. The axis pointing downward (and to the
right) is also meant to point towards the observer, whereas the “middle” axis
is meant to point away from the observer. The red circle is parallel to the
horizontal xy-plane and indicates rotation from the x-axis to the y-axis (in
both cases). Hence the red arrow passes in front of the z-axis. Figure is
another attempt at depicting a right-handed coordinate system. Again, there is
an ambiguity caused by projecting the three-dimensional coordinate system
into the plane. This corresponds to the two possible orientations of the
coordinate system. Thus the “correct” way to view is to imagine the x-axis as
pointing towards the observer and thus seeing a concave corner.
Representing a Vector in the Standard Basis
A point in space in a Cartesian coordinate system may also be
represented by a position vector, which can be thought of as an arrow
pointing from the origin of the coordinate system to the point. If the
coordinates represent spatial positions (displacements), it is common to
represent the vector from the origin to the point of interest as r. In two
dimensions, the vector from the origin to the point with Cartesian coordinates
(x, y) can be written as:
r = xi+yj
where i = (1,0) , and j = (0,1) are unit vectors in the direction of the x-axis
and y-axis respectively, generally referred to as the standard basis (in some
application areas these may also be referred t o a s ve r so r s ). S im i l ar l y,
in three d imensions, the vec tor fr om the origin to the point with Cartesian
coordinates ( x, y, z) can be written as:
r = xi+yj+zk
where k = (0,0,1) is the unit vector in the direction of the z-axis.
There is no natural interpretation of multiplying vectors to obtain
another vector that works in all dimensions, however there is a way to use
complex numbers to provide such a multiplication. In a two dimensional
cartesian plane, identify the point with coordinates (x, y) with the complex
number z = x + iy.
Here, i is the complex number whose square is the real number “1 and is
identified with the point with coordinates (0,1), so it is not the unit vector in
the direction of the x-axis (this confusion is just an unfortunate historical
accident).
Since the complex numbers can be multiplied giving another complex
number, this identification provides a means to “multiply” vectors. In a three
dimensional cartesian space a similar identification can be made with a subset
of the quaternions.
Applications
Each axis may have different units of measurement associated with it
(such as kilograms, seconds, pounds, etc.). Although four-and higher-
dimensional spaces are difficult to visualize, the algebra of Cartesian
coordinates can be extended relatively easily to four or more variables, so
that certain calculations involving many variables can be done. (This sort of
algebraic extension is what is used to define the geometry of higher-
dimensional spaces.) Conversely, it is often helpful to use the geometry of
Cartesian coordinates in two or three dimensions to visualize algebraic
relationships between two or three of many non-spatial variables.
The graph of a function or relation is the set of all points satisfying that
function or relation. For a function of one variable, f, the set of all points (x,y)
where y = f(x) is the graph of the function f. For a function of two variables,
g, the set of all points (x,y,z) wher e z = g(x,y) is the graph of the function g.
A sketch of the graph of such a function or relation would consist of all the
salient parts of the function or relation which would include its relative
extrema, its concavity and points of inflection, any points of discontinuity
and its end behaviour. All of these terms are more fully defined in calculus.
Such graphs are useful in calculus to understand the nature and behaviour of
a function or relation.
Note that positions on a surface in navigation use latitude and longitude
in a similar two dimensional system. However the co-ordinates are written in
the opposite sequence, effectively (y,x).
3: Polar Coordinate System
Figure: Points in the polar coordinate system with pole O and polar axis L. In green, the point with
radial coordinate 3 and angular coordinate 60 degrees, or (3,60°). In blue, the point (4,210°).
History
The concepts of angle and radius were already used by ancient peoples
of the 1st millennium BCE. The Greek astronomer and astrologer Hipparchus
(190–120 BCE) created a table of chord functions giving the length of the
chord for each angle, and there are references to his using polar coordinates
in establishing stellar positions. In On Spirals, Archimedes describes the
Archimedean spiral, a function whose radius depends on the angle. The
Greek work, however, did not extend to a full coordinate system.
From the 8th century CE onward, astronomers developed methods for
approximating and calculating the direction to Makkah (qibla)—and its
distance—from any location on the Earth. From the 9th century onward they
were using spherical trigonometry and map projection methods to determine
these quantities accurately. The calculation is essentially the conversion of
the equatorial polar coordinates of Mecca (i.e. its longitude and latitude) to its
polar coordinates (i.e. its qibla and distance) relative to a system whose
reference meridian is the great circle through the given location and the
Earth’s poles, and whose polar axis is the line through the location and its
antipodal point.
There are various accounts of the introduction of polar coordinates as
part of a formal coordinate system. The full history of the subject is described
in Harvard professor Julian Lowell Coolidge’s Origin of Polar Coordinates.
Grégoire de Saint-Vincent and Bonaventura Cavalieri independently
introduced the concepts in the mid-seventeenth century. Saint-Vincent wrote
about them privately in 1625 and published his work in 1647, while Cavalieri
published his in 1635 with a corrected version appearing in 1653. Cavalieri
first used polar coordinates to solve a problem relating to the area within an
Archimedean spiral. Blaise Pascal subsequently used polar coordinates to
calculate the length of parabolic arcs.
In Method of Fluxions (written 1671, published 1736), Sir Isaac Newton
examined the transformations between polar coordinates, which he referred to
as the “Seventh Manner; For Spirals”, and nine other coordinate systems. In
the journal Acta Eruditorum (1691), Jacob Bernoulli used a system with a
point on a line, called the pole and polar axis respectively. Coordinates were
specified by the distance from the pole and the angle from the polar axis.
Bernoulli’s work extended to finding the radius of curvature of curves
expressed in these coordinates.
The actual term polar coordinates has been attributed to Gregorio
Fontana and was used by 18th-century Italian writers. The term appeared in
English in George Peacock’s 1816 translation of Lacroix’s Differential and
Integral Calculus. Alexis Clairaut was the first to think of polar coordinates
in three dimensions, and Leonhard Euler was the first to actually develop
them.
Conventions
The radial coordinate is often denoted by r, and the angular coordinate
by θ or t. Angles in polar notation are generally expressed in either degrees or
radians (2* rad being equal to 360°) Degrees are traditionally used in
navigation, surveying, and many applied disciplines, while radians are more
common in mathematics and mathematical physics.
In many contexts, a positive angular coordinate means that the angle θ is
measured counterclockwise from the axis. In mathematical literature, the
polar axis is often drawn horizontal and pointing to the right.
Uniqueness of Polar Coordinates
Adding any number of full turns (360°) to the angular coordinate does
not change the corresponding direction. Also, a negative radial coordinate is
best interpreted as the corresponding positive distance measured in the
opposite direction. Therefore the same point can be expressed with an infinite
number of different polar coordinates (r, θ± n×360°) or (-r, θ± (2n + 1)180°),
where n is any integer. Moreover, the pole itself can be expressed as (0, θ) for
any angle θ.
Where a unique representation is needed for any point, it is usual to limit
r to non-negative numbers (r ≥ 0) and θ to the interval [0, 360°) or (-180°,
180°] (in radians, [0, 2n) or (-n, n]). One must also choose a unique azimuth
for the pole, e.g., θ= 0.
the solution with a minus sign in front of the square root giving the same
curve.
Line
Radial lines (those running through the pole) are represented by the
equation
θ=φ
where φ is the angle of elevation of the line; that is, φ = arctan m where
m is the slope of the line in the Cartesian coordinate system.
The non-radial line that crosses the radial line θ = φ perpendicularly at
the point (r0, φ) has the equation
r( θ) =r0sec(θ-φ).
Otherwise stated (r0, φ) is the point in which the tangent intersects the
imaginary circle of radius r0.
Polar Rose
A polar rose is a famous mathematical curve that looks like a petaled
flower, and that can be expressed as a simple polar equation,
r( θ) = acos(kθ + φ0)
for any constant φ0 (including 0).
If k is an integer, these equations will produce a k-petaled rose if k is odd,
or a 2k-petaled rose if k is even.
If k is rational but not an integer, a rose-like shape may form but with
overlapping petals.
Note that these equations never define a rose with 2, 6, 10, 14, etc.
petals. The variable a represents the length of the petals of the rose
Archimedean Spiral
The Archimedean spiral is a famous spiral that was discovered by
Archimedes, which also can be expressed as a simple polar equation. It is
represented by the equation
r(θ) = a + bθ.
Changing the parametre a will turn the spiral, while b controls the
distance between the arms, which for a given spiral is always constant.
The Archimedean spiral has two arms, one for θ > 0 and one for θ < 0.
The two arms are smoothly connected at the pole. Taking the mirror
image of one arm across the 90°/270° line will yield the other arm.
This curve is notable as one of the first curves, after the conic sections, to
be described in a mathematical treatise, and as being a prime example of a
curve that is best defined by a polar equation
Complex Numbers
Figure: A cylindrical coordinate system with origin O, polar axis A, and longitudinal axis L. The dot is
the point with radial distance ρ = 4, angular coordinate ϕ = 130°, and height z = 4.
The origin of the system is the point where all three coordinates can be
given as zero. This is the intersection between the reference plane and the
axis.
The axis is variously called the cylindrical or longitudinal axis, to
differentiate it from the polar axis, which is the ray that lies in the reference
plane, starting at the origin and pointing in the reference direction.
The distance from the axis may be called the radial distance or radius,
while the angular coordinate is sometimes referred to as the angular position
or as the azimuth. The radius and the azimuth are together called the polar
coordinates, as they correspond to a two-dimensional polar coordinate
system in the plane through the point, parallel to the reference plane. The
third coordinate may be called the height or altitude (if the reference plane is
considered horizontal), longitudinal position, or axial position.
Cylindrical coordinates are useful in connection with objects and
phenomena that have some rotational symmetry about the longitudinal axis,
such as water flow in a straight pipe with round cross-section, heat
distribution in a metal cylinder, electromagnetic fields produced by an
electric current in a long, straight wire, and so on
It is sometimes called “cylindrical polar coordinate” and “polar
cylindrical coordinate”, and is sometime used to specify the position of stars
in a galaxy (“galactocentric cylindrical polar coordinate”).
Definition
The three coordinates (ρ φ, z) of a point P are defined as:
• The radial distance ρ is the Euclidean distance from the z axis to the
point P.
• The azimuth pis the angle between the reference direction on the
chosen plane and the line from the origin to the projection of P on the
plane.
• The height z is the signed distance from the chosen plane to the point
P.
Unique Cylindrical Coordinates
As in polar coordinates, the same point with cylindrical coordinates (ρ,
φ, z) has infinitely many equivalent coordinates, namely (ρ φ± n360°, z) and
(-ρ φ± (2n + 1)×180°, z), where n is any integer. Moreover, if the radius ρ is
zero, the azimuth is arbitrary.
In situations where one needs a unique set of coordinates for each point,
one may restrict the radius to be non-negative (ρ ≥ 0) and the azimuth pto lie
in a specific interval spanning 360°, such as (-180°,+180°] or [0,360°).
Coordinate System Conversions
The cylindrical coordinate system is one of many three-dimensional
coordinate systems. The following formula may be used to convert between
them.
Cartesian Coordinates
For the conversion between cylindrical and Cartesian coordinate
systems, it is convenient to assume that the reference plane of the former is
the Cartesian x-y plane (with equation z = 0), and the cylindrical axis is the
Cartesian z axis.
Then the z coordinate is the same in both systems, and the
correspondence between cylindrical (p,φ) and Cartesian (x,y) are the same as
for polar coordinates, namely
in the other. The arcsin function is the inverse of the sine function, and is
assumed to return an angle in the range [- π/2,+π/2] = [-90°,+90°].
These formulas yield an azimuth φin the range [- 90°,+270°]. Many
modern programming languages provide a function that will compute the
correct azimuth φ, in the range (-π, π], given x and y, without the need to
perform a case analysis as above.
For example, this function is called by atan2(y,x) in the C programming
language, and atan(y,x) in Common Lisp.
Cylindrical Harmonics
The solutions to the Laplace equation in a system with cylindrical
symmetry are called cylindrical harmonics.
Geographic Coordinates
To a first approximation, the geographic coordinate system uses
elevation angle (latitude) in degrees north of the equator plane, in the range
-90° ≤ φ≤ 90°, instead of inclination. Latitude is either geocentric latitude,
measured at the Earth’s centre and designated variously by ψ, q, φ’, φ,φ or
geodetic latitude, measured by the observer’s local vertical, and g commonly
designated φ. The azimuth angle (longitude), commonly denoted by t is
measured in degrees east or west from some conventional reference meridian
(most commonly the IERS Reference Meridian), so its domain is -180° ≤ λ ≤
180°. For positions on the Earth or other solid celestial body, the reference
plane is usually taken to be the plane perpendicular to the axis of rotation.
The polar angle, which is 90° minus the latitude and ranges from 0 to
180°, is called colatitude in geography. Instead of the radial distance,
geographers commonly use altitude above some reference surface, which
may be the sea level or “mean” surface level for planets without liquid
oceans. The radial distance r can be computed from the altitude by adding the
mean radius of the planet’s reference surface, which is approximately
6,360±11 km for Earth.
However, modern geographical coordinate systems are quite complex,
and the positions implied by these simple formulae may be wrong by several
kilometres. The precise standard meanings of latitude, longitude and altitude
are currently defined by the World Geodetic System (WGS), and take into
account the flattening of the Earth at the poles (about 21 km) and many other
details.
In astronomy there are a series of spherical coordinate systems that
measure the elevation angle from different fundamental planes. These
reference planes are the observer’s horizon, the celestial equator (defined by
the Earth’s rotation), the plane of the ecliptic (defined by Earth’s orbit around
the sun), and the galactic equator (defined by the rotation of the galaxy).
Kinematics
In spherical coordinates the position of a point is written,
Figure: Rational Bézier curve – polynomial curve defined in homogeneous coordinates (blue) and its
projection on plane – rational curve (red)
Homogeneous coordinates have a range of applications, including
computer graphics and 3D computer vision, where they allow affine
transformations and, in general, projective transformations to be easily
represented by a matrix. If the homogeneous coordinates of a point are
multiplied by a non-zero scalar then the resulting coordinates represent the
same point.
An additional condition must be added on the coordinates to ensure that
only one set of coordinates corresponds to a given point, so the number of
coordinates required is, in general, one more than the dimension of the
projective space being considered. For example, two homogeneous
coordinates are required to specify a point on the projective line and three
homogeneous coordinates are required to specify a point on the projective
plane.
