File r37/packages/geometry/geometry.tex artifact 7def30cf12 part of check-in a57e59ec0d


\documentclass{article}

\title{\geo\ : A Small Package for Mechanized (Plane) Geometry
Manipulations\\[20pt] Version 1.1}

\author{Hans-Gert Gr\"abe, Univ. Leipzig, Germany}

\date{September 7, 1998}

\newenvironment{code}{\tt \begin{tabbing}
\hspace*{1cm}\=\hspace*{1cm}\=\hspace*{1cm}\=
\hspace*{1cm}\=\hspace*{1cm}\=\kill
}{\end{tabbing}}
\newcommand{\formel}[1]{\[\begin{array}{l} #1\end{array}\]}
\newcommand{\iks}{{\bf x}}
\newcommand{\uhh}{{\bf u}}
\newcommand{\vau}{{\bf v}}
\newcommand{\geo}{{\sc Geometry}}
\newcommand{\gr}{{Gr\"obner}}
\newcommand{\xxyy}[2] {\noindent{\tt #1} \\\hspace*{1cm}
\parbox[t]{9cm}{#2} \\[6pt]}

\begin{document}

\maketitle

\section{Introduction}

Geometry is not only a part of mathematics with ancient roots but also
a vivid area of modern research. Especially the field of geometry,
called by some negligence ``elementary'', continues to attract the
attention also of the great community of leisure mathematicians.  This
is probably due to the small set of prerequisites necessary to
formulate the problems posed in this area and the erudition and non
formal approaches ubiquitously needed to solve them. Examples from
this area are also an indispensable component of high school
mathematical competitions of different levels upto the International
Mathematics Olympiad (IMO) \cite{IMO}.
\medskip

The great range of ideas involved in elementary geometry theorem
proving inspired mathematicians to search for a common toolbox that
allows to discover such geometric statements or, at least, to prove
them in a more unified way. These attempts again may be traced back
until ancient times, e.g., to Euclid and his axiomatic approach to
geometry. 

Axiomatic approaches are mainly directed towards the introduction of
coordinates that allow to quantify geometric statements and to use the
full power of algebraic and even analytic arguments to prove geometry
theorems. Different ways of axiomatization lead to different, even
non-commutative, {\em rings of scalars}, the basic domain of
coordinate values, see \cite{Wu:94}.

Taking rational, real or even complex coordinates for granted (as we
will do in the following) it turns out that geometry theorems may be
classified due to their symmetry group as statements in, e.g.,
projective, affine or Euclidean (Cartesian) geometry. Below such a
distinction will be important for the freedom to choose appropriate
coordinate systems.
\medskip

It may be surprising that tedious but mostly straightforward
manipulations of the algebraic counterparts of geometric statements
allow to prove many theorems in geometry with even ingenious ``true
geometric'' proofs. With the help of a Computer Algebra System
supporting algebraic manipulations this approach obtains new power.
The method is not automatic, since one often needs a good feeling how
to encode a problem efficiently, but mechanized in the sense that one
can develop a tool box to support this encoding and some very standard
tools to derive a (mathematically strong~!) proof from these encoded
data.

The attempts to algorithmize this part of mathematics found their
culmination in the 80's in the work of W.-T. Wu \cite{Wu:94} on ``the
Chinese Prover'' and the fundamental book \cite{Chou:88} of S.-C. Chou
who proved 512 geometry theorems with this mechanized method, see also
\cite{Chou:84}, \cite{Chou:90}, \cite{Wu:84a}, \cite{Wu:84b}.

Since the geometric interpretation of algebraic expressions depends
heavily on the properties of the field of scalars, we get another
classification of geometry theorems: Those with coordinate version
valid over the algebraically closed field ${\bf C}$ and those with
coordinate version valid (or may be even formulated) only over ${\bf
R}$. The latter statements include {\em ordered geometry}, that uses
the distinction between ``inside'' and ``outside'', since ${\bf C}$
doesn't admit monotone orderings.
\medskip

This package \geo, written in the algebraic mode of Reduce, should
provide the casual user with a couple of procedures that allow him/her
to mechanize his/her own geometry proofs. Together with the Reduce
built-in simplifier for rational functions, the {\tt solve} function,
and the \gr\ utilities\footnote{Unfortunately, the built in \gr\
package of Reduce doesn't admit enough flexibility for our purposes.}
of the author's package CALI \cite{CALI} (part of the Reduce library)
it allows for proving a wide range of theorems of unordered geometry,
see the examples below and in the test file {\tt geometry.tst}.

This package grew up from a course of lectures for students of
computer science on this topic held by the author at the Univ. of
Leipzig in fall 1996 and was updated after a similar lecture in spring
1998.

\section{Mechanizing Geometry Proving}

Most geometric statements are of the following form:
\begin{quote}
Given certain (more or less) arbitrarily chosen points and/or lines we
construct certain derived points and lines from them. Then the
(relative) position of these geometric objects is of a certain
specific kind regardless of the (absolute) position of the chosen
data. 
\end{quote}

To obtain evidence for such a statement (recommended before attempting
to prove it~!) one makes usually one or several drawings, choosing the
independent data appropriately and constructing the dependent ones out
of them (best with ruler and compass, if possible). A computer may be
helpful in such a task, since the constructions are purely algorithmic
and computers are best suited for algorithmic tasks. Given appropriate
data structures such construction steps may be encoded into {\em
functions} that afterwards need only to be called with appropriate
parameters.

Even more general statements may be transformed into such a form and
must be transformed to create drawings. This may sometimes involve
constructions that can't be executed with ruler and compass as, e.g.,
angle trisection in Morley's theorem or construction of a conic in
Pascal's theorem.

\subsection{Algorithmization of (plane) geometry}

The representation of geometric objects through coordinates is best
suited for both compact (finite) data encoding and regeometrization of
derived objects, e.g., through graphic output. Note that the target
language for realization of these ideas on a computer can be almost
every computer language and is not restricted to those supporting
symbolic computations. Different geometric objects may be collected
into a {\em scene}. Rapid graphic output of such a scene with
different parameters may be collected into animations or even
interactive drag-and-move pictures if supported by the programming
system. (All this is not (yet) supported by \geo.)
\medskip

We will demonstrate this approach on geometric objects, containing
points and lines, represented as pairs {\tt P:Point=$(p_{1},p_{2})$}
or tripels {\tt g:Line=$(g_{1}, g_{2}, g_{3})$} of a certain basic
type {\tt Scalar}, e.g., floating point reals. Here $g$ represents the
homogeneous coordinates of the line $\{(x,y)\, :\, g_{1}x + g_{2}y +
g_{3}=0\}$.  In this setting geometric constructions may be understood
as functions constructing new geometric objects from given ones.
Implementing such functions variables occur in a natural way as formal
parameters that are assigned with special values of the correct type
during execution.
\medskip

1) For example, the equation
\[(x-p_{1})(q_{2}-p_{2}) - (y-p_{2})(q_{1}-p_{1})=0\] 
 of the line through two given points $P=(p_{1},p_{2}),\,
Q=(q_{1},q_{2})$ yields the function
\begin{code}
pp\_line(P,Q:Point):Line == 
	$(q_{2}-p_{2},p_{1}-q_{1},p_{2}q_{1}-p_{1}q_{2})$
\end{code}
that returns the (representation of the) line through these two
points. In this function $P$ and $Q$ are neither special nor general
points but formal parameters of type {\tt Point}.

