Вы находитесь на странице: 1из 8

EPGY Summer Institute

Special and General Relativity 2013


Lecture 4: Geometry of Spacetime
In this lecture a more sophisticated discussion of relativity, spacetime, and geometry will be introduced. Though
not entirely necessary to solve problems in relativity it will allow you to read more advanced treatments of
relativity (and impress your friends with complicated sounding terminology). The goal is to understand the
difference between the geometry of spacetime and the more familiar geometry of Euclidean geometry. So to begin
we need to make sure we are clear what we mean by ‘geometry’ in order to understand the difference.

Geometry
The mathematical area of geometry, as it will concern us, is the study of spaces and their properties. A space,
at the most abstract level, is a set of points. These points represent locations in the space. We will restrict
our discussion to physical space (the 3-dimensional space R3 that we reside in and later spacetime) and consider
the points to be physical locations. At this abstract level and considering only the set of points and subsets
of points (without any other restrictions or relations) the study of such structures is the mathematical topic of
topology. If we impose a little more structure and introduce a notion of ‘nearness’ (i.e. given two points can it be
determined if they are in the same neighborhood) we begin to get into the study of geometry. Suffice to say for
now, if neighborhoods of points can be defined and also a means to measure distances and directions a geometry
of the space can be discussed. The means by which distances are determined (via the metric) will be the central
concept that will concern us. In what follows only well defined, smooth spaces will be considered. The full study
of (differential) geometry and topology spends much time on pathological and problematic spaces.

Geometry of the flat Euclidean plane, R2


The simplest, most intuitive, geometry to explore first is that of flat two dimensional Euclidean space. This is the
elementary geometry learned in grade school and introduced, formally, by Euclid (325-265 BC) over two thousand
years ago. Though a brief review review follows, basic familiarity with Euclidean geometry is assumed. First, we
note that the two dimensional flat plane is the product space of the real line, R with itself, i.e. R2 ≡ R × R (every
element of the product space is parameterized by an element in the subspaces –basically x and y).
The core of Euclidean geometry is the 5 postulates set forth by Euclid for points, lines, and curves in the flat
plane. A number of axioms are given that are assumed evident from which the following postulates are stated as
evident.

1. A straight line can be drawn between any two points.


2. A straight line can be extended to any length.
3. Given any straight line segment, a circle can be drawn having the segment as radius and one endpoint as
center.
4. All right angles are equal (congruent) to one another.
5. Given a line and a point not on the line, it is possible to draw exactly one line through the given point
parallel to the line.

In elementary geometry courses these axioms and postulates are used to prove various theorems about geo-
metrical structures in the plane. An extremely important theorem that can be derived is the Pythagorean theorem
which was known before Euclid but stems from the postulates.
Pythagorean theorem: The square of the hypotenuse, h, of a right triangle is equal to the sum of
the squares of the other two sides (a and o).

h2 = o2 + a2

In the flat plane the Pythagorean theorem can be used to find the distance between any two points once
a coordinate system is established. A fundamental result is that the distance between two points in the plane
is independent of the coordinate system used as long as the axes are perpendicular. Put another way, if one
translates and/or rotates the coordinate system the distance between the two points remains the same. Put even
another way, the set of translations and rotations of the two dimensional plane leave the distance between any
two points in the plane invariant.