Introduction
The projective plane can be thought of as the Euclidean plane with
additional points, so called points at infinity, added. There is a point at
infinity for each direction, informally defined as the limit of a point that
moves in that direction away from a fixed point. Parallel lines in the
Euclidean plane are said to intersect at a point at infinity corresponding to
their common direction.
A given point (x, y) on the Euclidean plane is identified with two ratios
(X/Z, Y/Z), so the point (x, y) corresponds to the triple (X, Y, Z) = (xZ, yZ, Z)
where Z ≠ 0. Such a triple is a set of homogeneous coordinates for the point
(x, y).
Note that, since ratios are used, multiplying the three homogeneous
coordinates by a common, non-zero factor does not change the point
represented – unlike Cartesian coordinates, a single point can be represented
by infinitely many homogeneous coordinates.
The equation of a line through the point (a, b) may be written l(x - a) +
m(y - b) = 0 where l and m are not both 0. In parametric form this can be
written x = a + mt, y = b - lt. Let Z=1/t, so the coordinates of a point on the
line may be written (a + m/Z, b - l/ Z)=((aZ + m)/Z, (bZ - l)/Z). In
homogeneous coordinates this becomes (aZ + m, bZ - l, Z).
In the limit as t approaches infinity, in other words as the point moves
away from (a, b), Z becomes 0 and the homogeneous coordinates of the point
become (m, -l, 0). So (m, -l, 0) are defined as homogeneous coordinates of the
point at infinity corresponding to the direction of the line l(x - a) + m(y - b) =
0.
To summarize:
• Any point in the projective plane is represented by a triple (X, Y, Z),
called the homogeneous coordinates of the point, where X, Y and Z
are not all 0.
• The point represented by a given set of homogeneous coordinates is
unchanged if the coordinates are multiplied by a common factor.
• Conversely two sets of homogeneous coordinates represent the same
point if and only if one is obtained from the other by multiplying by a
common factor.
• When Z is not 0 the point represented is the point (X/Z, Y/Z) in the
Euclidean plane.
• When Z is 0 the point represented is a point at infinity.
Note that the triple (0, 0, 0) is omitted and does not represent any point.
The origin is represented by (0, 0, 1).
Notation
Some authors use different notations for homogeneous coordinates
which help distinguish them from Cartesian coordinates. The use of colons
instead of commas, for example (x:y:z) instead of (x, y, z), emphasizes that
the coordinates are to be considered ratios. Brackets, as in [x, y, z] emphasize
that multiple sets of coordinates are associated with a single point. Some
authors use a combination of colons and brackets as in [x:y:z].
Homogeneity
Homogeneous coordinates are not uniquely determined by a point, so a
function defined on the coordinates, say f(x, y, z), does not determine a
function defined on points as with Cartesian coordinates. But a condition f(x,
y, z) = 0 defined on the coordinates, as might be used to describe a curve,
determines a condition on points if the function is homogeneous.
Specifically, suppose there is a k such that
Duality as a Mapping
A (plane) duality is a map from a projective plane C = (P,L,I) to its dual
plane C* = (L,P,I*) which preserves incidence. That is, a (plane) duality σ
will map points to lines and lines to points (Pσ = L and Lσ = P) in such a way
that if a point Q is on a line m ( denoted by Q I m) then Q> I* mσ Ô! mσ I Qσ
A (plane) duality which is an isomorphism is called a correlation.[1] The
existence of a correlation means that the projective plane C is self-dual.
In the special case that the projective plane is of the PG(2,K) type, with
K a division ring, a duality is called a reciprocity. These planes are always
self-dual. By the Fundamental theorem of projective geometry a reciprocity is
the composition of an automorphic function of K and a homography. If the
automorphism involved is the identity, then the reciprocity is called a
projective correlation. A correlation of order two (an involution) is called a
polarity. If a correlation φ is not a polarity then φ2 is a nontrivial collineation.
This duality mapping concept can also be extended to higher dimensional
spaces so the modifier “(plane)” can be dropped in those situations.
Higher Dimensional Duality
Duality in the projective plane is a special case of duality for projective
spaces, transformations of PG(n,K) (also denoted by KPn) with K a field, that
interchange objects of dimension r with objects of dimension n - 1 - r ( =
codimension r + 1). That is, in a projective space of dimension n, the points
(dimension 0) are made to correspond with hyperplanes (codimension 1), the
lines joining two points (dimension 1) are made to correspond with the
intersection of two hyperplanes (codimension 2), and so on.
The points of PG(n,K) can be taken to be the nonzero vectors in the (n +
1)-dimensional vector space over K, where we identify two vectors which
differ by a scalar factor. Another way to put it is that the points of n-
dimensional projective space are the lines through the origin in Kn + 1, which
are 1-dimensional vector subspaces. Also the n- vector dimensional
subspaces of Kn + 1 represent the (n - 1)- geometric dimensional hyperplanes
of projective n-space over K.
A nonzero vector u = (u0,u1,...,un) in Kn + 1 also determines an (n - 1) -
geometric dimensional subspace (hyperplane) Hu, by
Hu = {(x0,x1,...,xn) : u0x0 + … + unxn = 0 }.
When a vector u is used to define a hyperplane in this way it shall be
denoted by uH, while if it is designating a point we will use uP. In terms of the
usual dot product, Hu = {xP : uH • xP = 0}. Since K is a field, the dot product
is symmetrical, meaning uH•xP = u0x0 + u1x1 + ... + unxn = x0u0 + x1u1 + ... +
xnun = xH•uP. A reciprocity can be given by uP ↔ Hu between points and
hyperplanes. This extends to a reciprocity between the line generated by two
points and the intersection of two such hyperplanes, and so forth.
In the projective plane, PG(2,K), with K a field we have the reciprocity
given by: points in homogeneous coordinates (a,b,c) ↔ lines with equations
ax + by + cz = 0. In a corresponding projective space, PG(3,K), a reciprocity
is given by: points in homogeneous coordinates (a,b,c,d) ↔ planes with
equations ax + by + cz + dw = 0. This reciprocity would also map a line
determined by two points (a1,b1,c1,d1) and (a2,b2,c2,d2) to the line which is
the intersection of the two planes with equations a1x + b1y + c1z + d1w = 0
and a2x + b2y + c2z + d2w = 0.
Three Dimensions
In a polarity of real projective 3-space, PG(3,R), points correspond to
planes, and lines correspond to lines. By restriction the duality of polyhedra
in solid geometry is obtained, where points are dual to faces, and sides are
dual to sides, so that the icosahedron is dual to the dodecahedron, and the
cube is dual to the octahedron.
Geometric Construction of a Reciprocity
The reciprocity of PG(2,K), with K a field, given by homogeneous
coordinates can also be described geometrically. This uses the model of the
real projective plane which is a “unit sphere with antipodes identifie≤, or
equivalently, the model of li n e s a n d p lan es t h r o ugh t h e or ig in of th e
vec to r sp a c e K3. Associate a line through the origin with the unique plane
through the origin which is perpendicular (orthogonal) to the line. When, in
the model, these lines are considered to be the points and the planes the lines
of the projective plane PG(2,K), this association becomes a reciprocity
(actually a polarity) of the projective plane. The sphere model is obtained by
intersecting the lines and planes through the origin with a unit sphere
centreed at the origin. The lines meet the sphere in antipodal points which
must then be identified to obtain a point of the projective plane, and the
planes meet the sphere in great circles which are thus the lines of the
projective plane.
That this association “preserves” incidence is most easily seen from the
lines and planes model. A point incident with a line in the projective plane
corresponds to a line lying in a plane in the model. Applying the association,
the plane becomes a line through the origin perpendicular to the plane it is
associated with. This image line is perpendicular to every line of the plane
which passes through the origin, in particular the original line (point of the
projective plane). All lines that are perpendicular to the original line at the
origin lie in the unique plane which is orthogonal to the original line, that is,
the image plane under the association. Thus, the image line lies in the image
plane and the association preserves incidence.
Poles and Polars
Figure: Pole and polar with respect to circle O. P = Q’, q is polar of Q, Q is pole of q.
In the Euclidean plane, fix a circle C with centre O and radius r. For each
point P other than O define an image point P so that OP • OP = r2. The
mapping defined by P → P is called inversion with respect to circle C. The
line through P which is perpendicular to the line OP is called the polar of the
point P with respect to circle C. Let m be a line not passing through O. Drop
a perpendicular from O to m, meeting m at the point Q (this is the point of m
that is closest to O).
The image of Q under inversion with respect to C is called the pole of m.
If a point P (different from O) is on a line m (not passing through O) then the
pole of m lies on the polar of P and viceversa. The incidence preserving
process, in which points and lines are transformed into their polars and poles
with respect to C is called reciprocation. In order to turn this process into a
reciprocity, the Euclidean plane (which is not a projective plane) needs to be
expanded to the extended euclidean plane by adding a line at infinity and
points at infinity which lie on this line. In this expanded plane, we define the
polar of the point O to be the line at infinity (and O is the pole of the line at
infinity), and the poles of the lines through O are the points of infinity where,
if a line has slope s (* 0) its pole is the infinite point associated to the parallel
class of lines with slope -1/s. The pole of the x-axis is the point of infinity of
the vertical lines and the pole of the y-axis is the point of infinity of the
horizontal lines.
The construction of a reciprocity based on inversion in a circle given
above can be generalized by using inversion in a conic section (in the
extended real plane). The reciprocities constructed in this manner are
projective correlations of order two, that is, polarities.
Mapping the Sphere onto the Plane
The unit sphere modulo “1 model of the projective plane is isomorphic
(w r.t. incidence properties) to the planar model: the affine plane extended
with a projective line at infinity.
To map a point on the sphere to a point on the plane, let the plane be
tangent to the sphere at some point which shall be the origin of the plane’s
coordinate system (2-D origin). Then construct a line passing through the
centre of the sphere (3-D origin) and the point on the sphere. This line
intersects the plane at a point which is the projection of the point on the
sphere onto the plane (or vice versa).
This projection can be used to define a one-to-one onto mapping
Also, lines in the planar model are projections of great circles of the
sphere. This is so because through any line in the plane pass an infinitude of
different planes: one of these planes passes through the 3-D origin, but a
plane passing through the 3-D origin intersects the sphere along a great circle.
As we have seen, any great circle in the unit sphere has a projective point
perpendicular to it, which can be defined as its dual. But this point is a pair of
antipodal points on the unit sphere, through both of which passes a unique 3-
D line, and this line extended past the unit sphere intersects the tangent plane
at a point, which means that there is a geometric way to associate a unique
point on the plane to every line on the plane, such that the point is the dual of
the line.
Plücker Coordinates
In geometry, Plücker coordinates, introduced by Julius Plücker in the
19th century, are a way to assign six homogenous coordinates to each line in
projective 3-space, P3. Because they satisfy a quadratic constraint, they
establish a one-to-one correspondence between the 4-dimensional space of
lines in P3 and points on a quadric in P5 (projective 5-space). A predecessor
and special case of Grassmann coordinates (which describe k-dimensional
linear subspaces, or flats, in an n-dimensional Euclidean space), Plücker
coordinates arise naturally in geometric algebra. They have proved useful for
computer graphics, and also can be extended to coordinates for the screws
and wrenches in the theory of kinematics used for robot control.
Geometric Intuition
Therefore, the determinant of the left side 2×2 matrix equals the product
of the determinants of the right side 2×2 matrices, the latter of which is a
fixed scalar, det Λ .
Primary Coordinates
With this motivation, we define Plücker coordinate pij as the determinant
of rows i and j of M,
This implies pii = 0 and pij = -pji, reducing the possibilities to only six (4
choose 2) independent quantities. As we have seen, the sixtuple
(p01 : p02 : p03 : p23 : p31 : p12 )
is uniquely determined by L, up to a common nonzero scale factor.
Furthermore, all six components cannot be zero, because if they were, all
2×2 subdeterminants in M would be zero and the rank of M at most one,
contradicting the assumption that x and y are distinct.
Thus the Plücker coordinates of L, as suggested by the colons, may be
considered homogeneous coordinates of a point in a 5-dimensional projective
space.
Plücker Map
Denote the set of all lines (linear images of P1) in P3 by G1,3. We thus
have a map :
Dual Coordinates
Alternatively, let L be a line contained in distinct planes a and b with
homogeneous coefficients (a0:a1:a2:a3) and (b0:b1:b2:b3), respectively. (The
first plane equation is 0 = -k akxk, for example.) Let N be the 2×4 matrix with
these coordinates as rows.
Geometry
To relate back to the geometric intuition, take x0 = 0 as the plane at
infinity; thus the coordinates of points not at infinity can be normalized so
that x0 = 1. Then M becomes
has rank 2, and so its columns are distinct points defining a line L. When the
P5 coordinates, qij, satisfy the quadratic Plücker relation, they are the Plücker
coordinates of L. To see this, first normalize q 01 to 1. Then we immediately
have that for the Plücker coordinates computed from M, pij = qij, except for
p23 = – q03 q12 – q02 q31 .
But if the qij satisfy the Plücker relation q+qq 3+qq 1 = 0, then p = q2,
completing the set of identities Consequently, α is a surjection onto the
algebraic variety consisting of the set of zeros of the quadratic polynomial
p01 p23 +p02 p31 +p03 p12.
And since α is also an injection, the lines in P3 are thus in bijective
correspondence with the points of this quadric in P5, called the Plücker
quadric or Klein quadric.
Uses
Plücker coordinates allow concise solutions to problems of line geometry
in 3-dimensional space, especially those involving incidence.
Line-line Join
In the event that two lines are coplanar but not parallel, their common
plane has equation
0 = (m•d2 )x0 + (d×d2 )•x ,
where x = (x1,x2,x3).
The slightest perturbation will destroy the existence of a common plane,
and near-parallelism of the lines will cause numeric difficulties in finding
such a plane even if it does exist.
Line-line Meet
Dually, two coplanar lines, neither of which contains the origin, have
common point
(x 0 : x) = (d•m2 :m×m2 ).
To handle lines not meeting this restriction.
Plane-line Meet
Given a plane with equation
0 = a0x0 + a1x1 + a2x2+ a3x3,
or more concisely 0 = a0x0+a•x; and given a line not in it with Plücker
coordinates (d:m), then their point of intersection is
(x0 : x) = (a•d : a×m - a0d) .
The point coordinates, (x0:x1:x2:x3), can also be expressed in terms of
Plücker coordinates as
Point-line Join
Dually, given a point (y0:y) and a line not containing it, their common
plane has equation
0 = (y•m) x0 + (y×≤y0m)•x.