2) The (coordinates of the) intersection point of two lines may be
computed solving the corresponding system of linear equations. We get
a partially defined function, since there is no or a not uniquely
defined intersection point, if the two lines are parallel. In this
case our function terminates with an error message.
\begin{code}
intersection\_point(a,b:Line):Point == \+\\
d:=$a_{1}b_{2}-a_{2}b_{1}$;\\
if $d=0$ then error ``Lines are parallel''\\
else return $((a_{2}b_{3}-a_{3}b_{2})/d,(a_{3}b_{1}-a_{1}b_{3})/d)$
\end{code}
Again $a$ and $b$ are formal parameters, here of the type {\tt Line}.

3) In the same way we may define a line $l$ through a given point $P$ 
perpendicular to a second line $a$ as
\begin{code}
lot(P:Point,a:Line):Line == $(a_{2},-a_{1},a_{1}p_{2}-a_{2}p_{1})$
\end{code}
and a line through $P$ parallel to $a$ as
\begin{code}
par(P:Point,a:Line):Line == $(a_{1},a_{2},-a_{1}p_{1}-a_{2}p_{2})$
\end{code}

4) All functions so far returned objects with coordinates being
rational expressions in the input parameters, thus especially well
suited for algebraic manipulations. To keep this nice property we
introduce only the {\em squared Euclidean distance}
\begin{code}
sqrdist(P,Q:Point):Scalar == $(p_{1}-q_{1})^{2}+(p_{2}-q_{2})^{2}$
\end{code}

5) Due to the relative nature of geometric statements some of the
points and lines may be chosen arbitrarily or with certain
restrictions. Hence we need appropriate constructors for points and
lines given by their coordinates
\begin{code}
Point(a,b:Scalar):Point == $(a,b)$\\
Line(a,b,c::Scalar):Line == $(a,b,c)$
\end{code}
and also for a point on a given line. For this purpose we provide two
different functions
\begin{code}
choose\_Point(a:Line,u:Scalar):Point == \+\\
if $a_{2}=0$ then\+\\ 
if $a_{1}=0$ then error ``a is not a line''\\
else return $(-a_{3}/a_{1},u)$\-\\
else return $(u,-(a_{3}+a_1\,u)/a_{2})$
\end{code}
that chooses a point on a line $a$ and
\begin{code}
varPoint(P,Q:Point,u:Scalar):Point == 
$(u\,a_{1}+(1-u)\,b_{1},u\,a_{2}+(1-u)\,b_{2})$
\end{code}
that chooses a point on the line through two given points. The main
reason to have also the second definition is that $u$ has a well
defined geometric meaning in this case. For example, the midpoint of
$PQ$ corresponds to $u={1\over 2}$:
\begin{code}
midPoint(P,Q:Point):Point == varPoint(P,Q,1/2)
\end{code}

6) One can compose these functions to get more complicated geometric
objects as, e.g., the pedal point of a perpendicular
\begin{code}
pedalPoint(P:Point,a:Line):Point == intersection\_point(lot(P,a),a),
\end{code}
the midpoint perpendicular of $BC$ 
\begin{code}
mp(B,C:Point):Line == lot(midPoint(B,C),line(B,C)),
\end{code}
the altitude to $BC$ in the triangle $\Delta\,ABC$
\begin{code}
altitude(A,B,C:Point):Line == lot(A,line(B,C))
\end{code}
and the median line
\begin{code}
median(A,B,C:Point):Line == line(A,midPoint(B,C))
\end{code}

7) We can also test geometric conditions to be fulfilled, e.g., whether
two lines $a$ and $b$ are parallel or orthogonal
\begin{code}
parallel(a,b:Line):Boolean == $(a_{1}b_{2}-a_{2}b_{1}=0)$
\end{code}
resp.
\begin{code}
orthogonal(a,b:Line):Boolean == $(a_{1}b_{1}+a_{2}b_{2}=0)$
\end{code}
or whether a given point is on a given line
\begin{code}
point\_on\_line(P:Point,a:Line):Boolean ==
$(a_{1}p_{1}+a_{2}p_{2}+a_{3}=0)$
\end{code}
The corresponding procedures implemented in the package return the
value of the expression to be equated to zero instead of a boolean.

Even more complicated conditions may be checked as, e.g., whether
three lines have a point in common or whether three points are on a
common line. For a complete collection of the available procedures we
refer to the section \ref{description}.
\medskip

Note that due to the linearity of points and lines all procedures
considered so far return data with coordinates that are rational in
the input parameters. One can easily enlarge the ideas presented in
this section to handle also non linear objects as circles and angles,
compute intersection points of circles, tangent lines etc., if the
basic domain {\tt Scalar} admits to solve non-linear (mainly
quadratic) equations. Since non-linear equations usually have more
than one solution, branching ideas should be incorporated, too. For
example, intersecting a circle and a line the program should consider
both intersection points.

\subsection{Mechanized evidence of geometric statements}

With a computer and these prerequisites at hand one may obtain
evidence of geometric statements not only from plots but also
computationally, converting the statement to be checked into a
function depending on the variable coordinates as parameters and
plugging in different values for them. 

For example, the following function tests whether the three midpoint
perpendiculars in a triangle given by the coordinates of its vertices
$A,B,C$ pass through a common point
\begin{code}
test(A,B,C:Point):Boolean ==\\\>\>\> 
concurrent(mp(A,B,C),mp(B,C,A),mp(C,A,B))
\end{code}
Plugging in different values for $A,B,C$ we can verify the theorem for
many different special geometric configurations. Of course this is
not yet a {\bf proof}.
\medskip

Lets add another remark: {\tt Point} and {\tt Line} are not only the
basic data types of our geometry, but data type functions parametrized
by the data type {\tt Scalar}. To have the full functionality of our
procedures {\tt Scalar} must be a field with effective zero test.

\section{Geometry Theorems of Constructive Type}

Implementing the functions described above in a system, that admits
also symbolic computations, we can execute the same computations also
with symbolic values, i.e. taking a pure transcendental extension of
${\bf Q}$ as scalars. The procedures then return (simplified) symbolic
expressions that specialize under (almost all) substitutions of
``real'' values for these symbolic ones to the same values as if they
were computed by the original procedures with the specialized input.
This leads to the notion of {\em generic geometric configurations}. A
geometric statement holds in this generic configuration, i.e., the
corresponding symbolic expression simplifies to zero, if and only if
it is ``generically true'', i.e., holds for all special coordinate
values except degenerate ones.

\subsection{Geometric configurations of constructive type}

This approach is especially powerful, if all geometric objects
involved into a configuration may be constructed step by step and have
{\em rational} expressions in the algebraically independent variables
as symbolic coordinates.
\medskip

{\sc Definition: } We say that a geometric configuration is of {\em
constructive type}\footnote{This notion is different from
\cite{Chou:88}.}, if its generic configuration may be constructed step
by step in such a way, that the coordinates of each successive
geometric object may be expressed as rational functions of the
coordinates of objects already available or algebraically independent
variables, and the conclusion may be expressed as vanishing of a
rational function in the coordinates of the available geometric
objects.
\medskip

Substituting the corresponding rational expressions of the coordinates
of the involved geometric objects into the coordinate slots of newly
constructed objects and finally into the conclusion expression, we
obtain successively rational expressions in the given algebraically
independent variables.
\begin{quote}\it
A geometry theorem of constructive type is generically true if and
only if (its configuration is not contradictory and) the conclusion
expression simplifies to zero.
\end{quote}

Indeed, if this expression simplifies to zero, the algebraic version
of the theorem will be satisfied for all ``admissible'' values of the
parameters. If the expression doesn't simplify to zero, the theorem
fails for almost all such parameters.