1
If a Cartesian coordinate system is established in a flat plane then the distance between any two points,
P1 = (x1 , y1 ), P2 = (x2 , y2 ), within it may be defined by,
p
∆l = ∆x2 + ∆y 2 where, ∆x = x2 − x1 , ∆y = y2 − y1
This is not the only way to define the distance between points, however it is the only form in the flat plane that
will concern us.
Example: Rotations in R2 Consider the act of rotating the flat plane by an angle θ in the coun-
terclockwise direction. We consider how the new unit vectors1 x̂′ , ŷ ′ , are represented in terms of the
old unit vectors, x̂, ŷ. This can be represented as an operator R̂(θ) which maps the coordinates (x, y)
into new coordinates (x′ , y ′ ) as follows
y x̂′ = x̂ cos θ + ŷ sin θ
y′
ŷ ′ = −x̂ sin θ + ŷ cos θ
x′
θ
θ
x
Of utmost importance is to note that the distance between two points does not change under this
transformation. The change in coordinate for each point is easily obtained by attaching subscripts 1
and 2 to the above formula. This results in the new coordinate representation of the two points,
∆x′ = ∆x cos θ + ∆y sin θ
∆y ′ = −∆x sin θ + ∆y cos θ
Returning to the Pythagorean theorem in the new coordinates we have,
∆l′2 = ∆x′2 + ∆y ′2 = (∆x cos θ + ∆y sin θ)2 + (−∆x sin θ + ∆y cos θ)2
= ∆x2 cos2 θ + 2∆x cos θ∆y sin θ + ∆y 2 sin2 θ
+∆x2 sin2 θ − 2∆x cos θ∆y sin θ + ∆y 2 cos2 θ
= ∆x2 (cos2 θ + sin2 θ) + ∆y 2 (cos2 θ + sin2 θ) = ∆x2 + ∆y 2 = ∆l2
The distance is invariant under rotations.
To repeat, the distance between two points in two dimensional space can be calculated using the Pythagorean
Theorem in two dimensions. The squared distance, d2 , between two points is,
d2 = (x2 − x1 )2 + (y2 − y1 )2 = ∆x2 + ∆y 2 .
No matter which coordinate system we use the distance always comes out the same2 .
Consider a map of the San Francisco Bay Area and the distance between Stanford and Berkeley (on this small
scale we can neglect the curvature of the Earth). The distance between them remains the same and notice that
the Pythagorean theorem always gives the same results; the distance between Stanford and Berkeley is always 64
km.

1A unit vector is a vector of length 1, it is essentially just a pointer vector. We have x̂ · x̂ = ŷ · ŷ = 1 and x̂ · ŷ = 0.
2 N.B. In the above equation, and in what follows it is important to note that the terms on the right hand side of the equation,
∆x2 for e.g., actually represent (∆x)2 . This is conventional shorthand that sometimes leads to confusion. Whenever ∆a2 is shown
you should understand it as (∆a)2 unless otherwise stated).

2
The Pythagorean theorem in three dimensions is essentially the same, it is obtained by simply adding on the
extra dimension squared,

d2 = ∆x2 + ∆y 2 + ∆z 2

Again if we set our coordinate axes in space now we still observe the distance to be 64 km. What is important
is that this formula defines the geometry of the space. Knowing how to measure distances in a space tells just
about everything about the geometry.
To reiterate, we see that no matter how we set up our coordinate systems the distance between two points
remains the same. This is a very important property and we term it the Invariance of Distance in three
dimensional space.

Analytic geometry
Thus far we have been discussing the basic geometry that you should be familiar with already. Let’s begin again
in a more formal, powerful, way of describing the geometry. This is based on the notion that in the constructions
in basic Euclidean geometry the tools that are necessary are to be able to measure distances and also to measure
angles. If we consider the flat plane as a vector space and examine vectors within this space we note that the
mathematical operation that yields distances and angles between vectors is the dot product.
~ in R2 ,
Recall the two definitions of the dot product of two vectors ~v and w

~v · w
~ = vx wx + vy wy or with the same vector ~v · ~v = vx2 + vy2
~v · w
~ = |~v ||w|
~ cos θ

Consider two points in the plane with a Cartesian coordinate system (x, y) established. The positions can be
given by two position vectors, ~r1 , ~r2 , that point from the origin to each point respectively. The displacement
vector between these two points is defined by the difference of these two vectors, ∆~r = ~r2 − ~r1 and spans the
distance between the two points. Note well that the distance between points 1 and 2 is given by (the square root
of) the dot product of this displacement vector by itself. That is, labeling the distance as d1−2 , we have,
√ p
d1−2 = ∆~r · ∆~r = ∆x2 + ∆y 2 (1)

The dot product in the flat plane embodies the Pythagorean theorem.
In addition, to find the angle between two line segments we introduce two vectors parallel to each line and
the dot product yields cos θ between them. The dot product embodies the geometry of the flat plane within it.
The generalization to three dimensional Euclidean space should be clear

Inner products, the metric formula


Thus we see that for ordinary (i.e. Euclidean) two and three dimensional space the dot product of vectors is the
key relation that allows us to measure distances and relate directions. This product will be of central importance
when we discuss more unfamiliar spaces (e.g. spacetime). A more generalized concept of the dot product is
introduced in differential geometry and is called the inner product, the term dot product is usually only refers to
1,2, and 3 dimensional Euclidean spaces.
To understand the inner product let’s begin with a simple example of two dimensional spaces. Let’s begin
by positing that we can uniquely label each point in this two dimensional space by two coordinates (x1 , x2 ).
Displacement vectors can then be constructed that span between two points in this space, ∆~r, as mentioned
above. (Technically we would like to keep the two points very close together but we will ignore this subtlety for
now3 ). To find the distance between these two points we take the inner product of this vector with itself, this gives
the squared distance. Since our space may change from point to point the inner product may change at different
points. Also, it might be the case that at some points the two dimensions (x1 , x2 ) may not be perpendicular. To
incorporate these possibilities the inner product is expressed as,