The plane coordinates, (a0:a1:a2:a3), can also be expressed in terms of
dual Plücker coordinates as
Line Families
Because the Klein quadric is in P5, it contains linear subspaces of
dimensions one and two (but no higher). These correspond to one- and two-
parametre families of lines in P3.
For example, suppose L and L2 are distinct lines in P3 determined by
points x, y and x2 , y2 , respectively. Linear combinations of their
determining points give linear combinations of their Plücker coordinates,
generating a one-parametre family of lines containing L and L2 . This
corresponds to a one-dimensional linear subspace belonging to the Klein
quadric.
Lines in Plane
If three distinct and non-parallel lines are coplanar; their linear
combinations generate a two-parametre family of lines, all the lines in the
plane. This corresponds to a two-dimensional linear subspace belonging to
the Klein quadric.
Lines Through Point
If three distinct and non-coplanar lines intersect in a point, their linear
combinations generate a two-parametre family of lines, all the lines through
the point. This also corresponds to a two-dimensional linear subspace
belonging to the Klein quadric.
Ruled Surface
A ruled surface is a family of lines that is not necessarily linear. It
corresponds to a curve on the Klein quadric. For example, a hyperboloid of
one sheet is a quadric surface in P3 ruled by two different families of lines,
one line of each passing through each point of the surface; each family
corresponds under the Plücker map to a conic section within the Klein
quadric in P5.
Line Geometry
During the nineteenth century, line geometry was studied intensively. In
terms of the bijection given above, this is a description of the intrinsic
geometry of the Klein quadric.
Ray Tracing
Line geometry is extensively used in ray tracing application where the
geometry and intersections of rays need to be calculated in 3D. An
implementation is described in Introduction to Pluecker Coordinates written
for the Ray Tracing forum by Thouis Jones.
Application to Bézout’s Theorem
Bézout’s theorem predicts that the number of points of intersection of
two curves is equal to the product of their degrees (assuming an algebraically
complete field and with certain conventions followed for counting
intersection multiplicities). Bézout’s theorem predicts there is one point of
intersection of two lines and in general this is true, but when the lines are
parallel the point of intersection is infinite.
Homogeneous coordinates can be used to locate the point of intersection
in this case. Similarly, Bézout’s theorem predicts that a line will intersect a
conic at two points, but in some cases one or both of the points is infinite and
homogeneous coordinates must be used to locate them.
For example, y = x2 and x = 0 have only one point of intersection in the
finite plane. To find the other point of intersection, convert the equations into
homogeneous form, yz = x2 and x = 0. This produces x = yz = 0 and, assuming
not all of x, y and z are 0, the solutions are x = y = 0, z ≠ 0 and x = z = 0, y ≠ 0.
This first solution is the point (0, 0) in Cartesian coordinates, the finite point
of intersection.
The second solutions gives the homogeneous coordinates (0, 1, 0) which
corresponds to the direction of the y-axis. For the equations xy = 1 and x = 0
there are no finite points of intersection. Converting the equations into
homogeneous form gives xy = z2 and x = 0. Solving produces the equation z2
= 0 which has a double root at z = 0.
From the original equation, x = 0, so y ≠ 0 since at least one coordinate
must be non-zero. Therefore (0, 1, 0) is the point of intersection counted with
multiplicity 2 in agreement with the theorem.Circular points at infinity
In projective geometry the circular points at infinity in the complex
projective plane (also called cyclic points or isotropic points) are
(1: i: 0) and (1: - i: 0).
Here the coordinates are homogeneous coordinates (x: y: z); so that the
line at infinity is defined by z = 0. These points at infinity are called circular
points at infinity because they lie on the complexification of every real circle.
In other words, both points satisfy the homogeneous equations of the type
Ax2 + Ay2 + 2B1xz + 2B2yz - Cz2 = 0.
The case where the coefficients are all real gives the equation of a
general circle (of the real projective plane). In general, an algebraic curve that
passes through these two points is called circular.
The circular points at infinity are the points at infinity of the isotropic
lines.
The circular points are invariant under translation and rotation.
Change of Coordinate Systems
Just as the selection of axes in the Cartesian coordinate is somewhat
arbitrary, the selection of a single system of homogeneous coordinates out of
all possible systems is somewhat arbitrary.
Therefore it is useful to know how the different systems are related to
each other.
Let (x, y, z) be the homogeneous coordinates of a point in the projective
plane and for a fixed matrix
Manifold
In mathematics (specifically in geometry and topology), a manifold is a
geometrical object that, near each point, resembles Euclidean space of a fixed
dimension, called the dimension of the manifold. In mathematical terms, a
manifold of dimension n is a topological space such that each point has a
neighbourhood that is homeomorphic to the Euclidean space of dimension n.
Lines and circles (but not figure eights) are one-dimensional manifolds (1-
manifolds). The plane, the sphere and the torus are examples of 2-manifolds.
Although near each point, a manifold resembles Euclidean space,
globally a manifold might not. For example, the surface of the sphere is not a
Euclidean space, but in a region it can be charted by means of geographic
maps: map projections of the region into the Euclidean plane. When a region
appears in two neighbouring maps (in the context of manifolds they are
called charts), the two representations do not coincide exactly and a
transformation is needed to pass from one to the other, called a transition
map.
The concept of a manifold is central to many parts of geometry and
modern mathematical physics because it allows more complicated structures
to be described and understood in terms of the relatively well-understood
properties of Euclidean space.
Manifolds naturally arise as solution sets of systems of equations and as
graphs of functions. Manifolds may have additional features. One important
class of manifolds is the class of differentiable manifolds.
This differentiable structure allows calculus to be done on manifolds. A
Riemannian metric on a manifold allows distances and angles to be
measured. Symplectic manifolds serve as the phase spaces in the Hamiltonian
formalism of classical mechanics, while four-dimensional Lorentzian
manifolds model space-time in general relativity.
Motivational Examples
Circle
After a line, the circle is the simplest example of a topological manifold.
Topology ignores bending, so a small piece of a circle is treated exactly the
same as a small piece of a line. Consider, for instance, the top half of the unit
circle, x2 + y2 = 1, where the y-coordinate is positive (indicated by the yellow
arc in F i g ure 1 ). Any point of this semicircle can be uniquely described by
its x-coordinate. So, projection onto the first coordinate is a continuous, and
invertible, mapping from the upper semicircle to the open interval (– 1,1):
Figure: The four charts each map part of the circle to an open interval, and together cover the whole
circle.
Such functions along with the open regions they map are called charts.
Similarly, there are charts for the bottom (red), left (blue), and right (green)
parts of the circle.
Together, these parts cover the whole circle and the four charts form an
atlas for the circle.
The top and right charts overlap: their intersection lies in the quarter of
the circle where both the x- and the y-coordinates are positive.
The two charts χtop an d χright each map this part into the interval (0, 1).
Thus a function T from (0, 1) to itself can be constructed, which first
uses the inverse of the top chart to reach the circle and then follows the right
chart back to the interval. Let a be any number in (0, 1), then:
Figure: A circle manifold chart based on slope, covering all but one point of the circle.
The top, bottom, left, and right charts show that the circle is a manifold,
but they do not form the only possible atlas. Charts need not be geometric
projections, and the number of charts is a matter of some choice. Consider the
charts
Here s is the slope of the line through the point at coordinates (x,y) and
the fixed pivot point (-1, 0); t is the mirror image, with pivot point (+1, 0).
The inverse mapping from s to (x, y) is given by
Generally manifolds are taken to have a fixed dimension (the space must
be locally homeomorphic to a fixed n-ball), and such a space is called an n-
manifold; however, some authors admit manifolds where different points can
have different dimensions. If a manifold has a fixed dimension, it is called a
pure manifold. For example, the sphere has a constant dimension of 2 and is
therefore a pure manifold whereas the disjoint union of a sphere and a line in
three-dimensional space is not a pure manifold. Since dimension is a local
invariant (i.e. the map sending each point to the dimension of its
neighbourhood over which a chart is defined, is locally constant), each
connected component has a fixed dimension.
Scheme-theoretically, a manifold is a locally ringed space, whose
structure sheaf is locally isomorphic to the sheaf of continuous (or
differentiable, or complex-analytic, etc.) functions on Euclidean space. This
definition is mostly used when discussing analytic manifolds in algebraic
geometry.
Broad Definition
The broadest common definition of manifold is a topological space
locally homeomorphic to a topological vector space over the reals. This omits
the point-set axioms, allowing higher cardinalities and non-Hausdorff
manifolds; and it omits finite dimension, allowing structures such as Hilbert
manifolds to be modeled on Hilbert spaces, Banach manifolds to be modeled
on Banach spaces, and Fréchet manifolds to be modeled on Fréchet spaces.
Usually one relaxes one or the other condition: manifolds with the point-
set axioms are studied in general topology, while infinite-dimensional
manifolds are studied in functional analysis.
Figure: The chart maps the part of the sphere with positive z coordinate to a disc.
Perhaps the simplest way to construct a manifold is the one used in the
example above of the circle. First, a subset of R2 is identified, and then an
atlas covering this subset is constructed. The concept of manifold grew
historically from constructions like this. Here is another example, applying
this method to the construction of a sphere:
Sphere with Charts
A sphere can be treated in almost the same way as the circle. In
mathematics a sphere is just the surface (not the solid interior), which can be
defined as a subset of R3:
The sphere is two-dimensional, so each chart will map part of the sphere
to an open subset of R2. Consider the northern hemisphere, which is the part
with positive z coordinate (coloured red in the picture on the right). The
function χ defined by
χ (x y z) = (x y)
maps the northern hemisphere to the open unit disc by projecting it on the (x,
y) plane. A similar chart exists for the southern hemisphere. Together with
two charts projecting on the (x, z) plane and two charts projecting on the (y, z)
plane an atlas of six charts is obtained which covers the entire sphere. This
can be easily generalized to higher-dimensional spheres.
Patchwork
A manifold can be constructed by gluing together pieces in a consistent
manner, making them into overlapping charts. This construction is possible
for any manifold and hence it is often used as a characterisation, especially
for differentiable and Riemannian manifolds. It focuses on an atlas, as the
patches naturally provide charts, and since there is no exterior space involved
it leads to an intrinsic view of the manifold.
The manifold is constructed by specifying an atlas, which is itself
defined by transition maps. A point of the manifold is therefore an
equivalence class of points which are mapped to each other by transition
maps. Charts map equivalence classes to points of a single patch. There are
usually strong demands on the consistency of the transition maps. For
topological manifolds they are required to be homeomorphisms; if they are
also diffeomorphisms, the resulting manifold is a differentiable manifold.
This can be illustrated with the transition map t = 1D from the second
half of the circle example. Start with two copies of the line. Use the
coordinate s for the first copy, and t for the second copy. Now, glue both
copies together by identifying the point t on the second copy with the point s
= 1D on the first copy (the points t = 0 and s = 0 are not identified with any
point on the first and second copy, respectively). This gives a circle.
Intrinsic and Extrinsic View
The first construction and this construction are very similar, but they
represent rather different points of view. In the first construction, the
manifold is seen as embedded in some Euclidean space. This is the extrinsic
view. When a manifold is viewed in this way, it is easy to use intuition from
Euclidean spaces to define additional structure. For example, in a Euclidean
space it is always clear whether a vector at some point is tangential or normal
to some surface through that point
The patchwork construction does not use any embedding, but simply
views the manifold as a topological space by itself This abstract point of view
is called the intrinsic view. It can make it harder to imagine what a tangent
vector might be, and there is no intrinsic notion of a normal bundle, but
instead there is an intrinsic stable normal bundle.
n-Sphere as a Patchwork
The n-sphere Sn is a generalisation of the idea of a circle (1-sphere) and
sphere (2-sphere) to higher dimensions. An n-sphere Sn can be constructed
by gluing together two copies of Rn . The transition map between them is
defined as
This function is its own inverse and thus can be used in both directions.
As the transition map is a smooth function, this atlas defines a smooth
manifold. In the case n = 1, the example simplifies to the circle example
given earlier.
Identifying Points of a Manifold
It is possible to define different points of a manifold to be same. This can
be visualized as gluing these points together in a single point, forming a
quotient space.
There is, however, no reason to expect such quotient spaces to be
manifolds. Among the possible quotient spaces that are not necessarily
manifolds, orbifolds and CW complexes are considered to be relatively well-
behaved. An example of a quotient space of a manifold that is also a manifold
is the real projective space identified as a quotient space of the corresponding
sphere.
One method of identifying points (gluing them together) is through a
right (or left) action of a group, which acts on the manifold. Two points are
identified if one is moved onto the other by some group element. If M is the
manifold and G is the group, the resulting quotient space is denoted by M / G
(or G \ M).
Manifolds which can be constructed by identifying points include tori
and real projective spaces (starting with a plane and a sphere, respectively).
Gluing Along Boundaries
Two manifolds with boundaries can be glued together along a boundary.
If this is done the right way, the result is also a manifold. Similarly, two
boundaries of a single manifold can be glued together.
Formally, the gluing is defined by a bijection between the two
boundaries. Two points are identified when they are mapped onto each other.
For a topological manifold this bijection should be a homeomorphism,
otherwise the result will not be a topological manifold. Similarly for a
differentiable manifold it has to be a diffeomorphism. For other manifolds
other structures should be preserved.
A finite cylinder may be constructed as a manifold by starting with a
strip [0, 1] × [0, 1] and gluing a pair of opposite edges on the boundary by a
suitable diffeomorphism. A projective plane may be obtained by gluing a
sphere with a hole in it to a Möbius strip along their respective circular
boundaries.
6-sphere Coordinates
In mathematics, 6-sphere coordinates are the coordinate system created
by inverting the Cartesian coordinates across the unit sphere. The three
coordinates are
Since inversion is its own inverse, the equations for x, y, and z in terms
of u, v, and w are similar:
Abscissa
In mathematics, abscissa (plural abscissae or abscissæ) refers to that
element of an ordered pair which is plotted on the horizontal axis of a two-
dimensional Cartesian coordinate system, as opposed to the ordinate. It is the
first of the two terms (often labelled x and y) which define the location of a
point in such a coordinate system.
The usage of the word abscissa is first recorded in 1659 by Stefano degli
Angeli, a mathematics professor in Rome, according to Moritz Cantor. Soon
thereafter, Leibniz used the term extensively in Latin in his Mathematische
Schriften (1692), after which it became a standardized mathematical term.