Note that due to cancelation of denominators the domain of definition
of the simplified expression may be greater than the (common) domain
of definition of the different parts of the unsimplified
expression. The correct non degeneracy conditions describing
``admissibility'' may be collected during the computation. Collecting
up the zero expression indicates, that the geometric configuration is
contradictory. Hence the statement, that a certain geometric
configuration of constructive type is contradictory, is of
constructive type, too.

The package \geo\ provides procedures {\tt clear\_ndg(), print\_ndg()}
to manage and print these non degeneracy conditions and also a
procedure {\tt add\_ndg(d)} as a hook for their user driven
management.

\subsection{Some one line proofs}

Take independent variables $a_1,a_2,b_1,b_2,c_1,c_2$ and
\begin{code}
A:=Point(a1,a2);\ B:=Point(b1,b2);\ C:=Point(c1,c2); 
\end{code}
as the vertices of a generic triangle. We can prove the following
geometric statements about triangles computing the corresponding
(compound) symbolic expressions and proving that they simplify to
zero. Note that Reduce does simplification automatically.
\medskip

\noindent 1) The midpoint perpendiculars of $\Delta\,ABC$ pass through
a common point since
\begin{code}\+\>
concurrent(mp(A,B),mp(B,C),mp(C,A));
\end{code}
simplifies to zero.
\medskip

\noindent 2) The intersection point of the midpoint perpendiculars
\begin{code}\+\>
M:=intersection\_point(mp(A,B),mp(B,C));
\end{code}
is the center of the circumscribed circle since
\begin{code}\+\>
sqrdist(M,A) - sqrdist(M,B);	
\end{code}
simplifies to zero.
\medskip

\noindent 3) {\em Euler's line}: 
\begin{quote}
The center $M$ of the circumscribed circle, the orthocenter $H$ and
the barycenter $S$ are collinear and $S$ divides $MH$ with ratio 1:2.
\end{quote}
Compute the coordinates of the corresponding points
\begin{code}\+\>
M:=intersection\_point(mp(a,b,c),mp(b,c,a));\\
H:=intersection\_point(altitude(a,b,c),altitude(b,c,a));\\
S:=intersection\_point(median(a,b,c),median(b,c,a));
\end{code}
and then prove that
\begin{code}\+\>
		collinear(M,H,S);\\
                sqrdist(S,varpoint(M,H,2/3));
\end{code}
both simplify to zero.
\medskip

\noindent 4) {\em Feuerbach's circle}: 
\begin{quote}
The midpoint $N$ of $MH$ is the center of a circle that passes through
nine special points, the three pedal points of the altitudes, the
midpoints of the sides of the triangle and the midpoints of the upper
parts of the three altitudes.
\end{quote}
\begin{code}\+\>
         N:=midpoint(M,H);\\[8pt]

                sqrdist(N,midpoint(A,B))-sqrdist(N,midpoint(B,C));\\
                sqrdist(N,midpoint(A,B))-sqrdist(N,midpoint(H,C));\\[8pt]

         D:=intersection\_point(pp\_line(A,B),pp\_line(H,C));\\
                sqrdist(N,midpoint(A,B))-sqrdist(N,D);
\end{code}
Again the last expression simplifies to zero thus proving the theorem.

\section{Non-linear Geometric Objects}

\geo\ provides several functions to handle angles and circles as
non-linear geometric objects.

\subsection{Angles and bisectors}

(Oriented) angles between two given lines are presented as tangens of
the difference of the corresponding slopes. Since
\[\tan(\alpha-\beta) = \frac{\tan(\alpha)-\tan(\beta)}{1+
\tan(\alpha)\, \tan(\beta)}\] 
we get for the angle between two lines $g,h$
\begin{code}\>
l2\_angle(g,h:Line):Scalar == $\frac{g_2h_1-g_1h_2}{g_1h_1+g_2h_2}$
\end{code}

Note that in unordered geometry we can't distinguish between inner and
outer angles. Hence we cannot describe (rationally) the parameters of
the angle bisector of a triangle. For a point $P$ the equation 
\begin{code}\>
l2\_angle(pp\_line(A,B),pp\_line(P,B)) =\\\>\>\>
	l2\_angle(pp\_line(P,B),pp\_line(C,B))
\end{code}
i.e., $\angle\,ABP=\angle\,PBC$, describes the condition to be located
on either the inner or outer bisector of $\angle\,ABC$. Clearing
denominators yields a procedure
\begin{code}\>
point\_on\_bisector(P,A,B,C)
\end{code}
that returns on generic input a polynomial of (total) degree 4 and
quadratic in the coordinates of $P$ that describes the condition for
$P$ to be on (either the inner or the outer) bisector of
$\angle\,ABC$.
\medskip

With some more effort one can also employ such indirect geometric
descriptions. For example, we can prove the following unordered
version of the bisector intersection theorem.
\medskip

\noindent 5) There are four common points on the three bisector pairs
of a given triangle $\Delta\,ABC$. Indeed, due to Cartesian symmetry
we may choose a special coordinate system with origin $A$ and (after
scaling) $x$-axes unit point $B$. The remaining point $C$ is
arbitrary. Then the corresponding generic geometric configuration is
described with two independent parameters $u_1,u_2$ -- the coordinates
of $C$:
\begin{code}\>
A:=Point(0,0); B:=Point(1,0); C:=Point(u1,u2);
\end{code}
A point {\tt P:=Point(x1,x2)} is an intersection point of three
bisectors iff it is a common zero of the polynomial system
\begin{code}
polys:=\{\>\>\+\+ point\_on\_bisector(P,A,B,C),\\ 
point\_on\_bisector(P,B,C,A),\\
point\_on\_bisector(P,C,A,B)\},
\end{code}
i.e., of the polynomial system
\begin{quote}
$\hspace*{-2ex}\{\ {x_1}^{2}\,u_2 -2\,x_1\,x_2\,u_1 +2\,x_1\,x_2
-2\,x_1\,u_2 -{x_2}^{2}\,u_2 +2\,x_2\,u_1 -2 \,x_2 +u_2,\\
2\,{x_1}^{2}\,u_1\,u_2 -{x_1}^{2}\,u_2 -2\,x_1\,x_2\,{u_1}^{2}
+2\,x_1\,x_2\,u_1 +2\,x_1\,x_2\,{u_2}^{2} -2\,x_1\,{u_1}^{2}\,u_2
-2\,x_1\,{u_2}^{3} -2\,{x_2}^{2}\,u_1\, u_2 +{x_2}^{2}\,u_2
+2\,x_2\,{u_1}^{3} -2\,x_2\,{u_1}^{2} +2\,x_2\,u_1\,{u_2}^{2} -2\,x_2
\,{u_2}^{2} +{u_1}^{2}\,u_2 +{u_2}^{3},\\ {x_1}^{2}\,u_2
-2\,x_1\,x_2\,u_1 -{x_2}^{2}\,u_2\}$
\end{quote}
with indeterminates $x_1,x_2$ over the coefficient field ${\bf
Q}(u_1,u_2)$. A \gr\ basis computation with CALI
\begin{code}\>\+
load cali;\\
setring(\{$x_1,x_2$\},\{\},lex);\\
setideal(polys,polys);\\
gbasis polys;
\end{code}
yields the following equivalent system:
\begin{quote}
$\hspace*{-2ex}\{\ 4\,{x_2}^{4}\,u_2 -8\,{x_2}^{3}\,{u_1}^{2}
+8\,{x_2}^{3}\,u_1 -8\,{x_2}^{3}\,{u_2}^{2 }
+4\,{x_2}^{2}\,{u_1}^{2}\,u_2 -4\,{x_2}^{2}\,u_1\,u_2
+4\,{x_2}^{2}\,{u_2}^{3} -4\,{ x_2}^{2}\,u_2 +4\,x_2\,{u_2}^{2}
-{u_2}^{3},\\ 2\,x_1\,u_1\,{u_2}^{2} -x_1\,{u_2}^{2} +2\,{
x_2}^{3}\,u_2 -4\,{x_2}^{2}\,{u_1}^{2} +4\,{x_2}^{2}\,u_1
-2\,{x_2}^{2}\,{u_2}^{2} -2\, x_2\,{u_1}^{2}\,u_2 +2\,x_2\,u_1\,u_2
-2\,x_2\,u_2 -u_1\,{u_2}^{2} +{u_2}^{2}\} $
\end{quote}
The first equation has 4 solutions in $x_2$ and each of them may be
completed with a single value for $x_1$ determined from the second
equation. Hence the system $polys$ has four generic solutions
corresponding to the four expected intersection points.  The solutions
have algebraic coordinates of degree 4 over the generic field of
scalars ${\bf Q}(u_1,u_2)$ and specialize to the correct ``special''
intersection points for almost all values for the parameters $u_1$ and
$u_2$.