∆~r · ∆~r = g11 (∆r1 )2 + g12 (∆r1 ∆r2 ) + g21 (∆r2 ∆r1 ) + g22 (∆r2 )2

In this expression the coefficients g are possible functions of x1 and x2 and may take (just about) any value. This
means that the dot product varies from point to point in this space. The set of these coefficients, {g}, is known
3 For those who like completeness and calculus, the reason that these two points should be considered infinitesimally separated (and

more appropriately expressed as d~ r ) is that the inner product


√ might vary over macroscopic distances. If so, one needs to integrate
R R
the inner product to find the distance, L = path dl = path d~ r . When we consider curved spacetimes in general relativity this
r · d~
will be the case.

3
as the metric and the whole expression is sometimes called the metric formula or equation. This general form is
what is known as a Reimannian metric since it has two powers of the vector components4 . We have introduced a
fair amount of math that will not be necessary as we proceed, so let’s restrict this metric formula to types that
will be of interest: we will not be interested in cases where the cross terms are present. I.e. we will only be
concerned with cases where the dimensions are perpendicular at each point in space. Thus the general form we
will concentrate on is,

∆~r · ∆~r = g11 (∆r1 )2 + g22 (∆r2 )2 = (length)2

The key point to understand here is that it is the metric {g} that determines the geometry of this space at
every point (i.e. the local geometry).

Metric for 2-dimensional Euclidean space in Cartesian coordinates


To find the metric formula for our simplest example we note that the inner product is just the dot product
given by the square of (1) above. We see this is obtained by simply setting gxx = 1 & gyy = 1 (and labeling
x1 = x, x2 = y) giving,

∆~r · ∆~r = (∆x)2 + (∆y)2 (2)

For three dimensional Euclidean space we would have 9 g’s but only the diagonal terms are present: ∆~r · ∆~r =
(∆x)2 + (∆y)2 + (∆z)2 . A space that has all diagonal coefficients that are constants are called flat (or Euclidean
spaces). Note that by redefining the coordinate any such metric can be converted into the form (2) above.

Introducing Spacetime
In Newtonian physics we have an absolute time, which can be viewed as a one-dimensional Euclidean space, R
(the number line), plus an absolute three dimensional space, R3 . Time and space are two separate spaces and
the movement in one space (time) is independent of movement through the other (space).
What the postulates of special relativity tell us is that movement through space is no longer independent
of movement through time. We conclude that space and time are no longer independent and that the rate of
movement through time depends on your movement through space (velocity).
The result of this is that space and time are linked; we can no longer treat them as separate or absolute.
To quote Minkowski (1909), “Henceforth space by itself, and time by itself, are doomed to fade away into mere
shadows, and only a union of the two will preserve an independent reality”.
To mathematically manifest this idea we introduce a new entity –spacetime– where time becomes the fourth
dimension tacked on to three dimensional space. What we want to do is incorporate this concept into our
understanding of SR and see if it is useful. To do this we need to describe the geometry of this new entity,
spacetime.
Recall that for three-dimensional space, the Pythagorean theorem told us how to measure the distance between
two points. This distance remains the same, or is invariant, when coordinate systems are translated or rotated
(a coordinate transformation). As mentioned in the previous section specifying how to measure between points
in a space gives almost all of the information to specify the geometry.
Our goal now is to extend this to spacetime. How do we extend this formula to measure (the equivalent of)
distance in this new space? Well the simplest idea is to follow what we do when we extend the formula for distance
from two to three dimensions – we add the square of the difference between the two points in the new dimension.
Naively, this would mean adding ∆t2 to the metric equation for three dimensions However, we must be careful.
The new dimension we plan to add is the time dimension, which has units of seconds whereas the unit of distance
in space is meters. Hence, it is mathematically inconsistent to simply add the squared time difference to the
metric equation. To rectify this we can include a constant, say k, to convert seconds to meters. (Equivalently, we
could measure all distances in time by multiplying all spatial distances by a constant to convert them to seconds).
Including this constant, the metric equation which results by simply ‘tacking on’ the time dimension becomes a
four-dimensional Euclidean (R4 ) metric equation,

∆s2 = k∆t2 + ∆x2 + ∆y 2 + ∆z 2 .