The first occurrence of the term in English is found in An Institution of
Fluxions by the English mathematician Humphry Ditton (1706), where he
spells the word abscisse, possibly denoting the plural.
Examples
For the point (-7, 3), -7 is called the abscissa and 3 the ordinate.
5: Affine Coordinate System
Note that Rj - O are difference vectors with the origin in O and ends in R.
An affine space cannot have a coordinate system with n less than its
dimension, but n may indeed be greater, which means that the coordinate
map is not necessary surjective. Examples of n-coordinate system in an (n-1)-
dimensional space are barycentric coordinates and affine homogeneous
coordinates (1, x1, … x 1). In the latter case the x0 coordinate is equal to 1 on
all space, n - but this “reserved” coordinate facilitates matrix representation of
affine maps.
Alpha-numeric Grid
a b c d e f
1 a1 b1 c1 d1 e1 f1
2 a2 b2 c2 d2 e2 f2
3 a3 b3 c3 d3 e3 f3
4 a4 b4 c4 d4 e4 f4
5 a5 b5 c5 d5 e5 f5
6 a6 b6 c6 d6 e6 f6
Bipolar Coordinates
Bipolar coordinates are a two-dimensional orthogonal coordinate system.
There are two commonly defined types of bipolar coordinates. The other
system is two-centre bipolar coordinates. There is also a third coordinate
system that is based on two poles (biangular coordinates).
The first is based on the Apollonian circles. The curves of constant σ and
of τ are circles that intersect at right angles. The coordinates have two foci F1
and F2, which are generally taken to be fixed at (“a, 0) and (a, 0),
respectively, on the x-axis of a Cartesian coordinate system.
Bipolar coordinates form the basis for several sets of three-dimensional
orthogonal coordinates. The bipolar cylindrical coordinates are produced by
projecting in the z-direction. The bispherical coordinates are produced by
rotating the bipolar coordinates about the -axis, i.e., the axis connecting the
foci, whereas the toroidal coordinates are produced by rotating the bipolar
coordinates about the y-axis, i.e., the axis separating the foci.
(Recall that F1 and F2 are located at (-a, 0) and (a, 0), respectively.)
Equivalently
where the σcoordinate of a point P equals the angle F1PF2 and the
τcoordinate equals the natural logarithm of the ratio of the distances d1 and d2
to the focal lines
(Recall that the focal lines F1 and F2 are located at x = -a and x = +a,
respectively.)
Surfaces of constant σ correspond to cylinders of different radii
that all pass through the focal lines and are not concentric. The surfaces of
constant rare non-intersecting cylinders of different radii
that surround the focal lines but again are not concentric. The focal lines and
all these cylinders are parallel to the z-axis (the direction of projection). In the
z = 0 plane, the centres of the constant-σ and constant- cylinders lie on the
and axes, respectively.
Scale Factors
The scale factors for the bipolar coordinates σ and τare equal
Bispherical Coordinates
Illustration of bispherical coordinates, which are obtained by rotating a
two-dimensional bipolar coordinate system about the axis joining its two foci.
The foci are located at distance 1 from the vertical z-axis. The red self-
interecting torus is the o=45° isosurface, the blue sphere is the x=0.5
isosurface, and the yellow half-plane is the φ=60° isosurface. The green half-
plane marks the x-z plane, from which φ is measured. The black point is
located at the intersection of the red, blue and yellow isosurfaces, at Cartesian
coordinates roughly (0.841, -1.456, 1.239).
Bispherical coordinates are a three-dimensional orthogonal coordinate
system that results from rotating the two-dimensional bipolar coordinate
system about the axis that connects the two foci. Thus, the two foci F1 and F
in bipolar coordinates remain points (on the z-axis, the axis of rotation) in the
bispherical coordinate system.
Hyperbolic Coordinates
In mathematics, hyperbolic coordinates are a method of locating points
in Quadrant I of the Cartesian plane
Identity Line
In a 2-dimensional Cartesian coordinate system, with x representing the
abscissa and y the ordinate, the identity line is the y = x line. The line,
sometimes called the 1:1 line, has a slope of 1. When the abscissa and
ordinate are on the same scale, the identity line forms a 45° angle with the
abscissa, and is thus also, informally, called the 45° line.
The line is often used as a reference in a 2-dimensional scatter plot
comparing two sets of data expected to be identical under ideal conditions.
When the corresponding data points from the two data sets are equal to each
other, the corresponding scatters fall exactly on the identity line.
Jacobi Coordinates
In the theory of many-particle systems, Jacobi coordinates often are used
to simplify the mathematical formulation. These coordinates are particularly
common in treating polyatomic molecules and chemical reactions, and in
celestial mechanics. An algorithm for generating the Jacobi coordinates for N
bodies may be based upon binary trees. In words, the algorithm is described
as follows:
Let mj and mk be the masses of two bodies that are replaced by a new
body of virtual mass M = mj + mk. The position coordinates xj and xk are
replaced by their relative position rjk = xj – xk and by the vector to their centre
of mass Rjk = (mj qj + mkqk)/(mj + mk). The node in the binary tree
corresponding to the virtual body has mj as its right child and mk as its left
child. The order of children indicates the relative coordinate points from xk to
xj. Repeat the above step for N - 1 bodies, that is, the N - 2 original bodies
plus the new virtual body.
For the four-body problem the result is:
Line Coordinates
In geometry, line coordinates are used to specify the position of a line
just as point coordinates (or simply coordinates) are used to specify the
position of a point.
Lines in the Plane
There are several possible ways to specify the position of a line in the
plane. A simple way is by the pair (m, b) where the equation of the line is y
=mx + b. Here m is the slope and b is the x-intercept. This system specifies
coordinates for all lines that are not vertical.
However, it is more common and simpler algebraically to use
coordinates (l, m) where the equation of the line is lx + my + 1 = 0. This
system specifies coordinates for all lines except those that pass through the
origin. The geometrical interpretations of l and m are the negative reciprocals
of the x and y-intercept respectively.
The exclusion of lines passing through the origin can be resolved by
using a system of three coordinates (l, m, n) to specify the line in which the
equation, lx + my + n = 0. Here l and m may not both be 0. In this equation,
only the ratios between l, m and n are significant, in other words if the
coordinates are multiplied by a non-zero scalar then line represented remains
the same. So (l, m, n) is a system of homogeneous coordinates for the line.
If points in the plane are represented by homogeneous coordinates (x, y,
z), the equation of the line is lx + my + nz = 0. In this context, l, m and n may
not all be 0. In particular, (0, 0, 1) represents the line z = 0, which is the line
at infinity in the projective plane. The coordinates (0, 1, 0) and (1, 0, 0)
represent the x and y-axes respectively.
Tangential Equations
Just as f(x, y) = 0 can represent a curve as a subset of the points in the
plane, the equation ϕ(l, m) = 0 represents a subset of the lines on the plain.
The set of lines on the plane may, in an abstract sense, be thought of as the
set of points in a projective plane, the dual of the original plane. The equation
ϕ(l, m) = 0 then represents a curve in the dual plane.
For a curve f(x, y) = 0 in the plane, the tangents to the curve form a curve
in the dual space called the dual curve. If ϕ(l, m) = 0 is the equation of the
dual curve, then it is called the tangential equation, for the original curve. A
given equation ϕ(l, m) = 0 represents a curve in the original plane determined
as the envelope of the lines that satisfy this equation.
Similarly, if ϕ(l, m, n) is a homogeneous function then ϕ(l, m, n) = 0
represents a curve in the dual space given in homogeneous coordinates, and
may be called the homogeneous tangential equation of the enveloped curve.
Tangential equations are useful in the study of curves defined as
envelopes, just as Cartesian equations are useful in the study of curves
defined as loci.
Tangential Equation of a Point
A linear equation in line coordinates has the form al + bm + c = 0, where
a, b and c are constants. Suppose (l, m) is a line that satisfies this equation. If
c is not 0 then lx + my + 1 = 0, where x = a/c and y = b/c, so every line
satisfying the original equation passes though the point (x, y).
Conversely, any line through (x, y) satisfies the original equation, so al +
bm + c = 0 is the equation of set of lines through (x, y). For a given point (x,
y), the equation of the set of lines though it is lx + my + 1 = 0, so this may be
defined as the tangential equation of the point. Similarly, for a point (x, y, z)
given in homogeneous coordinates, then the equation of the point in
homogeneous tangential coordinates is (lx, my, nz) = 0.
Formulas
The intersection of the lines (l1, m1) and (l2, m2) is the solution to the
linear equations
The lines (l1, m1), (l2, m2), and (l3, m3) are concurrent when the
determinant
For homogeneous coordinates, the intersection of the lines (l1, m1, n1)
and (l2, m2, n2) is
(m1n2 -m2n1,l2n1 -l1n2,l1m2 -l2m1).
The lines (l1, m1, n1), (l2, m2, n2) and (l3, m3, n3) are concurrent when the
determinant
Dually, the coordinates of the line containing (x1, y1, z1) and (x2, y2, z2)
are
(y1z2 - y2z1,x2z1 - x1z2 ,x1y2 - x2y1).
Log-polar Coordinates
Log-polar coordinates (or logarithmic polar coordinates) is a coordinate
system in two dimensions, where a point is identified by two numbers, one
for the logarithm of the distance to a certain point, and one for an angle.
Log-polar coordinates are closely connected to polar coordinates, which
are usually used to describe domains in the plane with some sort of rotational
symmetry. In areas like harmonic and complex analysis, the log-polar
coordinates are more canonical than polar coordinates.
Definition and Coordinate Transformations
Log-polar coordinates in the plane consist of a pair of real numbers
(ρ,θ), where θ is the logarithm of the distance to a given point (the origin) and
is the angle between a line of reference (the x-axis) and the line through the
origin and the point. The angular coordinate is the same as for polar
coordinates, while the radial coordinate is transformed according to the rule
r = eρ.
where r is the distance to the origin. The formulas for transformation from
Cartesian coordinates to log-polar coordinates are given by
has the same simple expression as in Cartesian coordinates. This is true for
all coordinate systems where the transformation to Cartesian coordinates is
given by a conformal mapping.
Thus, when considering Laplace’s equation for a part of the plane with
rotational symmetry e.g. a circular disk, log-polar coordinates is the natural
choice.
Dirichlet-to-Neumann Function
The latter coordinate system is for instance suitable for dealing with
Dirichlet-to-Neumann functions. If the discrete coordinate system is
interpreted as an undirected graph in the unit disc, it can be considered as a
model for an electrical network.
To every line segment in the graph is associated a conductance given by
a function/.
The electrical network will then serve as a discrete model for the
Dirichlet problem in the unit disc, where the Laplace equation takes the form
of Kirchhoff’s law
On the nodes on the boundary of the circle, an electrical potential
(Dirichlet data) is defined, which induces an electrical current (Neumann
data) through the boundary nodes.
The linear function A, mapping Dirichlet data to Neumann data is called
a Dirichlet-to-Neumann function, and depends on the topology and
conductance of the network.
In the case with the continuous disc, it follows that if the conductance is
homogeneous, let’s say y = 1 everywhere, then the Dirichlet-to-Neumann
function satisfies the following equation
In order to get a good discrete model of the Dirichlet problem, it would
be useful to find a graph in the unit disc whose (discrete) Dirichlet-to-
Neumann function has the same property.
Even though polar coordinates don’t give us any answer, this is exactly
what the spiral network given by log-polar coordinates provides us with.
Image Analysis
Already at the end of the 1970s, applications for the discrete spiral
coordinate system were given in image analysis. To represent an image in
this coordinate system rather than in Cartesian coordinates, gives
computational advantages when rotating or zooming in an image.
Also, the photo receptors in the retina in the human eye are distributed in
a way that has big similarities with the spiral coordinate system. It can also be
found in the Mandelbrot fractal.
Log-polar coordinates can also be used to construct fast methods for the
Radon transform and its inverse.
6: Orthogonal Coordinates
where d is the dimension and the scaling functions (or scale factors)
equal the square roots of the diagonal components of the metric tensor, or the
lengths of the local basis vectors ek described below. These scaling functions
hi ar e used to calculate differential operators in the new coordinates, e.g., the
gradient, the Laplacian, the divergence and the curl.
A simple method for generating orthogonal coordinates systems in two
dimensions is by a conformal mapping of a standard two-dimensional grid of
Cartesian coordinates (x, y). A complex number z = x + iy can be formed
from the real coordinates x and y, where i represents the square root of -1.
Any holomorphic function w = f(z) with non-zero complex derivative will
produce a conformal mapping; if the resulting complex number is written w =
u + iv, then the curves of constant u and v intersect at right angles, just as the
original lines of constant x and y did. Orthogonal coordinates in three and
higher dimensions can be generated from an orthogonal two-dimensional
coordinate system, either by projecting it into a new dimension (cylindrical
coordinates) or by rotating the two-dimensional system about one of its
symmetry axes. However, there are other orthogonal coordinate systems in
three dimensions that cannot be obtained by projecting or rotating a two-
dimensional system, such as the ellipsoidal coordinates. More general
orthogonal coordinates may be obtained by starting with some necessary
coordinate surfaces and considering their orthogonal trajectories.
Basis Vectors
Covariant Vasis
In Cartesian coordinates, the basis vectors are fixed (constant). In the
more general setting of curvilinear coordinates, a point in space is specified
by the coordinates, and at every such point there is bound a set of basis
vectors, which generally are not constant: this is the essence of curvilinear
coordinates in general and is a very important concept. What distinguishes
orthogonal coordinates is that, though the basis vectors vary, they are always
orthogonal with respect to each other. In other words,
These basis vectors are by definition the tangent vectors of the curves
obtained by varying one coordinate, keeping the others fixed:
Figure: Visualization of 2D orthogonal coordinates. Curves obtained by holding all but one coordinate
constant are shown, along with basis vectors. Note that the basis vectors aren’t of equal length: they
need not be, they only need to be orthogonal.
where r is some point and qi is the coordinate for which the basis vector is
extracted. In other words, a curve is obtained by fixing all but one coordinate;
the unfixed coordinate is varied as in a parametric curve, and the derivative of
the curve with respect to the parametre (the varying coordinate) is the basis
vector for that coordinate.