Although it is hard to give an explicit description through radicals
of these symbolic values, one can compute with them knowing their
minimal polynomials. Since in this situation $x_2$ is the distance
from $P$ to the line $AB$, we can prove that each of the four points
has equal distance to each of the 3 lines through two vertices of
$\Delta\,ABC$, i.e., that these points are the centers of its incircle
and the three excircles. First we compute the differences of the
corresponding squared distances
\begin{code}\>\+
con1:=sqrdist(P,pedalpoint(p,pp\_line(A,C)))-x2\^{}2;\\
con2:=sqrdist(p,pedalpoint(p,pp\_line(B,C)))-x2\^{}2;
\end{code}
The numerator of each of these two expressions should simplify to zero
under the special algebraic values of $x_1,x_2$. This may be verified
computing their normal forms with respect to the above \gr\ basis: 
\begin{code}\>\+
con1 mod gbasis polys;\\
con2 mod gbasis polys;
\end{code}
Note that \cite{Wu:94} proposes also a constructive proof for the
bisector intersection theorem:

Start with $A,B$ and the intersection point $P$ of the bisectors
through $A$ and $B$. Then $g(AC)$ and $g(BC)$ are symmetric to $g(AB)$
wrt.\ $g(AP)$ and $g(BP)$ and $P$ must be on their bisector:
\begin{code}\>\+
A:=Point(0,0); B:=Point(1,0); P:=Point(u1,u2);\\
l1:=pp\_line(A,B);\\
l2:=symline(l1,pp\_line(A,P));\\
l3:=symline(l1,pp\_line(B,P));\\[6pt]

point\_on\_bisector(P,A,B,intersection\_point(l2,l3));
\end{code}
As desired the last expression simplifies to zero.


\subsection{Circles}

The package \geo\ supplies two different types for encoding circles.
The first type is {\tt Circle1} that stores the pair $(M,s)$, the
center and the squared radius of the circle. The implementation of
{\tt point\_on\_circle1(P,c)} and \linebreak {\tt p3\_circle1(A,B,C)}
is almost straightforward. The latter function finds the circle
through 3 given points, computing its center as the intersection point
of two midpoint perpendiculars.

For purposes of analytic geometry it is often better to work with the
representation {\tt Circle} derived from the description of the circle
as the set of points $(x,y)$ for which the expression
\[(x-m_1)^2+(y-m_2)^2-r^2 = (x^2+y^2) -2\,m_1\,x -2\,m_2\,y
+m_1^2+m_2^2-r^2 \] 
vanishes. We use homogeneous coordinates {\tt k:Circle}
$=(k_1,k_2,k_3,k_4)$ for the circle
\[k:=\{\,(x,y)\ :\ k_1*(x^2+y^2)+k_2*x+k_3*y+k_4 = 0\}\]
since they admit denominator free computations and include also lines
as special circles with infinite radius: The line $g=(g_1,g_2,g_3)$ is
the circle $(0,g_1,g_2,g_3)$.

Its easy to derive formulas {\tt circle\_center(k)} for the center of
the circle $k$ and {\tt circle\_sqradius(k)} for its squared radius.
It is also straightforward to test {\tt point\_on\_circle(P,k)}.  The
parameters of the circle {\tt p3\_circle(A,B,C)} through 3 given
points 
\begin{code}\>
A:=Point($a_1,a_2$); B:=Point($b_1,b_2$); C=Point($c_1,c_2$); 
\end{code}
may be obtained from a nontrivial solution of the corresponding
homogeneous linear system with coefficient matrix
\[\left(\begin{array}{cccc} 
a_1^2+a_2^2 & a_1 & a_2 & 1 \\ b_1^2+b_2^2 & b_1 & b_2 & 1 \\
c_1^2+c_2^2 & c_1 & c_2 & 1 \\
\end{array}\right)
\] 
The condition that 4 points are on a common circle then may be
expressed as
\begin{code}\>
p4\_circle(A,B,C,D) == point\_on\_circle(D,p3\_circle(A,B,C));
\end{code}
For generic points $A,B,C,D$ this yields a polynomial $p_4$ of degree
4 in their coordinates.

Note that this condition is equivalent to the circular angle theorem:
For generic points $A,B,C,D$
\begin{code}\+\>
  u:=angle(pp\_line(A,D),pp\_line(B,D));\\
  v:=angle(pp\_line(A,C),pp\_line(B,C));\\
  (num(u)*den(v)-den(u)*num(v));
\end{code}
yields the same condition $p_4$. The common denominator {\tt
den(u)*den(v)} corresponds to the degeneracy condition that either
$A,B,C$ or $A,B,D$ are collinear.
\medskip

This condition is also equivalent to {\em Ptolemy's theorem}:
\begin{quote}
For points $A,B,C,D$ are (in that order) on a circle iff
\[l(AB)*l(CD)+l(AD)*l(BC) = l(AC)*l(BD),\]
i.e., the sum of the products of the lengths of opposite sides of the
cyclic quadrilateral $ABCD$ equals the product of the lengths of its
diagonals. 
\end{quote}
For an elementary proof see \cite[2.61]{Coxeter:67}. To get a
mechanized proof with the tools developed so far we are faced with
several problems. First the theorem invokes distances and not their
squares. Second the theorem uses the order of the given points.
Unordered geometry can't even distinguish between sides and diagonals
of a quadrilateral.