To distinguish the two cases we define the ‘distance’ in this new four-dimensional space as s (which we will
eventually call the interval), this is to contrast the label d for distance in space. The term Euclidean is simply
4 There are other types of metrics but we will not be concerned with those. All relativity uses Riemannian metric formulas

4
to express that this new four dimensional space is flat, i.e. distances between points are determined by the
Pythagorean Theorem for four dimensions.
2
The constant k is unknown except for its dimensions. It must have the units of m s2 in order to convert the
2 2
squared time dimension (sec ) to squared distance dimension (m ). This is required for mathematical consistency.

The big question here is: Is the geometry of four-dimensional spacetime Euclidean?, i.e. R4 ? To check we will
need to show that what we know, thus far, about special relativity is true in such a geometric space. Recall that
the principle of special relativity implies the constancy of the speed of light – we can check whether the speed of
light remains the same in different frames within this geometry.
To keep things simple let’s consider two frames of reference S and S ′ that have collinear x and x′ axes and
are in relative motion along this axis. If we define all events to occur on this axis, then y = z = 0 for all events
and we can just examine the two-dimensional spacetime. In this case our metric formula simplifies to,

∆s2 = k∆t2 + ∆x2 . (3)

We will consider two events that connected by a light ray. Say, event 1 is the emission of a light pulse at
x1 = x′1 = 0 and event 2 is the reception some distance away, at x2 &x′2 6= 0. This will give a value of ∆s that
must be the same value in both frames. Since these two events are separated by a light ray we must satisfy the
following constraints.
a) ∆t 6= ∆t′ and ∆x 6= ∆x′ since these two frames are in relative motion along the collinear axis.
∆x ′ ∆x′
b) vlight = c = ∆t must equal vlight =c= ∆t′ . This stems from the constancy of c.
c) The undetermined constant, k, must be the same in S and S ′ since it relates to the geometry of spacetime.
d) The inner product is the same in all frames: ∆s = ∆s′ . This is equivalent to the invariance of distance
discussed before for Euclidean spaces.
By d) we have then that,

∆s2 = ∆s2 −→ k∆t2 + ∆x2 = k∆t′2 + ∆x′2


∆x2
   
2 ′2 ∆x′2
−→ ∆t k + = ∆t k+ (4)
∆t2 ∆t′2

Writing the speed of light in the two frames as vlight and vlight we have,
 
2
2
∆t2 k + vlight

= ∆t′2 k + v ′ light
 
2
∆t 2 k + v light
=
k + v ′ 2light

∆t′2

From a) and c) we conclude that vlight 6= v ′ light which violates b). Thus the Euclidean metric (3) is not consistent
with the constancy of the speed of light and therefore the principle of special relativity. This metric does not
describe the geometry of spacetime.
This result could have been anticipated by a more subtle argument. Since (3) is exactly the same as the
Euclidean metric for the ordinary two dimensional plane, and that rotations in the x − y plane leave the distance
invariant, then ordinary rotations in the t − x plane leave the metric invariant in this case. However, we know
that light will be represented by 45o lines in a frame where light travels at c and a rotation to a different frame
would yield a light line that is not at 45o , i.e. not at c.
Thus, we are left with the task of finding the metric for spacetime. Let’s generalize (3) by including undeter-
mined constants gtt = a and gxx = b,

∆s2 = ka∆t2 + b∆x2 . (5)

To satisfy d) we must have


∆x2 ∆x′2
ka∆t2 + b∆x2 = ka∆t′2 + b∆x′2 −→ ∆t2 (ka + b ) = ∆t ′2
(ka + b )
∆t2 ∆t′2
and now forcing b) to be true we can express this as

∆t2 (ka + bc2 ) = ∆t′2 (ka + bc2 )

5
Since ∆t 6= ∆t′ and all of the terms in the parentheses are constants (ka + bc2 ) there is no way to solve this except
when both sides are zero. If the term in parenthesis vanishes then we get 0 = 0 which is at least consistent. Thus
we are left with,

ka + bc2 = 0 ka = −bc2

There are various ways to solve this but we start with k = c2 which leaves us with a = −b and,