Note that the vectors are not necessarily of equal length. The normalized
basis vectors are notated with a hat and obtained by dividing by the length:
this follows from the fact that, by definition, ei ·ej =δij , using the Kronecker
delta. Note that:
The position of the indices represent how the components are calculated
(upper indices should not be confused with exponentiation). Note that the
summation symbols Σ (capital Sigma) and the summation range, indicating
summation over all basis vectors (i = 1, 2, ..., d), are often omitted. The
components are related simply by:
hi2xi =xi
There is no distinguishing widespread notation in use for vector
components with respect to the normalized basis; in this article we’ll use
subscripts for vector components and note that the components are calculated
in the normalized basis.
Vector Algebra
Vector addition and negation are done component-wise just as in
Cartesian coordinates with no complication. Extra considerations may be
necessary for other vector operations. Note however, that all of these
operations assume that two vectors in a vector field are bound to the same
point (in other words, the tails of vectors coincide). Since basis vectors
generally vary in orthogonal coordinates, if two vectors are added whose
components are calculated at different points in space, the different basis
vectors require consideration.
Dot Product
The dot product in Cartesian coordinates (Euclidean space with an
orthonormal basis set) is simply the sum of the products of components. In
orthogonal coordinates, the dot product of two vectors x and y takes this
familiar form when the components of the vectors are calculated in the
normalized basis:
Vector Calculus
Differentiation
Looking at an infinitesimal displacement from some point, it’s apparent
that
that open in the opposite direction, i.e., towards -y . The foci of all these
parabolic cylinders are located along the line defined by x = y = 0 . The radius
r has a simple formula as well
that proves useful in solving the Hamilton-Jacobi equation in parabolic
coordinates for the inverse-square central force problem of mechanics.
Scale Factors
The scale factors for the parabolic cylindrical coordinates σand τare:
The parabolic cylinder harmonics for (m,n) are now the product of the
solutions. The combination will reduce the number of constants and the
general solution to Laplace’s equation may be written:
Applications
The classic applications of parabolic cylindrical coordinates are in
solving partial differential equations, e.g., Laplace’s equation or the
Helmholtz equation, for which such coordinates allow a separation of
variables.
A typical example would be the electric field surrounding a flat semi-
infinite conducting plate.
where the centres of this coordinate system are at (+a, 0) and (a, 0).
Polar Coordinates
To polar coordinates from two-centre bipolar coordinates
Darboux’s Theorem
Darboux’s theorem is a theorem in the mathematical field of differential
geometry and more specifically differential forms, partially generalizing the
Frobenius integration theorem. It is a foundational result in several fields, the
chief among them being symplectic geometry The theorem is named after
Jean Gaston Darboux who established it as the solution of the Pfaff problem.
One of the many consequences of the theorem is that any two symplectic
manifolds of the same dimension are locally symplectomorphic to one
another. That is, every 2n-dimensional symplectic manifold can be made to
look locally like the linear symplectic space Cn with its canonical symplectic
form. There is also an analogous consequence of the theorem as applied to
contact geometry.
Statement and First Consequences
The precise statement is as follows. Suppose that θ is a differential 1-
form on an n dimensional manifold, such that dθ has constant rank p. If
then there is a local system of coordinates x1,...,x , y1, ..., yin which n - p p
Isothermal Coordinates
In mathematics, specifically in differential geometry, isothermal
coordinates on a Riemannian manifold are local coordinates where the metric
is conformal to the Euclidean metric. This means that in isothermal
coordinates, the Riemannian metric locally has the form
where λ and μ are smooth with λ > 0 and |μ| < 1. In fact
has a diffeomorphic solution. Such a solution has been proved to exist in any
neighbourhood where | |μ|| _ < 1.
Gaussian Curvature
In the isothermal coordinates (u, v), the Gaussian curvature takes the
simpler form
where ρ = e*.
Normal Coordinates
In differential geometry, normal coordinates at a point p in a
differentiable manifold equipped with a symmetric affine connection are a
local coordinate system in a neighbourhood of p obtained by applying the
exponential map to the tangent space at p. In a normal coordinate system, the
Christoffel symbols of the connection vanish at the point p, thus often
simplifying local calculations. In normal coordinates associated to the Levi
Civita connection of a Riemannian manifold, one can additionally arrange
that the metric tensor is the Kronecker delta at the point p, and that the first
partial derivatives of the metric at p vanish.
A basic result of differential geometry states that normal coordinates at a
point always exist on a manifold with a symmetric affine connection. In such
coordinates the covariant derivative reduces to a partial derivative (at p only),
and the geodesics through p are locally linear functions of t (the affine
parametre). This idea was implemented in a fundamental way by Albert
Einstein in the general theory of relativity: the equivalence principle uses
normal coordinates via inertial frames. Normal coordinates always exist for
the Levi-Civita connection of a Riemannian or Pseudo-Riemannian manifold.
By contrast, there is no way to define normal coordinates for Finsler
manifolds (Busemann 1955).
Geodesic Normal Coordinates
Geodesic normal coordinates are local coordinates on a manifold with an
affine connection afforded by the exponential map
given by any basis of the tangent space at the fixed basepoint p -M. If the
additional structure of a Riemannian metric is imposed, then the basis defined
by E may be required in addition to be orthonormal, and the resulting
coordinate system is then known as a Riemannian normal coordinate system.
Normal coordinates exist on a normal neighbourhood of a point p in M. A
normal neighbourhood U is a subset of M such that there is a proper
neighbourhood V of the origin in the tangent space T M and exp acts as a
diffeomorphism between U and V. Now let U be a normal neighbourhood of
p in M then the chart is given by:
Quadratic Formula
A quadratic equation with real or complex coefficients has two solutions,
called roots. These two solutions may or may not be distinct, and they may or
may not be real.
Having
ax2+bx + c = 0,
the roots are given by the quadratic formula
which produces
In the case that b ‘Š 0, there are two distinct roots, but if the polynomial
is irreducible, they cannot be expressed in terms of square roots of numbers
in the coefficient field. Instead, define the 2-root R(c) of c to be a root of the
polynomial x2 + x + c, an element of the splitting field of that polynomial.
One verifies that R(c) + 1 is also a root. In terms of the 2-root operation, the
two roots of the (non-monic) quadratic ax2 + bx + c are
and
Trigonometric Functions
In mathematics, the trigonometric functions (also called circular
functions) are functions of an angle. They are used to relate the angles of a
triangle to the lengths of the sides of a triangle. Trigonometric functions are
important in the study of triangles and modelling periodic phenomena, among
many other applications.
The most familiar trigonometric functions are the sine, cosine, and
tangent. In the context of the standard unit circle with radius 1, where a
triangle is formed by a ray originating at the origin and making some angle
with the x-axis, the sine of the angle gives the length of the y-component
(rise) of the triangle, the cosine gives the length of the x-component (run),
and the tangent function gives the slope (y-component divided by the x-
component). More precise definitions are detailed below. Trigonometric
functions are commonly defined as ratios of two sides of a right triangle
containing the angle, and can equivalently be defined as the lengths of
various line segments from a unit circle. More modern definitions express
them as infinite series or as solutions of certain differential equations,
allowing their extension to arbitrary positive and negative values and even to
complex numbers.
Trigonometric functions have a wide range of uses including computing
unknown lengths and angles in triangles (often right triangles). In this use,
trigonometric functions are used, for instance, in navigation, engineering, and
physics. A common use in elementary physics is resolving a vector into
Cartesian coordinates. The sine and cosine functions are also commonly used
to model periodic function phenomena such as sound and light waves, the
position and velocity of harmonic oscillators, sunlight intensity and day
length, and average temperature variations through the year. In modern
usage, there are six basic trigonometric functions, tabulated here with
equations that relate them to one another. Especially with the last four, these
relations are often taken as the definitions of those functions, but one can
define them equally well geometrically, or by other means, and then derive
these relations.
Right-angled Triangle Definitions
Figure: (Top): Trigonometric function sinè for selected angles θ, π -θ, π + θ and 2π-θ in the four
quadrants. (Bottom) Graph of sine function versus angle. Angles from the top panel are identified.
The notion that there should be some standard correspondence between
the lengths of the sides of a triangle and the angles of the triangle comes as
soon as one recognizes that similar triangles maintain the same ratios
between their sides.
That is, for any similar triangle the ratio of the hypotenuse (for example)
and another of the sides remains the same. If the hypotenuse is twice as long,
so are the sides. It is these ratios that the trigonometric functions express.
To define the trigonometric functions for the angle A, start with any right
triangle that contains the angle A. The three sides of the triangle are named as
follows:
• The hypotenuse is the side opposite the right angle, in this case side
h. The hypotenuse is always the longest side of a right-angled
triangle.
• The opposite side is the side opposite to the angle we are interested in
(angle A), in this case side a.
• The adjacent side is the side having both the angles of interest (angle
A and right-angle C), in this case side b.
In ordinary Euclidean geometry, according to the triangle postulate, the
inside angles of every triangle total 180° (n radians). Therefore, in a right-
angled triangle, the two non-right angles total 90° (n/2 radians), so each of
these angles must be in the range of (0°,90°) as expressed in interval notation.
The following definitions apply to angles in this 0° - 90° range. They can
be extended to the full set of real arguments by using the unit circle, or by
requiring certain symmetries and that they be periodic functions.
For example, the figure shows sin θ for angles θ, n- θ, n+ θ, and 2π
depicted on the unit circle (top) and as a graph (bottom).
The value of the sine repeats itself apart from sign in all four quadrants,
and if the range of 0is extended to additional rotations, this behaviour repeats
periodically with a period 2n.
The trigonometric functions are summarized in the following table and
described in more detail below.
The angle θ is the angle between the hypotenuse and the adjacent line -
the angle at A in the accompanying diagram.
Note that this ratio does not depend on size of the particular right triangle
chosen, as long as it contains the angle A, since all such triangles are similar.
The cosine of an angle is the ratio of the length of the adjacent side to the
length of the hypotenuse: so called because it is the sine of the
complementary or co-angle. In our case
The tangent of an angle is the ratio of the length of the opposite side to
the length of the adjacent side: so called because it can be represented as a
line segment tangent to the circle, that is the line that touches the circle, from
Latin linea tangens or touching line (cf. tangere, to touch). In our case
The acronyms “SOHCAHTOA” (“Soak-a-toe”, “Sock-a-toa”, “Soak-a-
toa”) and “OHSAHCOAT” are commonly used mnemonics for these ratios.
Reciprocal Functions
The remaining three functions are best defined using the above three
functions.
The cosecant csc(A), or cosec(A), is the reciprocal of sin(A), i.e. the ratio
of the length of the hypotenuse to the length of the opposite side:
The secant sec(A) is the reciprocal of cos(A), i.e. the ratio of the length of
the hypotenuse to the length of the adjacent side:
It is so called because it represents the line that cuts the circle (from
Latin: secare, to cut).
The cotangent cot(A) is the reciprocal of tan(A), i.e. the ratio of the
length of the adjacent side to the length of the opposite side:
Slope Definitions
Equivalent to the right-triangle definitions, the trigonometric functions
can also be defined in terms of the rise, run, and slope of a line segment
relative to horizontal. The slope is commonly taught as “rise over run” or D
run. The three main trigonometric functions are commonly taught in the order
sine, cosine, tangent. With a line segment length of 1 (as in a unit circle), the
following mnemonic devices show the correspondence of definitions:
1. “Sine is first, rise is first” meaning that Sine takes the angle of the
line segment and tells its vertical rise when the length of the line is 1.
2. “Cosine is second, run is second” meaning that Cosine takes the
angle of the line segment and tells its horizontal run when the length
of the line is 1.
3. “Tangent combines the rise and run” meaning that Tangent takes the
angle of the line segment and tells its slope; or alternatively, tells the
vertical rise when the line segment’s horizontal run is 1.
This shows the main use of tangent and arctangent: converting between
the two ways of telling the slant of a line, i.e., angles and slopes. (Note that
the arctangent or “inverse tangent” is not to be confused with the cotangent,
which is cosine divided by sine.) While the length of the line segment makes
no difference for the slope (the slope does not depend on the length of the
slanted line), it does affect rise and run. To adjust and find the actual rise and
run when the line does not have a length of 1, just multiply the sine and
cosine by the line length. For instance, if the line segment has length 5, the
run at an angle of 7° is 5 cos(7°)
Series Definitions
Trigonometric functions are analytic functions. Using only geometry and
properties of limits, it can be shown that the derivative of sine is cosine and
the derivative of cosine is the negative of sine. (Here, and generally in
calculus, all angles are measured in radians.) One can then use the theory of
Taylor series to show that the following identities hold for all real numbers x:
These identities are sometimes taken as the definitions of the sine and
cosine function. They are often used as the starting point in a rigorous
treatment of trigonometric functions and their applications (e.g., in Fourier
series), since the theory of infinite series can be developed, independent of
any geometric considerations, from the foundations of the real number
system.
The differentiability and continuity of these functions are then
established from the series definitions alone.
Combining these two series gives Euler’s formula: cos x + i sin x = eix.
Other series can be found. For the following trigonometric functions:
Un is the nth up/down number,
Bn is the nth Bernoulli number, and
En (below) is the nth Euler number.
Tangent
When this series for the tangent function is expressed in a form in which
the denominators are the corresponding factorials, the numerators, called the
“tangent numbers”, have a combinatorial interpretation: they enumerate
alternating permutations of finite sets of odd cardinality.
Cosecant
When this series for the secant function is expressed in a form in which
the denominators are the corresponding factorials, the numerators, called the
“secant numbers”, have a combinatorial interpretation: they enumerate
alternating permutations of finite sets of even cardinality.
Cotangent
This identity can be proven with the Herglotz trick. By combining the –
n-th with the n-th term, it can be expressed as an absolutely convergent
series:
Computation
The computation of trigonometric functions is a complicated subject,
which can today be avoided by most people because of the widespread
availability of computers and scientific calculators that provide built-in
trigonometric functions for any angle. This section, however, describes
details of their computation in three important contexts: the historical use of
trigonometric tables, the modern techniques used by computers, and a few
“important” angles where simple exact values are easily found.
The first step in computing any trigonometric function is range reduction
—reducing the given angle to a “reduced angle” inside a small range of
angles, say 0 to π/2, using the periodicity and symmetries of the
trigonometric functions. Prior to computers, people typically evaluated
trigonometric functions by interpolating from a detailed table of their values,
calculated to many significant figures. Such tables have been available for as
long as trigonometric functions have been described, and were typically
generated by repeated application of the half-angle and angle-addition
identities starting from a known value (such as sin(π/2) = 1).