The fist problem may be solved by repeated squaring. Denoting the
lengths appropriately we get step by step
\[\begin{array}{c}
p\cdot r + q\cdot s = t\cdot u\\
(p\,r)^2+(q\,s)^2-(t\,u)^2 = -(2\,p\,q\,r\,s)\\
((p\,r)^2+(q\,s)^2-(t\,u)^2)^2 - (2\,p\,q\,r\,s)^2 = 0
\end{array}\]
arriving at an expression that contains only squared distances. This
expression 
\[{\tt poly:= }\ p^4\,r^4 - 2\,p^2\,q^2\,r^2\,s^2 -
2\,p^2\,r^2\,t^2\,u^2 + q^4\,s^4 - 2\,q^2\, s ^2\,t^2\,u^2 + t^4\,u^4
\] 
is symmetric in pairs of opposite sides thus solving also the second
problem. Substituting the corresponding squared distances of generic
points $A,B,C,D$ we obtain exactly the square of the condition $p_4$. 
\medskip

As for bisector coordinates the coordinates of intersection points of
a circle and a line generally can't be expressed rationally in terms
of the coordinates of the circles. For a generic circle {\tt c:=
Circle($c_1,c_2,c_3,c_4$)} and a generic line {\tt
d:=Line($d_1,d_2,d_3$)} we may solve the line equation for $y$ and
substitute the result into the circle equation to get a single
polynomial $q(x)$ of degree 2 with zeroes being the $x$-coordinate of
the two intersection points of {\tt c} and {\tt d}:
\begin{code}\>\+
vars:=\{x,y\};\\
polys:=\{c1*(x\^{}2+y\^{}2)+c2*x+c3*y+c4, d1*x+d2*y+d3\};\\
s:=solve(second polys,y);\\
q:=num sub(s,first polys);
\end{code}
$q:={x}^{2}\,c_1\,({d_1}^{2} +{d_2}^{2}) +x\,(2\,c_1\,d_1\,d_3 +
c_2\,{d_2}^{2} -c_3\,d_1\,d_2 ) +(c_1\,{d_3}^{2} -c_3\,d_2\,d_3
+c_4\,{d_2}^{2})$
\medskip

In many cases {\tt d} is the line through a specified point {\tt P:=
Point($p_1,p_2$)} on the circle. Fixing these coordinates as generic
ones we get the algebraic relations
\begin{code}\>
polys:=\{point\_on\_line(P,d), point\_on\_circle(P,c)\};
\end{code}
\formel{\{d_1\,p_1 +d_2\,p_2 +d_3, c_1\,{p_1}^{2} +c_1\,{p_2}^{2}
 +c_2\,p_1 +c_3\,p_2 +c_4\}}
between the coordinates of $c, d$ and $P$. This dependency may be
removed solving these equations for $d_3$ and $c_4$. In the new
coordinates the polynomial $q(x)$ factors
\begin{code}\>\+
s:=solve(polys,\{d3,c4\});\\
factorize sub(s,q);
\end{code}
into $x-p_1$ and a second factor that is linear in $x$. This yields
the coordinates for the intersection point of {\tt c} and {\tt d}
different from {\tt P} that are saved into a function {\tt
other\_cl\_point(P,c,d)}. Similarly we computed the coordinates of the
second intersection point of two circles $c_1$ and $c_2$ passing
through a common point {\tt P} and saved into a function {\tt
other\_cc\_point(P,c1,c2)}. 

Also conditions on the coordinates of a circle and a line resp.\ two
circles to be tangent may be derived in a similar way.
\medskip

\noindent 6) These functions admit a constructive proof of {\em
Miquels theorem}:
\begin{quote} Let $\Delta\,ABC$ be a triangle. Fix arbitrary points
$P,Q,R$ on the sides $AB, BC, AC$. Then the three circles through each
vertex and the chosen points on adjacent sides pass through a common
point.
\end{quote}
Take as above 
\begin{code}\>
A:=Point(0,0); B:=Point(1,0); C:=Point(c1,c2);
\end{code}
Generic points on the sides may be introduced with three auxiliary
indeterminates: 
\begin{code}\>\+
P:=choose\_pl(pp\_line(A,B),u1);\\
Q:=choose\_pl(pp\_line(B,C),u2);\\
R:=choose\_pl(pp\_line(A,C),u3);
\end{code}
Then 
\begin{code}\>
X:=other\_cc\_point(P,p3\_circle(A,P,R),p3\_circle(B,P,Q));
\end{code}
is the intersection point of two of the circles different from {\tt P}
(its generic coordinates contain 182 terms) and since
\begin{code}\>
point\_on\_circle(X,p3\_circle(C,Q,R));
\end{code}
simplifies to zero the third circle also passes through {\tt X}.

\section{Geometry Theorems of Equational Type}

As already seen in the last section non-linear geometric conditions
are best given through implicit polynomial dependency conditions on
the coordinates of the geometric objects. In this more general setting
a geometric statement may be translated into a {\em generic geometric
configuration}, involving different geometric objects with coordinates
depending on (algebraically independent) variables $\vau = (v_1,
\ldots, v_n)$, a system of {\em polynomial conditions} $F= \{f_1,
\ldots, f_r\}$ expressing the implicit geometric conditions and a
polynomial $g$ encoding the geometric conclusion, such that, for a
certain polynomial non degeneracy condition $h$, the following holds:
\begin{quote}\it
The geometric statement is true iff for all non degenerate correct
special geometric configurations, i.e., with coordinates, obtained
from the generic ones by specialization $v_i\mapsto c_i$ in such a
way, that $f({\bf c})=0$ for all $f\in F$ but $h({\bf c})\neq 0$, the
conclusion holds, i.e., $g({\bf c})$ vanishes.
\end{quote}
Denoting by $Z(F)$ the set of zeroes of the polynomial system $F$ and
writing $Z(h)=Z(\{h\})$ for short, we arrive at {\em geometry theorems
of equational type}, that may be shortly stated in the form
\[Z(F)\setminus Z(h) \subseteq Z(g).\]
Over an algebraically closed field, e.g. ${\bf C}$, this is equivalent
to the ideal membership problem
\[g\cdot h\in rad\ I(F),\]
where $rad\ I(F)$ is the radical of the ideal generated by $F$. Even
if $h$ is unknown a detailed analysis of the different components of
the ideal $I(F)$ allows to obtain more insight into the geometric
problem.
\medskip

Note the symmetry between $g$ and $h$ in the latter formulation of
geometry theorems. This allows to derive {\em non degeneracy
conditions} for a given geometry theorem of equational type from the
stable ideal quotient
\[h\in rad\ I(F):g^\infty.\]
Since every element of this ideal may serve as non degeneracy
condition there is no weakest condition among them, if the ideal is
not principal.

\subsection{Dependent and independent variables}

Let $S=R[v_1,\ldots,v_n]$ be the polynomial ring in the given
variables over the field of scalars $R$. The polynomial system $F$
describes algebraic dependency relations between these variables in
such a way that the values of some of the variables may be chosen
(almost) arbitrarily whereas the remaining variables are determined
upto a finite number of values by these choices.
\medskip

A set of variables $\uhh\subset\vau$ is called {\em independent} wrt.\
the ideal $I=I(F)$ iff $I\cap R[\uhh]=(0)$, i.e., the variables are
algebraically independent modulo $I$. If $\uhh$ is a maximal subset
with this property the remaining variables $\iks=\vau\setminus\uhh$
are called {\em dependent}.

Although a maximal set of independent variables may be read off from a
\gr\ basis of $I$ there is often a natural choice of dependent and
independent variables induced from the geometric problem. $\uhh$ is a
maximal independent set of variables iff $F$ has a finite number of
solutions as polynomial system in $\iks$ over the generic scalar field
$R(\uhh)$. In many cases this may be proved with less effort than
computing a \gr\ basis of $I$ over $S$.