∆s2 = c2 a∆t2 − a∆x2

We are left with an undetermined constant a. Recall that we can always redefine our coordinates to absorb this
constant, therefore we can simply set a = 1. What we are left is –what is claimed– is the geometry of spacetime
(in regions where there is no gravity) given by the metric

∆s2 = c2 ∆t2 − ∆x2 − ∆y 2 − ∆z 2 (6)


where we have returned the other two coordinates to give the full four-dimensional version. By evaluating the
interval (∆s) in any frame we know it must take the same form in any other frame –this is the power of this
approach. This property will be called the Invariance of the Interval. Clearly this expression looks a little
strange and will take some getting used to.
The task now is to verify that this geometry embodies the results of special relativity as we know them.
We constructed it to satisfy the constancy of c. Again, to keep things simple, we will restrict to two spacetime
dimensions and have two frames S and S ′ in relative motion along the x (or x′ ) axis.

Time dilation from the metric


To verify that this relation is consistent with time dilation let’s return to the light clock example we started with.
Alice is in frame S and Bob in S ′ and the two events in consideration are the emission of a light pulse from Bob’s
clock and the reception after the round trip. Since these two events occur at the same place for Bob, then we
have ∆x′ = 0 and ∆t′ = 1 ns. The speed of the clock as measured in Alice’s frame is v = ∆x ∆t . Let’s spell out all
of the measurements,

Alice in S : ∆t ∆x
Bob in S ′ : ∆t′ = 1 ns ∆x′ = 0

The interval between these two events is the same in both frames, giving,

∆s2 = c2 ∆t2 − ∆x2 = c2 ∆t′2


1 ∆x2
 
2 2
c ∆t 1 − 2 = c2 ∆t′2
c ∆t2
v2
 
∆t2 1 − 2 = ∆t′2
c
∆t′2 ∆t′
∆t2 = 2 −→ ∆t = q = γ∆t′
1 − vc2 v2
1 − c2

Thus we see that the time dilation is ‘built into’ the metric equation. This quick derivation should convince you
of the power of this approach.

Length contraction from the metric


Consider two observers in frame S, one at the origin (x = x′ = 0) and the other a distance L away as measured
in S and L′ as measured in frame S ′ . The same arrangement as in the previous example is established here. In
frame S Bob passes with speed v between observer A1 and A2 which takes a time ∆t = Lv in frame S. We will
define the two events in question as passing A1 (event 1) and Bob passing A2 (event 2). Note that these two
events occur at the same location in frame S ′ . Thus the specification of the two events is as follows,
L
A1 &A2 in S : ∆t = v ∆x = L
′ L′
Bob in S : ∆t = ′
v ∆x′ = 0

6
We can now invoke the invariance of the interval,

L L′
∆s2 = c2 ∆t2 − ∆x2 = c2 ∆t′2 −→ c2 ( )2 − L2 = c2 ( )2
v v
c2 c2 ′2 v 2
L2 ( − 1) = L −→ L2 (1 − 2 ) = L′2
v2 v 2r c
v2
L′ = L 1−
c2
This is the expression for length contraction we found before where L is the length measured at rest and L′ is
the length measured when moving.

(Lack of) simultaneity from the metric


The fact that two events in one frame are seen as simultaneous may not be seen as simultaneous in another frame
is very clearly seen through the invariance of the interval. Consider two events in frame S that are simultaneous,
that is, ∆t = 0, and say they are separated by ∆x. If the arrangement of S and S ′ is as before then we know
that ∆x 6= ∆x′ and the invariance of the interval gives,
∆s2 = −∆x2 = c2 ∆t′ − ∆x′
Thus ∆t 6= 0 and the two events do not appear simultaneous in S ′ .