Modern computers use a variety of techniques. One common method,
especially on higher-end processors with floating point units, is to combine a
polynomial or rational approximation (such as Chebyshev approximation,
best uniform approximation, and Padé approximation, and typically for
higher or variable precisions, Taylor and Laurent series) with range reduction
and a table lookup—they first look up the closest angle in a small table, and
then use the polynomial to compute the correction. Devices that lack
hardware multipliers often use an algorithm called CORDIC (as well as
related techniques), which uses only addition, subtraction, bitshift, and table
lookup. These methods are commonly implemented in hardware floating-
point units for performance reasons.
For very high precision calculations, when series expansion convergence
becomes too slow, trigonometric functions can be approximated by the
arithmetic-geometric mean, which itself approximates the trigonometric
function by the (complex) elliptic integral.
Finally, for some simple angles, the values can be easily computed by
hand using the Pythagorean theorem, as in the following examples. For
example, the sine, cosine and tangent of any integer multiple of π/ 60 radians
(3°) can be found exactly by hand. Consider a right triangle where the two
other angles are equal, and therefore are both π/4 radians (45°). Then the
length of side b and the length of side a are equal; we can choose a=b=1 .
The values of sine, cosine and tangent of an angle of π/4 radians (45°) can
then be found using the Pythagorean theorem:
In this formula the angle at C is opposite to the side c. This theorem can
be proven by dividing the triangle into two right ones and using the
Pythagorean theorem.
The law of cosines can be used to determine a side of a triangle if two
sides and the angle between them are known. It can also be used to find the
cosines of an angle (and consequently the angles themselves) if the lengths of
all the sides are known.
Law of Tangents
The following all form the law of tangents
The explanation of the formulae in words would be cumbersome, but the
patterns of sums and differences; for the lengths and corresponding opposite
angles, are apparent in the theorem.
Law of Cotangents
If
(the semi-perimetre for the triangle), then the following all form the law of
cotangents
In words the theorem is: the cotangent of a half-angle equals the ratio of
the semi-perimetre minus the opposite side to the said angle, to the inradius
for the triangle.
History
While the early study of trigonometry can be traced to antiquity, the
trigonometric functions as they are in use today were developed in the
medieval period. The chord function was discovered by Hipparchus of
Nicaea (180–125 BC) and Ptolemy of Roman Egypt (90–165 AD). The
functions sine and cosine can be traced to the jyâ and koti-jyâ functions used
in Gupta period Indian astronomy (Aryabhatiya, Surya Siddhanta), via
translation from Sanskrit to Arabic and then from Arabic to Latin.
All six trigonometric functions in current use were known in Islamic
mathematics by the 9th century, as was the law of sines, used in solving
triangles. al-Khwârizmî produced tables of sines, cosines and tangents. They
were studied by authors including Omar Khayyαm, Bhâskara II, Nasir al-Din
al-Tusi, Jamshîd al-Kâshî (14th century), Ulugh Beg (14th century),
Regiomontanus (1464), Rheticus, and Rheticus’ student Valentinus Otho
Madhava of Sangamagrama (c. 1400) made early strides in the analysis
of trigonometric functions in terms of infinite series. The first published use
of the abbreviations ‘sin’, ‘cos’, and ‘tan’ is by the 16th century French
mathematician Albert Girard. In a paper published in 1682, Leibniz proved
that sin x is not an algebraic function of x. Leonhard Euler’s Introductio in
analysin infinitorum (1748) was mostly responsible for establishing the
analytic treatment of trigonometric functions in Europe, also defining them as
infinite series and presenting “Euler’s formula”, as well as the near-modern
abbreviations sin., cos., tang., cot., sec., and cosec. A few functions were
common historically, but are now seldom used, such as the chord (crd(^ = 2
sin(02)), the versine (versing = 1 - cos(^ = 2 sin2(02)) (which appeared in the
earliest tables ), the haversine (haversin(^ = versing/ 2 = sin2(02)), the
exsecant (exsec(^ = sec(^ - 1) and the excosecant (excsc(^ = exsec(W2 - θ) =
csc(^ -1). Many more relations between these functions are listed in the
article about trigonometric identities.
Etymologically the word sine derives from the Sanskrit word for half the
chord, jya-ardha, abbreviated to jiva. This was transliterated in Arabic as
jiba, written jb, vowels not being written in Arabic. Next, this transliteration
was mis-translated in the 12th century into Latin as sinus, under the mistaken
impression that jb stood for the word jaib, which means “bosom” or “bay” or
“fold” in Arabic, as does sinus in Latin. Finally, English usage converted the
Latin word sinus to sine. The word tangent comes from Latin tangens
meaning “touching”, since the line touches the circle of unit radius, whereas
secant stems from Latin secans - “cutting” -since the line cuts the circle.
Conic Section
In mathematics, a conic section (or just conic) is a curve obtained as the
intersection of a cone (more precisely, a right circular conical surface) with a
plane. In analytic geometry a conic may be defined as a plane algebraic curve
of degree 2. There are a number of other geometric definitions possible One
of the most useful, in that it involves only the plane, is that a conic consists of
those points whose distances to some point, called a focus, and some line,
called a directrix, are in a fixed ratio, called the eccentricity. Traditionally,
the three types of conic section are the hyperbola, the parabola, and the
ellipse. The circle is a special case of the ellipse, and is of sufficient interest
in its own right that it is sometimes called the fourth type of conic section.
The type of a conic corresponds to its eccentricity, those with eccentricity
less than 1 being ellipses, those with eccentricity equal to 1 being parabolas,
and those with eccentricity greater than 1 being hyperbolas. In the focus-
directrix definition of a conic the circle is a limiting case with eccentricity 0.
In modern geometry certain degenerate cases, such as the union of two lines,
are included as conics as well.
The conic sections were named and studied at least since 200 BC, when
Apollonius of Perga undertook a systematic study of their properties.
History
Menaechmus
It is believed that the first definition of a conic section is due to
Menaechmus (died 320 BC). This work does not survive, however, and is
only known through secondary accounts. The definition used at that time
differs from the one commonly used today in that it requires the plane cutting
the cone to be perpendicular to one of the lines that generate the cone as a
surface of revolution (a generatrix).
Thus the shape of the conic is determined by the angle formed at the
vertex of the cone (between two opposite generatrices): If the angle is acute
then the conic is an ellipse; if the angle is right then the conic is a parabola;
and if the angle is obtuse then the conic is a hyperbola. Note that the circle
cannot be defined this way and was not considered a conic at this time.
Euclid ( fl. 300 BC ) is said to have written four books on conics but
these were lost as well. Archimedes (died c. 212 BC) is known to have
studied conics, having determined the area bounded by a parabola and an
ellipse. The only part of this work to survive is a book on the solids of
revolution of conics.
Apollonius of Perga
The greatest progress in the study of conics by the ancient Greeks is due
to Apollonius of Perga (died c.190 BC) ), whose eight volume Conic Sections
summarized the existing knowledge at the time and greatly extended it.
Apollonius’s major innovation was to characterize a conic using properties
within the plane and intrinsic to the curve; this greatly simplified analysis.
With this tool, it was now possible to show that any plane cutting the
cone, regardless of its angle, will produce a conic according to the earlier
definition, leading to the definition commonly used today.
Pappus of Alexandria (died c. 350 CE) is credited with discovering
importance of the concept of a focus of a conic, and the discovery of the
related concept of a directrix.
Al-Kuhi
An instrument for drawing conic sections was first described in 1000 CE
by the Islamic mathematician Al-Kuhi.
Omar Khayyám
Apollonius’s work was translated into Arabic (the technical language of
the time) and much of his work only survives through the Arabic version.
Persians found applications to the theory; the most notable of these was the
Persian mathematician and poet Omar Khayyám who used conic sections to
solve algebraic equations.
Europe
Johannes Kepler extended the theory of conics through the “principle of
continuity”, a precursor to the concept of limits. Girard Desargues and Blaise
Pascal developed a theory of conics using an early form of projective
geometry and this helped to provide impetus for the study of this new field.
In particular, Pascal discovered a theorem known as the hexagrammum
mysticum from which many other properties of conics can be deduced.
Meanwhile, René Descartes applied his newly discovered Analytic geometry
to the study of conics. This had the effect of reducing the geometrical
problems of conics to problems in algebra.
Properties
Just as two (distinct) points determine a line, five points determine a
conic. Formally, given any five points in the plane in general linear position,
meaning no three collinear, there is a unique conic passing through them,
which will be non-degenerate; this is true over both the affine plane and
projective plane. Indeed, given any five points there is a conic passing
through them, but if three of the points are collinear the conic will be
degenerate (reducible, because it contains a line), and may not be unique.
Irreducible conic sections are always “smooth”. This is important for many
applications, such as aerodynamics, where a smooth surface is required to
ensure laminar flow and to prevent turbulence.
Intersection at Infinity
An algebro-geometrically intrinsic form of this classification is by the
intersection of the conic with the line at infinity, which gives further insight
into their geometry:
• ellipses intersect the line at infinity in 0 points – rather, in 0 real
points, but in 2 complex points, which are conjugate;
• parabolas intersect the line at infinity in 1 double point,
corresponding to the axis – they are tangent to the line at infinity, and
close at infinity, as distended ellipses;
• hyperbolas intersect the line at infinity in 2 points, corresponding to
the asymptotes – hyperbolas pass through infinity, with a twist.
Going to infinity along one branch passes through the point at infinity
corresponding to the asymptote, then re-emerges on the other branch
at the other side but with the inside of the hyperbola (the direction of
curvature) on the other side – left vs. right (corresponding to the non-
orientability of the real projective plane) – and then passing through
the other point at infinity returns to the first branch. Hyperbolas can
thus be seen as ellipses that have been pulled through infinity and re-
emerged on the other side, flipped.
Degenerate Cases
There are five degenerate cases: three in which the plane passes through
apex of the cone, and three that arise when the cone itself degenerates to a
cylinder (a doubled line can occur in both cases). When the plane passes
through the apex, the resulting conic is always degenerate, and is either: a
point (when the angle between the plane and the axis of the cone is larger
than tangential); a straight line (when the plane is tangential to the surface of
the cone); or a pair of intersecting lines (when the angle is smaller than the
tangential). These correspond respectively to degeneration of an ellipse,
parabola, and a hyperbola, which are characterized in the same way by angle.
The straight line is more precisely a double line (a line with multiplicity 2)
because the plane is tangent to the cone, and thus the intersection should be
counted twice.
Where the cone is a cylinder, i.e. with the vertex at infinity, cylindric
sections are obtained; this corresponds to the apex being at infinity.
Cylindrical sections are ellipses (or circles), unless the plane is vertical
(which corresponds to passing through the apex at infinity), in which case
three degenerate cases occur: two parallel lines, known as a ribbon
(corresponding to an ellipse with one axis infinite and the other axis real and
non-zero, the distance between the lines), a double line (an ellipse with one
infinite axis and one axis zero), and no intersection (an ellipse with one
infinite axis and the other axis imaginary).
Cartesian Coordinates
In the Cartesian coordinate system, the graph of a quadratic equation in
two variables is always a conic section – though it may be degenerate, and all
conic sections arise in this way. The equation will be of the form
Ax2 + Bxy +Cy2 + Dx + Ey + F = 0 with A,B,C not all zero.
As scaling all six constants yields the same locus of zeros, one can
consider conics as points in the five-dimensional projective space P5.
Classification
Regular and degenerated conic sections can be distinguished based on
the determinant of AQ.
If det AQ = 0 , the conic is degenerate.
If Q is not degenerate, we can see what type of conic section it is by
computing the minor det A33 (that is, the determinant of the submatrix
resulting from removing the last row and the last column of AQ):
Centre
In the centre of the conic, the gradient of the quadratic form Q vanishes,
so:
We can calculate the centre by taking the first two rows of the associated
matrix AQ , multiplying each by (x, y, 1)T, setting both inner products equal to
0, and solving the system.
Tangents
Through a given point, P, there are generally two lines tangent to a
conic. Expressing P as a column vector, p, the two points of tangency are the
intersections of the conic with the line whose equation is
When P is on the conic, the line is the tangent there. When P is inside an
ellipse, the line is the set of all points whose own associated line passes
through P. This line is called the polar of the pole P with respect to the conic.
Just as P uniquely determines its polar line (with respect to a given
conic), so each line determines a unique P. This is thus an expression of
geometric duality between points and lines in the plane. As special cases, the
centre of a conic is the pole of the line at infinity, and each asymptote of a
hyperbola is a polar (a tangent) to one of its points at infinity.
Using the theory of poles and polars, the problem of finding the four
mutual tangents of two conics reduces to finding the intersection of two
conics.
Reduced Equation
The reduced equation of a conic section is the equation of a conic section
translated and rotated so that its centre lies in the centre of the coordinate
system and its axes are parallel to the coordinate axes. This is equivalent to
saying that the coordinates are moved to satisfy these properties.
Figure: If λ1 and λ2 are the eigenvalues of the matrix A33, the reduced equation can be written as
Plane (Geometry)
In mathematics, a plane is a flat, two-dimensional surface. A plane is the
two dimensional analogue of a point (zero-dimensions), a line (one-
dimension) and a space (three-dimensions). Planes can arise as subspaces of
some higher dimensional space, as with the walls of a room, or they may
enjoy an independent existence in their own right, as in the setting of
Euclidean geometry.
When working in two-dimensional Euclidean space, the definite article is
used, the plane, to refer to the whole space. Many fundamental tasks in
mathematics, geometry, trigonometry, graph theory and graphing are
performed in two-dimensional space, or in other words, in the plane.
Euclidean Geometry
Euclidean geometry is a mathematical system attributed to the
Alexandrian Greek mathematician Euclid, which he described in his textbook
on geometry: the Elements. Euclid’s method consists in assuming a small set
of intuitively appealing axioms, and deducing many other propositions
(theorems) from these. Although many of Euclid’s results had been stated by
earlier mathematicians, Euclid was the first to show how these propositions
could fit into a comprehensive deductive and logical system. The Elements
begins with plane geometry, still taught in secondary school as the first
axiomatic system and the first examples of formal proof. It goes on to the
solid geometry of three dimensions. Much of the Elements states results of
what are now called algebra and number theory, couched in geometrical
language.
For over two thousand years, the adjective “Euclidean” was unnecessary
because no other sort of geometry had been conceived. Euclid’s axioms
seemed so intuitively obvious that any theorem proved from them was
deemed true in an absolute, often metaphysical, sense. Today, however, many
other self-consistent non-Euclidean geometries are known, the first ones
having been discovered in the early 19th century. An implication of
Einstein’s theory of general relativity is that Euclidean space is a good
approximation to the properties of physical space only where the
gravitational field is weak.