If $F$ has an infinite number of solutions then $\uhh$ was independent
but not maximal. If $F$ has no solution then $\uhh$ was not
independent. 

\subsection{Geometry theorems of linear type}

We arrive at a particularly nice situation in the case when $F$ is a
non degenerate quadratic linear system of equations in $\iks$ over
$R(\uhh)$. Such geometry theorems are called {\em of linear type}.

In this case there is a unique (rational) solution $\iks = \iks(\uhh)$
that may be substituted for the dependent variables into the geometric
conclusion $g=g(\iks,\uhh)$. We obtain as for geometry theorems of
constructive type a rational expression in $\uhh$ and 
\begin{quote}\it
the geometry theorem holds (under the non degeneracy condition
$h=det(F)\in R[\uhh]$, where $det(F)$ is the determinant of the linear
system $F$) iff this expression simplifies to zero.
\end{quote}


\noindent 7) As an example consider the {\em theorem of Pappus}:
\begin{quote}
Let $A,B,C$ and $P,Q,R$ be two triples of collinear points. Then the
intersection points $g(AQ)\wedge g(BP), g(AR)\wedge g(CP)$ and
$g(BR)\wedge g(CQ)$ are collinear.
\end{quote}
The geometric conditions put no restrictions on $A,B,P,Q$ and one
restriction on each $C$ and $R$. Hence we may take as generic
coordinates
\begin{code}\>\+
A:=Point(u1,u2); B:=Point(u3,u4); C:=Point(x1,u5);\\
P:=Point(u6,u7); Q:=Point(u8,u9); R:=Point(u0,x2);
\end{code}
with $u_0,\ldots,u_9$ independent and $x_1,x_2$ dependent, as
polynomial conditions
\begin{code}\>
F:=\{collinear(A,B,C), collinear(P,Q,R)\};
\end{code}
and as conclusion
\begin{code}\+\+
con:=collinear(\\
	intersection\_point(pp\_line(A,Q),pp\_line(P,B)),\\
	intersection\_point(pp\_line(A,R),pp\_line(P,C)),\\
	intersection\_point(pp\_line(B,R),pp\_line(Q,C)));
\end{code}
a rational expression with 462 terms. The polynomial conditions are
linear in $x_1,x_2$ and already separated. Hence
\begin{code}\>\+
sol:=solve(polys,\{x1,x2\});\\
sub(sol,con);
\end{code}
proves the theorem since the expression obtained from $con$
substituting the dependent variables by their rational expressions in
$\uhh$ simplifies to zero.
\medskip

As for most theorems of linear type the linear system may be solved
``geometrically'' and the whole theorem may be translated into a
constructive geometric statement:
\begin{code}\>\+
A:=Point(u1,u2); B:=Point(u3,u4);\\ 
P:=Point(u6,u7); Q:=Point(u8,u9);\\[6pt] 

C:=choose\_pl(pp\_line(A,B),u5);\\
R:=choose\_pl(pp\_line(P,Q),u0);\\[6pt]

con:=collinear(\+\\
	intersection\_point(pp\_line(A,Q),pp\_line(P,B)),\\
	intersection\_point(pp\_line(A,R),pp\_line(P,C)),\\
	intersection\_point(pp\_line(B,R),pp\_line(Q,C)));
\end{code}

\subsection{Geometry theorems of non-linear type}

Lets return to the general situation of a polynomial system $F\subset
S$ that describes algebraic dependency relations, a subdivision
$\vau=\iks\cup\uhh$ of the variables into dependent and independent
ones, and the conclusion polynomial $g(\iks,\uhh)\in S$. The set of
zeros $Z(F)$ may be decomposed into irreducible components that
correspond to prime components $P_\alpha$ of the ideal $I=I(F)$
generated by $F$ over the ring $S=R[\iks,\uhh]$.

Since $P_\alpha\supset I$ the variables $\uhh$ may become dependent
wrt.\ $P_\alpha$. Prime components where $\uhh$ remains independent
are called {\em generic}, the other components are called {\em
special}.  Note that each special component contains a non zero
polynomial in $R[\uhh]$. Multiplying them all together yields a non
degeneracy condition $h=h(\uhh)\in R[\uhh]$ on the independent
variables such that a zero $P\in Z(F)$ with $h(P)\neq 0$ necessarily
belongs to one of the generic components. Hence they are the
``essential'' components and we say that the geometry theorem is {\em
generically true}, when the conclusion polynomial $g$ vanishes on all
these generic components.
\medskip

If we compute in the ring $S'=R(\uhh)[\iks]$, i.e., consider the
independent variables as parameters, exactly the generic components
remain visible. Indeed, this corresponds to a localization of $S$ by
the multiplicative set $R[\uhh]\setminus\{0\}$. Hence the geometry
theorem is generically true iff $g\in rad(I)\cdot S'$, i.e. $g$
belongs to the radical of the ideal $I$ in this special extension of
$S$. A sufficient condition can be derived from a \gr\ basis $G$ of
$F$ with the $\uhh$ variables as parameters: Test whether $g\ mod\ G
=0$, i.e., the normal form vanishes. More subtle examples may be
analyzed with the \gr\ factorizer or more advanced techniques from the
authors package CALI, \cite{CALI}.
\medskip

\noindent 8) As an application we consider the following nice theorem
from \cite[ch. 4, \S\ 2]{Coxeter:67} about Napoleon triangles:
\begin{quote}
Let $\Delta\,ABC$ be an arbitrary triangle and $P,Q$ and $R$ the third
vertex of equilateral triangles erected externally on the sides $BC,
AC$ and $AB$ of the triangle. Then the lines $g(AP), g(BQ)$ and
$g(CR)$ pass through a common point, the {\em Fermat point} of the
triangle $\Delta\,ABC$.
\end{quote}

A mechanized proof again will be faced with the difficulty that
unordered geometry can't distinguish between different sides wrt.\ a
line. A straightforward formulation of the geometric conditions starts
with independent coordinates for $A,B,C$ and dependent coordinates for
$P,Q,R$. W.l.o.g. we may fix the coordinates in the following way:
\begin{code}\+\>
A:=Point(0,0); B:=Point(0,2); C:=Point(u1,u2);\\
P:=Point(x1,x2); Q:=Point(x3,x4); R:=Point(x5,x6);
\end{code}
There are 6 geometric conditions for the 6 dependent variables.
\begin{code}\+\+
polys:=\{\>\>sqrdist(P,B)-sqrdist(B,C), sqrdist(P,C)-sqrdist(B,C),\\ 
	sqrdist(Q,A)-sqrdist(A,C), sqrdist(Q,C)-sqrdist(A,C),\\ 
	sqrdist(R,B)-sqrdist(A,B), sqrdist(R,A)-sqrdist(A,B)\};
\end{code}
\formel{
{x_1}^{2} +{x_2}^{2} -4\,x_2 -{u_1}^{2} -{u_2}^{2} +4\,u_2\\ 
{x_1}^{2} -2\,x_1\,u_1 +{ x_2}^{2} -2\,x_2\,u_2 +4\,u_2 -4\\ 
{x_3}^{2} +{x_4}^{2} -{u_1}^{2} -{u_2}^{2}\\ 
{x_3}^{2} -2\,x_3\,u_1 +{x_4}^{2} -2\,x_4\,u_2\\ 
{x_5}^{2} +{x_6}^{2} -4\,x_6\\ 
{x_5}^{2} +{x_6}^{2} -4}
These equations may be divided into three groups of two quadratic
relations for the coordinates of each of the points $P,Q,R$. Each of
this pairs has (only) two solutions, the inner and the outer triangle
vertex, since it may easily be reduced to a quadratic and a linear
equation, the line equation of the corresponding midpoint
perpendicular. Hence the whole system has 8 solutions and by geometric
reasons the conclusion
\begin{code}\>
con:=concurrent(pp\_line(A,P), pp\_line(B,Q), pp\_line(C,R));
\end{code}
will hold on at most two of them. Due to the special structure the
interreduced polynomial system is already a \gr\ basis and hence can't
be split by the \gr\ factorizer. A full decomposition into isolated
primes yields four components over $R(\uhh)$, each corresponding to a
pair of solutions over the algebraic closure. On one of them the
conclusion polynomial reduces to zero thus proving the geometry
theorem. 
\begin{code}\>\+
vars:=\{x1,x2,x3,x4,x5,x6\};\\
setring(vars,\{\},lex);\\
iso:=isolatedprimes polys;\\[6pt]