Isometries of spacetime
So far we have verified that the main results of special relativistic kinematics is incorporated into the metric
of spacetime, (6) but we want to check some other properties that we expect to hold. First of all, we expect
some of the old properties of three dimensional space to still hold. These being the invariance of the distance
between points if we translate the coordinates or rotate the spatial coordinates, as discussed before. This is
evident since within the spacetime metric (6) there is the old Euclidean metric, which is invariant under spatial
translations and rotations. To see this examine the geometry at one set time, so that ∆t = 0. Then we are left
with −(∆x2 + ∆y 2 + ∆z 2 ) which is invariant under,
3 Translations of spatial coordinates: x → x + ǫ (same for y and z).
3 Rotations of spatial coordinates: about z axis which mixes x and y (and about y and x axes).
These are the allowed (six) transformations that leave the distance (metric formula) invariant for Euclidean
three-dimensional space. These are still true in four-dimensional spacetime. When we include time another
transformation we expect to keep the metric invariant is temporal translations, t → t + ǫ. All that matters in
terms of the metric is the change in time, thus the absolute zero doe not matter. This brings us to 7 transformations
that leave the interval invariant.
These transformations are important in the more mathematical study of geometry. Those transformations
that leave the interval invariant are called isometries. It turns out that we still have three more isometries to go
over. These last three are the analogues of rotations in space. These last three mix the time coordinate with each
of the spatial coordinates (t − x, t − y, t − z), they are called the Lorentz transformations and are of central
importance in special relativity. The ‘scrunching’ transformation we have been performing with our spacetime
diagrams are exactly Lorentz transformations in graphical form. Later, we will examine the mathematical form
of the Lorentz transformations. To conclude the 10 isometries that leave interval invariant are called the Poincare
group of isometries and consist of,
3 Translations of spatial coordinates: x → x + ǫ (same for y and z).
3 Rotations of spatial coordinates: about z axis which mixes x and y (and about y and x. axes).
1 Translations of time, t → t + ǫ.
3 Lorentz transformations: which mixes t and x (as well as t and y, t and z).
This makes for a total of 10 transformations. Just prior to Einsteins 1905 paper, Poincare demonstrated that
these transformations form a mathematical group, now known as the Poincare group of transformations. We will
not go into what forms a mathematical group or the properties of such in these lectures.

7
y y′ y
T̂x (ǫ) T̂y (ǫ) y T̂z (ǫ) T̂t (ǫ)

T̂t (ǫ)t → t + ǫ

~ǫ ~ǫ

x ′
x x x
3: Spatial Translations
z z ′
z 1: Temporal translation
z
y R̂y (θ) y
R̂x (θ) y R̂z (θ)
′ y′
y
3: Spatial Rotations
θ x′
x′
θ
x x x
θ
z z z′ z
z′
Λ̂x (β) ct ct′ Λ̂y (β) ct′ Λ̂z (β) ct ct′ 3: Lorentz transformations
ct
v
β= c

x′ y′ z′

x y z
y
Appendix: Set theory notation
z x
Since our focus is on spaces, which are, at their most basic, a collection of point, the mathematical machinery dealing with
collections of points, set theory, will need to be made familiar. A full treatment of set theory would take us too far astray
but the language and notation of the subject will help the discussion to follow be clear and succinct.
A set is a collection of objects of any type, the elements of the set.. As long as distinctions can be made between
elements of the collection classifications of the elements can be made. For the discussion here the examples of sets of
concern will have many constraints imposed upon the elements of the set. For now we label the elements of a set, S, by
a, b, c, .... There can be a finite or infinite number of elements and sets can be made of collection of sets. The relations will
be abbreviated by the following set of mathematical symbols.
∈ is a member of a ∈ S, ‘element a is a member of the set S’.
⊂ is a subset of A ⊂ B ‘the set A is a subset of the set B’5
= equals A = B ‘the elements of set A are identical to those in B and vice versa.’
/ *, 6=.
negations. If the above are not true, they are represented by a slash: ∈,
∪ union of sets A ∪ B ‘the set of elements that belong to A and/or B’
∩ intersection of sets A ∩ B ‘the set of elements that belong to A and B’
Other symbols are taken from logic and other mathematical areas will be employed as well,
∃ there exists ∄ there does not exist
∀ for all e.g. ‘x + y = z ∈ S ∀ x & y ∈ S.’ (x plus y is in S for all elements that are in S.)
→ implies, or therefore. 9 does not imply
s.t. such that e.g. ‘x and y are chosen s.t. x & y are negative.’

Examples of relevant sets


∅ The null set, the set that contains no elements.
R The set of real numbers.
Z The set of integers.
R0 The set of rational numbers. (The set of irrationals: R − R0 ).
Familiarity with these sets is assumed in the following. The following should be obvious,
Z ⊂ R. The integers are a subset of the real numbers.
/ R0 . Pi is not a rational number. And, Z R − R0 = ∅, the integers are not irrational real numbers.
π ∈
T

∀ x, y ∈ Z, x + y = z ∈ Z ‘the addition of any two integers is an integer.’


5 All elements of A are also in the set B but the converse is not true.

Вам также может понравиться