The Elements
The Elements are mainly a systematization of earlier knowledge of
geometry. Its superiority over earlier treatments was rapidly recognized, with
the result that there was little interest in preserving the earlier ones, and they
are now nearly all lost. Books I–IV and VI discuss plane geometry. Many
results about plane figures are proved, e.g., If a triangle has two equal angles,
then the sides subtended by the angles are equal. The Pythagorean theorem is
proved.
Books V and VII–X deal with number theory, with numbers treated
geometrically via their representation as line segments with various lengths.
Notions such as prime numbers and rational and irrational numbers are
introduced. The infinitude of prime numbers is proved. Books XI–XIII
concern solid geometry. A typical result is the 1:3 ratio between the volume
of a cone and a cylinder with the same height and base.
The Parallel Postulate: If two lines intersect a third in such a way that
the sum of the inner angles on one side is less than two right angles, then the
two lines inevitably must intersect each other on that side if extended far
enough.
Axioms
Euclidean geometry is an axiomatic system, in which all theorems (“true
statements”) are derived from a small number of axioms. Near the beginning
of the first book of the Elements, Euclid gives five postulates (axioms) for
plane geometry, stated in terms of constructions (as translated by Thomas
Heath):
“Let the following be postulated”:
1. “To draw a straight line from any point to any point.”
2. “To produce [extend] a finite straight line continuously in a straight
line.”
3. “To describe a circle with any centre and distance [radius].”
4. “That all right angles are equal to one another.”
5. The parallel postulate: “That, if a straight line falling on two straight
lines make the interior angles on the same side less than two right
angles, the two straight lines, if produced indefinitely, meet on that
side on which are the angles less than the two right angles.”
Although Euclid’s statement of the postulates only explicitly asserts the
existence of the constructions, they are also taken to be unique.
The Elements also include the following five “common notions”:
1. Things that are equal to the same thing are also equal to one another.
2. If equals are added to equals, then the wholes are equal.
3. If equals are subtracted from equals, then the remainders are equal.
4. Things that coincide with one another equal one another.
5. The whole is greater than the part.
Parallel Postulate
To the ancients, the parallel postulate seemed less obvious than the
others. Euclid himself seems to have considered it as being qualitatively
different from the others, as evidenced by the organization of the Elements:
the first 28 propositions he presents are those that can be proved without it.
Many alternative axioms can be formulated that have the same logical
consequences as the parallel postulate. For example Playfair’s axiom states:
In a plane, through a point not on a given straight line, at most one line can be
drawn that never meets the given line.
Methods of Proof
Euclidean geometry is constructive. Postulates 1, 2, 3, and 5 assert the
existence and uniqueness of certain geometric figures, and these assertions
are of a constructive nature: that is, we are not only told that certain things
exist, but are also given methods for creating them with no more than a
compass and an unmarked straightedge. In this sense, Euclidean geometry is
more concrete than many modern axiomatic systems such as set theory,
which often assert the existence of objects without saying how to construct
them, or even assert the existence of objects that cannot be constructed within
the theory. Strictly speaking, the lines on paper are models of the objects
defined within the formal system, rather than instances of those objects.
For example a Euclidean straight line has no width, but any real drawn
line will. Though nearly all modern mathematicians consider nonconstructive
methods just as sound as constructive ones, Euclid’s constructive proofs often
supplanted fallacious nonconstructive ones—e.g., some of the Pythagoreans’
proofs that involved irrational numbers, which usually required a statement
such as “Find the greatest common measure of ...”
Euclid often used proof by contradiction. Euclidean geometry also
allows the method of superposition, in which a figure is transferred to another
point in space. For example, proposition I.4, side-angle-side congruence of
triangles, is proved by moving one of the two triangles so that one of its sides
coincides with the other triangle’s equal side, and then proving that the other
sides coincide as well. Some modern treatments add a sixth postulate, the
rigidity of the triangle, which can be used as an alternative to superposition.
System of Measurement and Arithmetic
Euclidean geometry has two fundamental types of measurements: angle
and distance. The angle scale is absolute, and Euclid uses the right angle as
his basic unit, so that, e.g., a 45-degree angle would be referred to as half of a
right angle. The distance scale is relative; one arbitrarily picks a line segment
with a certain length as the unit, and other distances are expressed in relation
to it.
A line in Euclidean geometry is a model of the real number line. A line
segment is a part of a line that is bounded by two end points, and contains
every point on the line between its end points. Addition is represented by a
construction in which one line segment is copied onto the end of another line
segment to extend its length, and similarly for subtraction.
Measurements of area and volume are derived from distances. For
example, a rectangle with a width of 3 and a length of 4 has an area that
represents the product, 12. Because this geometrical interpretation of
multiplication was limited to three dimensions, there was no direct way of
interpreting the product of four or more numbers, and Euclid avoided such
products, although they are implied, e.g., in the proof of book IX, proposition
20.
Euclid refers to a pair of lines, or a pair of planar or solid figures, as
“equal” (4σïò) if their lengths, areas, or volumes are equal, and similarly for
angles. The stronger term “congruent” refers to the idea that an entire figure
is the same size and shape as another figure. Alternatively, two figures are
congruent if one can be moved on top of the other so that it matches up with
it exactly. (Flipping it over is allowed.) Thus, for example, a 2x6 rectangle
and a 3x4 rectangle are equal but not congruent, and the letter R is congruent
to its mirror image. Figures that would be congruent except for their differing
sizes are referred to as similar.
8: Notation and Terminology
Later Work
Archimedes and Apollonius
Archimedes (ca. 287 BCE – ca. 212 BCE), a colourful figure about
whom many historical anecdotes are recorded, is remembered along with
Euclid as one of the greatest of ancient mathematicians. Although the
foundations of his work were put in place by Euclid, his work, unlike
Euclid’s, is believed to have been entirely original. He proved equations for
the volumes and areas of various figures in two and three dimensions, and
enunciated the Archimedean property of finite numbers.
Apollonius of Perga (ca. 262 BCE–ca. 190 BCE) is mainly known for his
investigation of conic sections.
17th Century: Descartes
René Descartes (1596–1650) developed analytic geometry, an alternative
method for formalizing geometry. In this approach, a point is represented by
its Cartesian (x, y) coordinates, a line is represented by its equation, and so
on.
In Euclid’s original approach, the Pythagorean theorem follows from
Euclid’s axioms. In the Cartesian approach, the axioms are the axioms of
algebra, and the equation expressing the Pythagorean theorem is then a
definition of one of the terms in Euclid’s axioms, which are now considered
theorems. The equation
Treatment of Infinity
Infinite Objects
Euclid sometimes distinguished explicitly between “finite lines” (e.g.,
Postulate 2) and “infinite lines” (book I, proposition 12). However, he
typically did not make such distinctions unless they were necessary. The
postulates do not explicitly refer to infinite lines, although for example some
commentators interpret postulate 3, existence of a circle with any radius, as
implying that space is infinite.
The notion of infinitesimally small quantities had previously been
discussed extensively by the Eleatic School, but nobody had been able to put
them on a firm logical basis, with paradoxes such as Zeno’s paradox
occurring that had not been resolved to universal satisfaction. Euclid used the
method of exhaustion rather than infinitesimals.
Later ancient commentators such as Proclus (410–485 CE) treated many
questions about infinity as issues demanding proof and, e.g., Proclus claimed
to prove the infinite divisibility of a line, based on a proof by contradiction in
which he considered the cases of even and odd numbers of points constituting
it.
At the turn of the 20th century, Otto Stolz, Paul du Bois-Reymond,
Giuseppe Veronese, and others produced controversial work on non-
Archimedean models of Euclidean geometry, in which the distance between
two points may be infinite or infinitesimal, in the Newton–Leibniz sense.
Fifty years later, Abraham Robinson provided a rigorous logical foundation
for Veronese’s work.
Infinite Processes
One reason that the ancients treated the parallel postulate as less certain
than the others is that verifying it physically would require us to inspect two
lines to check that they never intersected, even at some very distant point, and
this inspection could potentially take an infinite amount of time.
The modern formulation of proof by induction was not developed until
the 17th century, but some later commentators consider it implicit in some of
Euclid’s proofs, e.g., the proof of the infinitude of primes.
Supposed paradoxes involving infinite series, such as Zeno’s paradox,
predated Euclid. Euclid avoided such discussions, giving, for example, the
expression for the partial sums of the geometric series in IX.35 without
commenting on the possibility of letting the number of terms become infinite.
Logical Basis
Classical Logic
Euclid frequently used the method of proof by contradiction, and
therefore the traditional presentation of Euclidean geometry assumes classical
logic, in which every proposition is either true or false, i.e., for any
proposition P, the proposition “P or not P” is automatically true.
Modern Standards of Rigor
Placing Euclidean geometry on a solid axiomatic basis was a
preoccupation of mathematicians for centuries. The role of primitive notions,
or undefined concepts, was clearly put forward by Alessandro Padoa of the
Peano delegation at the 1900 Paris conference:
...when we begin to formulate the theory, we can imagine that the
undefined symbols are completely devoid of meaning and that
the unproved propositions are simply conditions imposed upon
the undefined symbols.
Then, the system of ideas that we have initially chosen is simply one
interpretation of the undefined symbols; but..this interpretation can be
ignored by the reader, who is free to replace it in his mind by another
interpretation.. that satisfies the conditions...
Logical questions thus become completely independent of empirical or
psychological questions...
The system of undefined symbols can then be regarded as the
abstraction obtained from the specialized theories that result when...the
system of undefined symbols is successively replaced by each of the
interpretations...
—Padoa, Essai d’une théorie algébrique des nombre entiers, avec une
Introduction logique à une théorie déductive qulelconque
That is, mathematics is context-independent knowledge within a
hierarchical framework. As said by Bertrand Russell:
If our hypothesis is about anything, and not about some one or
more particular things, then our deductions constitute
mathematics. Thus, mathematics may be defined as the subject in
which we never know what we are talking about, nor whether
what we are saying is true.
—Bertrand Russell, Mathematics and the metaphysicians
Such foundational approaches range between foundationalism and
formalism.
Axiomatic Formulations
Geometry is the science of correct reasoning on incorrect figures.
—George Polyá, How to Solve It, p. 208
• Euclid’s axioms: In his dissertation to Trinity College, Cambridge,
Bertrand Russell summarized the changing role of Euclid’s geometry
in the minds of philosophers up to that time. It was a conflict between
certain knowledge, independent of experiment, and empiricism,
requiring experimental input. This issue became clear as it was
discovered that the parallel postulate was not necessarily valid and its
applicability was an empirical matter, deciding whether the
applicable geometry was Euclidean or non-Euclidean.
• Hilbert’s axioms: Hilbert’s axiom s h ad the goal of ident ifying a
simple and complete set of independent axioms from which the most
important geometric theorems could be deduced. The outstanding
objectives were to make Euclidean geometry rigorous (avoiding
hidden assumptions) and to make clear the ramifications of the
parallel postulate.
• Birkhoff’s axioms: Birkhoff proposed four postulates for Euclidean
geometry that can be confirmed experimentally with scale and
protractor. The notions of angle and distance become primitive
concepts.
• Tarski‘s axioms: Tarski (1902–1983) and his students defined
elementary Euclidean geometry as the geometry that can be
expressed in first-order logic and does not depend on set theory for its
logical basis, in contrast to Hilbert’s axioms, which involve point
sets. Tarski proved that his axiomatic formulation of elementary
Euclidean geometry is consistent and complete in a certain sense:
there is an algorithm that, for every proposition, can be shown either
true or false. (This doesn’t violate Gödel’s theorem, because
Euclidean geometry cannot describe a sufficient amount of arithmetic
for the theorem to apply.) This is equivalent to the decidability of real
closed fields, of which elementary Euclidean geometry is a model.
Constructive Approaches and Pedagogy
The process of abstract axiomatization as exemplified by Hilbert’s
axioms reduces geometry to theorem proving or predicate logic. In contrast,
the Greeks used construction postulates, and emphasized problem solving.
For the Greeks, constructions are more primitive than existence
propositions, and can be used to prove existence propositions, but not vice
versa. To describe problem solving adequately requires a richer system of
logical concepts. The contrast in approach may be summarized:
• Axiomatic proof: Proofs are deductive derivations of propositions
from primitive premises that are ‘true’ in some sense. The aim is to
justify the proposition.
• Analytic proof: Proofs are non-deductive derivations of hypothesis
from problems. The aim is to find hypotheses capable of giving a
solution to the problem. One can argue that Euclid’s axioms were
arrived upon in this manner. In particular, it is thought that Euclid felt
the parallel postulate was forced upon him, as indicated by his
reluctance to make use of it, and his arrival upon it by the method of
contradiction.
Andrei Nicholaevich Kolmogorov proposed a problem solving basis for
geometry. This work was a precursor of a modern formulation in terms of
constructive type theory. This development has implications for pedagogy as
well.
If proof simply follows conviction of truth rather than contributing to its
construction and is only experienced as a demonstration of something already
known to be true, it is likely to remain meaningless and purposeless in the
eyes of students.
—Celia Hoyles,
The curricular shaping of students’ approach to proof
Line (Geometry)
The notion of line or straight line was introduced by ancient
mathematicians to represent straight objects with negligible width and depth.
Lines are an idealization of such objects. Thus, until seventeenth century,
lines were defined like this: “The line is the first species of quantity, which
has only one dimension, namely length, without any width nor depth, and is
nothing else than the flow or run of the point which [...] will leave from its
imaginary moving some vestige in length, exempt of any width. [...] The
straight line is that which is equally extended between its points”
Euclid described a line as “breadthless length”, and introduced several
postulates as basic unprovable properties from which he constructed the
geometry, which is now called Euclidean geometry to avoid confusion with
other geometries which have been introduced since the end of nineteenth
century (such as non-Euclidean geometry, projective geometry, and affine
geometry).
In modern mathematics, given the multitude of geometries, the concept
of a line is closely tied to the way the geometry is described. For instance, in
analytic geometry, a line in the plane is often defined as the set of points
whose coordinates satisfy a given linear equation, but in a more abstract
setting, such as incidence geometry, a line may be an independent object,
distinct from the set of points which lie on it.
When a geometry is described by a set of axioms, the notion of a line is
usually left undefined (a so-called primitive object). The properties of lines
are then determined by the axioms which refer to them. One advantage to this
approach is the flexibility it gives to users of the geometry. Thus in
differential geometry a line may be interpreted as a geodesic (shortest path
between points), while in some projective geometries a line is a 2-
dimensional vector space (all linear combinations of two independent
vectors). This flexibility also extends beyond mathematics and, for example,
permits physicists to think of the path of a light ray as being a line.