for each u in iso collect con mod u;
\end{code}
With a formulation as in \cite[p.~123]{Chou:88}, that uses oriented
angles, we may force all Napoleon triangles to be erected on the {\em
same} side (internally resp.\ externally) and prove a more general
theorem as above. Taking isosceles triangles with equal base angles
and (due to one more degree of freedom) $x_5$ as independent the
conclusion remains valid:
\begin{code}
polys2:=\{\>\>sqrdist(P,B)-sqrdist(P,C),\+\+\\ 
	sqrdist(Q,A)-sqrdist(Q,C),  \\
	sqrdist(R,A)-sqrdist(R,B), \\
	num(p3\_angle(R,A,B)-p3\_angle(P,B,C)), \\
	num(p3\_angle(Q,C,A)-p3\_angle(P,B,C))\};\-\-\\[6pt]

sol:=solve(polys2,\{x1,x2,x3,x4,x6\});\\
sub(first sol,con);
\end{code}
again simplifies to zero. Note that the new theorem is of linear type.

\section{The Procedures Supplied by \geo}\label{description}

This section contains a short description of all procedures available
in \geo. We refer to the data types {\tt Scalar, Point, Line, Circle1}
and {\tt Circle} described above. Booleans are represented as extended
booleans, i.e.\ the procedure returns a {\tt Scalar} that is zero iff
the condition is fulfilled. In some cases also a non zero result has a
geometric meaning. For example, {\tt collinear(A,B,C)} returns the
signed area of the corresponding parallelogram.
\bigskip

\xxyy{angle\_sum(a,b:Scalar):Scalar }{Returns $\tan(\alpha+\beta)$, if
$a=\tan(\alpha), b=\tan(\beta)$.} 
\xxyy{altitude(A,B,C:Point):Line }{The altitude from $A$ onto
$g(BC)$. } 
\xxyy{c1\_circle(M:Point,sqr:Scalar):Circle}{The circle with given
center and sqradius.}  
\xxyy{cc\_tangent(c1,c2:Circle):Scalar}{Zero iff $c_1$ and $c_2$ are
tangent.}  
\xxyy{choose\_pc(M:Point,r,u):Point}{Chooses a point on the circle
around $M$ with radius $r$ using its rational parametrization with
parameter  $u$.}
\xxyy{choose\_pl(a:Line,u):Point }{Chooses a point on $a$ using
parameter $u$.} 
\xxyy{Circle(c1,c2,c3,c4:Scalar):Circle}{The {\tt Circle} constructor.} 
\xxyy{Circle1(M:Point,sqr:Scalar):Circle1}{The {\tt Circle1}
constructor. } 
\xxyy{circle\_center(c:Circle):Point}{The center of $c$.} 
\xxyy{circle\_sqradius(c:Circle):Point}{The sqradius of $c$.} 
\xxyy{cl\_tangent(c:Circle,l:Line):Scalar}{Zero iff $l$ is tangent to
$c$.} 
\xxyy{collinear(A,B,C:Point):Scalar}{Zero iff $A,B,C$ are on a common
line. In general the signed area of the parallelogram spanned by
$\vec{AB}$ and $\vec{AC}$. }
\xxyy{concurrent(a,b,c:Line):Scalar}{Zero iff $a,b,c$ have a common
point.} 
\xxyy{intersection\_point(a,b:Line):Point}{The intersection point of
the lines $a,b$. }
\xxyy{l2\_angle(a,b:Line):Scalar}{Tangens of the angle between $a$ and
$b$. } 
\xxyy{Line(a,b,c:Scalar):Line}{The {\tt Line} constructor.}
\xxyy{lot(P:Point,a:Line):Line}{The perpendicular from $P$ onto $a$.}
\xxyy{median(A,B,C:Point):Line}{The median line from $A$ to $BC$.}
\xxyy{midpoint(A,B:Point):Point}{The midpoint of $AB$. }
\xxyy{mp(B,C:Point):Line}{The midpoint perpendicular of $BC$.}
\xxyy{orthogonal(a,b:Line):Scalar}{zero iff the lines $a,b$ are
orthogonal. }
\xxyy{other\_cc\_point(P:Point,c1,c2:Circle):Point}{ $c_1$ and $c_2$
intersect at $P$. The procedure returns the second intersection
point. }
\xxyy{other\_cl\_point(P:Point,c:Circle,l:Line):Point}{$c$ and $l$
intersect at $P$. The procedure returns the second intersection
point.}
\xxyy{p3\_angle(A,B,C:Point):Scalar}{Tangens of the angle between
$\vec{BA}$ and $\vec{BC}$. }
\xxyy{p3\_circle(A,B,C:Point):Circle\ {\rm or\ }\\
p3\_circle1(A,B,C:Point):Circle1}{The circle through 3 given points. }
\xxyy{p4\_circle(A,B,C,D:Point):Scalar}{Zero iff four given points
are on a common circle. }
\xxyy{par(P:Point,a:Line):Line}{The line through $P$ parallel to
$a$. } 
\xxyy{parallel(a,b:Line):Scalar}{Zero iff the lines $a,b$ are
parallel. }
\xxyy{pedalpoint(P:Point,a:Line):Point}{The pedal point of the
perpendicular from $P$ onto $a$.} 
\xxyy{Point(a,b:Scalar):Point}{The {\tt Point} constructor.}
\xxyy{point\_on\_bisector(P,A,B,C:Point):Scalar}{Zero iff $P$ is a
point on the (inner or outer) bisector of the angle $\angle\,ABC$.}  
\xxyy{point\_on\_circle(P:Point,c:Circle):Scalar\ {\rm or\ }\\
point\_on\_circle1(P:Point,c:Circle1):Scalar}{Zero iff $P$ is on the
circle $c$.} 
\xxyy{point\_on\_line(P:Point,a:Line):Scalar}{Zero iff $P$ is on the
line $a$. }
\xxyy{pp\_line(A,B:Point):Line}{The line through $A$ and $B$.}
\xxyy{sqrdist(A,B:Point):Scalar}{Square of the distance between $A$
and $B$.} 
\xxyy{sympoint(P:Point,l:Line):Point}{The point symmetric to $P$ wrt.\
the line $l$.} 
\xxyy{symline(a:Line,l:Line):Line}{The line symmetric to $a$ wrt.\ the
line $l$.} 
\xxyy{varpoint(A,B:Point,u):Point}{The point $D=u\cdot A+(1-u)\cdot
B$. }

\noindent \geo\ supplies as additional tools the functions
\bigskip

\xxyy{extractmat(polys,vars)}{Returns the coefficient matrix of the
list of equations $polys$ that are linear in the variables $vars$. }
\xxyy{red\_hom\_coords(u:\{Line,Circle\})}{Returns the reduced
homogeneous coordinates of $u$, i.e., divides out the content. }



\section{More Examples}

Here we give a more detailed explanation of some of the examples
collected in the test file {\tt geometry.tst} and give a list of
exercises. Their solutions can be found in the test file, too.