A line segment is a part of a line that is bounded by two distinct end
points and contains every point on the line between its end points. Depending
on how the line segment is defined, either of the two end points may or may
not be part of the line segment. Two or more line segments may have some of
the same relationships as lines, such as being parallel, intersecting, or skew.
Ray
If the concept of “order” of points of a line is defined, a ray, or half-line,
may be defined as well. A ray is part of a line which is finite in one direction,
but infinite in the other. It can be defined by two points, the initial point, A,
and one other, B. The ray is all the points in the line segment between A and
B together with all points, C, on the line through A and B such that the points
appear on the line in the order A, B, C.
The direction of the line is that of the vector b-a. Different choices of a
and b can yield the same line.
Projective Geometry
In projective geometry, a line is similar to that in Euclidean geometry but
has slightly different properties. In many models of projective geometry the
idea of the line rarely conforms to the notion of the “straight curve” as it is
visualised in Euclidean geometry. Elliptic geometry is a typical example of
when this happens.
Geodesics
The “straightness” of a line, interpreted as the property that it minimizes
distances between its points, can be generalized and leads to the concept of
geodesics in metric spaces.
Rotation (Mathematics)
In geometry and linear algebra, a rotation is a transformation in a plane
or in space that describes the motion of a rigid body around a fixed point. A
rotation is different from a translation, which has no fixed points, and from a
reflection, which “flips” the bodies it is transforming. A rotation and the
above-mentioned transformations are isometries; they leave the distance
between any two points unchanged after the transformation.
It is important to know the frame of reference when considering
rotations, as all rotations are described relative to a particular frame of
reference. In general for any orthogonal transformation on a body in a
coordinate system there is an inverse transformation which if applied to the
frame of reference results in the body being at the same coordinates.
For example in two dimensions rotating a body clockwise about a point
keeping the axes fixed is equivalent to rotating the axes counterclockwise
about the same point while the body is kept fixed.
Circle Group
In mathematics, the circle group, denoted by T, is the multiplicative
group of all complex numbers with absolute value 1, i.e., the unit circle in the
complex plane.
T = {z∈C:|z|=1}.
The circle group forms a subgroup of C, the multiplicative group of all
nonzero complex numbers. Since C is abelian, it follows that T is as well.
The circle group is also the group U(1) of 1 ×1 unitary matrices; these act on
the complex plane by rotation about the origin. The circle group can be
parametrized by the angle θ of rotation by
This is the exponential map for the circle group. The circle group plays a
central role in Pontryagin duality, and in the theory of Lie groups.
The notation T for the circle group stems from the fact that Tn (the direct
product of T with itself n times) is geometrically an n-torus. The circle group
is then a 1-torus.
Elementary Introduction
Figure: Multiplication on the circle group is equivalent to addition of angles
One way to think about the circle group is that it describes how to add
angles, where only angles between 0° and 360° are permitted. For example,
the diagram illustrates how to add 150° to 270°. The answer should be 150° +
270° = 420°, but when thinking in terms of the circle group, we need to
“forget” the fact that we have wrapped once around the circle. Therefore we
adjust our answer by 360° which gives 420° = 60° (mod 360°).
Another description is in terms of ordinary addition, where only numbers
between 0 and 1 are allowed (with 1 corresponding to a full rotation). To
achieve this, we might need to throw away digits occurring before the
decimal point. For example, when we work out 0.784 + 0.925 + 0.446, the
answer should be 2.155, but we throw away the leading 2, so the answer (in
the circle group) is just 0.155.
9: Topological and Analytic Structure
The circle group is more than just an abstract algebraic object. It has a
natural topology when regarded as a subspace of the complex plane. Since
multiplication and inversion are continuous functions on C, the circle group
has the structure of a topological group. Moreover, since the unit circle is a
closed subset of the complex plane, the circle group is a closed subgroup of C
(itself regarded as a topological group).
One can say even more. The circle is a 1-dimensional real manifold and
multiplication and inversion are real-analytic maps on the circle. This gives
the circle group the structure of a one-parametre group, an instance of a Lie
group. In fact, up to isomorphism, it is the unique 1-dimensional compact,
connected Lie group. Moreover, every n-dimensional compact, connected,
abelian Lie group is isomorphic to Tn.
Isomorphisms
The circle group shows up in a variety of forms in mathematics. We list
some of the more common forms here. Specifically, we show that
These representations are just the characters of the circle group. The
character group of T is clearly an infinite cyclic group generated byq1:
The irreducible real representations of the circle group are the trivial
representation (which is 1-dimensional) and the representations
taking values in SO(2). Here we only have positive integers n since the
representation p-n is equivalent to pn.
Besides just preserving length, rotations also preserve the angles between
vectors. This follows from the fact that the standard dot product between two
vectors u and v can be written purely in terms of length:
It follows that any length-preserving transformation in R3 preserves the
dot product, and thus the angle between vectors. Rotations are often defined
as linear transformations that preserve the inner product on R3. This is
equivalent to requiring them to preserve length.
Orthogonal and Rotation Matrices
Every rotation maps an orthonormal basis of R3 to another orthonormal
basis. Like any linear transformation of finite-dimensional vector spaces, a
rotation can always be represented by a matrix. Let R be a given rotation.
With respect to the standard basis (e1,e2 ,e3) of R3 the columns of R are
given by (Re1 ,Re2 ,Re3) . Since the standard basis is orthonormal, the
columns of R form another orthonormal basis. This orthonormality condition
can be expressed in the form
RTR=I
where RT denotes the transpose of R and I is the 3 × 3 identity matrix.
Matrices for which this property holds are called orthogonal matrices. The
group of all 3 × 3 orthogonal matrices is denoted O(3), and consists of all
proper and improper rotations.
In addition to preserving length, proper rotations must also preserve
orientation. A matrix will preserve or reverse orientation according to
whether the determinant of the matrix is positive or negative. For an
orthogonal matrix R, note that det RT = det R implies (det R)2 = 1 so that det
R = ±1. The subgroup of orthogonal matrices with determinant +1 is called
the special orthogonal group, denoted SO(3).
Thus every rotation can be represented uniquely by an orthogonal matrix
with unit determinant.
Moreover, since composition of rotations corresponds to matrix
multiplication, the rotation group is isomorphic to the special orthogonal
group SO(3).
Improper rotations correspond to orthogonal matrices with determinant
-1, and they do not form a group because the product of two improper
rotations is a proper rotation.
Group Structure
The rotation group is a group under function composition (or
equivalently the product of linear transformations). It is a subgroup of the
general linear group consisting of all invertible linear transformations of
Euclidean space.
Furthermore, the rotation group is nonabelian. That is, the order in which
rotations are composed makes a difference. For example, a quarter turn
around the positive x-axis followed by a quarter turn around the positive y-
axis is a different rotation than the one obtained by first rotating around y and
then x.
The orthogonal group, consisting of all proper and improper rotations, is
generated by reflections. Every proper rotation is the composition of two
reflections, a special case of the Cartan– Dieudonné theorem.
Axis of Rotation
Every nontrivial proper rotation in 3 dimensions fixes a unique 1-
dimensional linear subspace of R3 which is called the axis of rotation (this is
Euler’s rotation theorem).
Each such rotation acts as an ordinary 2-dimensional rotation in the plane
orthogonal to this axis. Since every 2-dimensional rotation can be represented
by an angle ϕ, an arbitrary 3-dimensional rotation can be specified by an axis
of rotation together with an angle of rotation about this axis. (Technically,
one needs to specify an orientation for the axis and whether the rotation is
taken to be clockwise or counterclockwise with respect to this orientation).
For example, counterclockwise rotation about the positive z-axis by
angle ϕ is given by
Given a unit vector n in R3 and an angle φ, let R(φ, n) represent a
counterclockwise rotation about the axis through n (with orientation
determined by n). Then
• R(0, n) is the identity transformation for any n
• R(j, n) = R(- j, - n)
• R(p + j, n) = R(p - j, - n).
Using these properties one can show that any rotation can be represented
by a unique angle φ in the range 0 ≤ φ ≤ π and a unit vector n such that
• n is arbitrary if φ = 0
• n is unique if 0 <φ<π
• n is unique up to a sign if φ = π (that is, the rotations R(π, ±n) are
identical).
Charts on SO(3)]
In mathematics, the special orthogonal group in three dimensions,
otherwise known as the rotation group SO(3), is a naturally occurring
example of a manifold. The various charts on SO(3) set up rival coordinate
systems: in this case there cannot be said to be a preferred set of parametres
describing a rotation. There are three degrees of freedom, so that the
dimension of SO(3) is three. In numerous applications one or other
coordinate system is used, and the question arises how to convert from a
given system to another.
The Space of Rotations
In geometry the rotation group is the group of all rotations about the
origin of three-dimensional Euclidean space R3 under the operation of
composition.[1] By definition, a rotation about the origin is a linear
transformation that preserves length of vectors (it is an isometry) and
preserves orientation (i.e. handedness) of space.
A length-preserving transformation which reverses orientation is called
an improper rotation. Every improper rotation of three-dimensional Euclidean
space is rotation followed by a reflection in a plane through the origin.
Composing two rotations results in another rotation; every rotation has a
unique inverse rotation; and the identity map satisfies the definition of a
rotation. Owing to the above properties, the set of all rotations is a group
under composition.
Moreover, the rotation group has a natural manifold structure for which
the group operations are smooth; so it is in fact a Lie group. The rotation
group is often denoted SO(3) for reasons explained below.
The space of rotations is isomorphic with the set of rotation operators
and the set of “orthonormal matrices with determinant +1”. It is also
isomorphic with the set of quaternions with their internal product, and also
equivalent to the set of rotation vectors, with a difficult internal composition
operation given by the product of their equivalent matrices.
Rotation vectors notation arise from the Euler’s rotation theorem which
states that any rotation in three dimensions can be described by a rotation by
some angle about some axis. Considering this, we can then specify the axis of
one of these rotations by two angles, and we can use the radius of the vector
to specify the angle of rotation. These vectors represent a ball in 3D with an
unusual topology.
This 3D solid sphere is equivalent to the surface of a 4D sphere, which is
also a 3D variety. For doing this equivalence, we will have to define how will
we represent a rotation with this 4D-embedded surface.
The elements of the rotation matrix are not all independent - as Euler’s
rotation theorem dictates, the rotation matrix has only three degrees of
freedom.
The rotation matrix has the following properties:
• A is a real, orthogonal matrix, hence each of its rows or columns
represents a unit vector.
• The eigenvalues of A are
{1, e±iθ} = {1, cos(θ) + i sin(θ), cos(θ) - i sin(θ)} where i is the
standard imaginary unit with the property i 2 = -1
• The determinant of A is +1, equivalent to the product of its
eigenvalues.
• The trace of A is 1+2cos(θ), equivalent to the sum of its eigenvalues.
The angle θwhich appears in the eigenvalue expression corresponds to
the angle of the Euler axis and angle representation.
The eigenvector corresponding with the eigenvalue of 1 is the
accompanying Euler axis, since the axis is the only (nonzero) vector which
remains unchanged by left-multiplying (rotating) it with the rotation matrix.
The above properties are equivalent to:
Albert, A.: Regression and the Moore-Penrose Pseudoinverse, Ac adem ic Press, New York, USA,
1972.
Bishop, R. L. and R. J.: Crittenden Geometry of Manifolds, Academic Press, New York, USA, 1964.
Bonola, R.: History of Non-Euclidean Geometry. Hafner Publishing Company, New York, USA, 1912.
Boothby, W. M.: An Introduction to Differentiable Manifolds and Riemannian Geometry, Academic
Press, New York, USA, 1975.
Cartan, E.: Geometry of Riemannian Spaces, Math Sc i Press, Br ooklin e, MA, USA, 1983.
Darling, R. W. R.: Differential Forms and Connections, Cambridge University Press, Cambridge, UK,
1994.
Do Carmo, M. P. : Differential Geometry of Curves and Surfaces, Prentice Hall, Englewood Cliffs, NJ,
USA, 1976.
———: Riemannian Geometry, Birkhauser, Boston, MA, USA, 1992.
Doolin, B. F. and C. F. Martin: Introduction to Differential Geometry for Engineers, Marcel Dekker,
Inc., New York, NY, USA, 1990.
Eisenhart, L. P. : An Introduction to Differential Geometry with Use of the Tensor Calculus, Princeton
University Press, Princeton, USA, 1940.
———: Riemannian Geometry, Princeton University Press, Princeton, USA, 1925.
Farebrother, R. W.: Linear Least Squares Computations, Marcel Dekker Inc., New York, USA, 1988.
Field, Judith V. : The Geometrical Work of Girard Desargues. Springer-Verlag, New York, 1987.
Gerretsen, J. C. H.: Lectures on Tensor Calculus and Differential Geometry,Academic Press, New
York, USA, 1982.
Gray, A.: Modern Differential Geometry of Curves and Surfaces, CRC Press, Boca Raton, FL, USA,
1993.
Gruber, M. H. J.: Regression Estimators: A Comparitive Study, Academic Press, San Diego CA, 1990.
Kendall, M. G.: A Course in the Geometry of n Dimensions, Hafner Publishing Company, New York,
USA, 1961.
Levi-Civita, T.: The Absolute Differential Calculus, Dover Publications, New York, USA, 1977.
Millman, R. S. and G. D. Parker: Elements of Differential Geometry,Dover Publications, New York,
USA, 2009.
Mlodinow, M.: Euclid’s window (the story of geometry from parallel lines to hyperspace), Harper &
Row, New York, USA, 1974.
Murray, M. K. and J. W. Rice: Differential Geometry and Statistics, Chapman & Hall, London, UK.
1993.
Nomizu, K. and T. Sasaki: Affine Differential Geometry, Cambridge University Press, Cambridge, UK,
1966.
Postnikov, M. M.: The Variational Theory of Geodesics, Dover, Publications, New York, 1983.
Robinson, E. A.: Least Squares Regression Analysis in Terms of Linear Algebra, Goose Pond Press,
Houston, USA, 1981.
Saville, D. J. and G. R. Wood: Statistical Methods: The Geometric Approach, Springer-Verlag, New
York, USA, 1991.
Weintraub , S . H . : Differential Forms: A Complement to Vector Calculus, Academic Press, San
Diego, CA, USA, 1997.