\subsection{Theorems that can be translated into theorems of
constructive or linear type}

There are many geometry theorems that may be reformulated as theorems
of constructive type.
\medskip


\noindent 9) The affine version of {\em Desargue's theorem}: 
\begin{quote}
If two triangles $\Delta\,ABC$ and $\Delta\,RST$ are in similarity
position,\linebreak i.e., $g(AB)\,\|g(RS),\ g(BC)\|g(ST)$ and
$g(AC)\|g(RT)$, then $g(AR),\linebreak g(BS)$ and $g(CT)$ pass through
a common point (or are parallel).
\end{quote}

The given configuration may be constructed step by step in the
following way: Take $A,B,C,R$ arbitrarily, choose $S$ arbitrarily on
the line through $R$ parallel to $g(AB)$ and $T$ as the intersection
point of the lines through $R$ parallel to $g(AC)$ and through $S$
parallel to $g(BC)$.
\begin{code}\>\+
A:=Point(a1,a2); B:=Point(b1,b2);\\ 
C:=Point(c1,c2); R:=Point(d1,d2);\\
S:=choose\_pl(par(R,pp\_line(A,B)),u);\\
T:=intersection\_point(\\\>\>
	par(R,pp\_line(A,C)),par(S,pp\_line(B,C)));\\[6pt]

con:=concurrent(pp\_line(A,R),pp\_line(B,S),pp\_line(C,T));
\end{code}
Another proof may be obtained translating the statement into a theorem
of linear type. Since the geometric conditions put no restrictions on
$A,B,C,R$, one restriction on $S$ ($g(AB)\|g(RS)$) and two
restrictions on $T$ ($g(BC)\|g(ST),\,$ $g(AC)\|g(RT)$),
we may take as generic coordinates
\begin{code}\>\+
A:=Point(u1,u2); B:=Point(u3,u4); C:=Point(u5,u6);\\
R:=Point(u7,u8); S:=Point(u9,x1); T:=Point(x2,x3);
\end{code}
with $u_1,\ldots,u_9$ independent and $x_1,x_2,x_3$ dependent, as
polynomial conditions
\begin{code}\+\+
polys:=\{\>\>parallel(pp\_line(R,S),pp\_line(A,B)),\\
	parallel(pp\_line(S,T),pp\_line(B,C)),\\
	parallel(pp\_line(R,T),pp\_line(A,C))\};
\end{code}
and as conclusion
\begin{code}\>
con:=concurrent(pp\_line(A,R),pp\_line(B,S),pp\_line(C,T));
\end{code}
The polynomial conditions are linear in $x_1,x_2,x_3$ and thus 
\begin{code}\>\+
sol:=solve(polys,\{x1,x2,x3\});\\
sub(sol,con);
\end{code}
proves the theorem since the expression obtained from $con$
substituting the dependent variables by their rational expressions in
$\uhh$ simplifies to zero.
\medskip

The general version of {\em Desargue's theorem}:
\begin{quote}
The lines $g(AR),\ g(BS)$ and $g(CT)$ pass through a common point iff
the intersection points $g(AB)\wedge g(RS),\ g(BC)\wedge g(ST)$ and
$g(AC)\wedge g(RT)$ are collinear.
\end{quote}
may be reduced to the above theorem by a projective transformation
mapping the line through the three intersection points to infinity.
Its algebraic formulation
\begin{code}\>\+
A:=Point(0,0); B:=Point(0,1); C:=Point(u5,u6);\\
R:=Point(u7,u8); S:=Point(u9,u1); T:=Point(u2,x1);\\[6pt]
con1:=collinear(\+\\intersection\_point(pp\_line(R,S),pp\_line(A,B)),\\
	intersection\_point(pp\_line(S,T),pp\_line(B,C)),\\
	intersection\_point(pp\_line(R,T),pp\_line(A,C)));\-\\

con2:=concurrent(pp\_line(A,R),pp\_line(B,S),pp\_line(C,T));
\end{code}
contains a polynomial $con_2$ linear in $x_1$ and a rational function
$con_1$ with numerator quadratic in $x_1$ that factors as
\[{\rm num}(con_1)=con_2\cdot {\rm collinear}(R,S,T)\]
thus also proving the general theorem.
\medskip

\noindent 10) Consider the following theorem about the {\em Brocard
points} (\cite[p.~336]{Chou:88})
\begin{quote}
Let $\Delta\,ABC$ be a triangle. The circles $c_1$ through $A,B$ and
tangent to $g(AC)$, $c_2$ through $B,C$ and tangent to $g(AB)$, and
$c_3$ through $A,C$ and tangent to $g(BC)$ pass through a common
point.
\end{quote}
It leads to a theorem of linear type that can't be translated into
constructive type in an obvious way. The circles may be described each
by 3 dependent variables and 3 conditions
\begin{code}\>\+
A:=Point(0,0); B:=Point(1,0); C:=Point(u1,u2);\\[6pt]

c1:=Circle(1,x1,x2,x3);\\
c2:=Circle(1,x4,x5,x6);\\
c3:=Circle(1,x7,x8,x9);\-\\[6pt]

polys:=\{\>\>
	cl\_tangent(c1,pp\_line(A,C)), \+\+\\
	point\_on\_circle(A,c1), \\
	point\_on\_circle(B,c1), \\
	cl\_tangent(c2,pp\_line(A,B)), \\
	point\_on\_circle(B,c2), \\
	point\_on\_circle(C,c2), \\
	cl\_tangent(c3,pp\_line(B,C)), \\
	point\_on\_circle(A,c3), \\
	point\_on\_circle(C,c3)\};
\end{code}
that are linear in the dependent variables. Hence the coordinates of
the circles and the intersection point of two of them may be computed
and checked for incidence with the third circle:
\begin{code}\>\+
vars:=\{x1,x2,x3,x4,x5,x6,x7,x8,x9\};\\
sol:=solve(polys,vars);\\[6pt]

P:=other\_cc\_point(C,sub(sol,c1),sub(sol,c2));\\
con:=point\_on\_circle(P,sub(sol,c3));
\end{code}
Again $con$ simplifies to zero thus proving the theorem.
\medskip

Even some theorems involving nonlinear objects as circles may be
translated into theorems of constructive type using a rational
parametrization of the non linear object. For a circle with radius $r$
and center $M=(m_1,m_2)$ we may use the rational parametrization
\[\{(\frac{1-u^2}{1+u^2}r+m_1,\frac{2u}{1+u^2}r+m_2)\ |\ u\in {\bf
C}\}.\] 
This way we can prove
\medskip

\noindent 11) {\em Simson's theorem} (\cite[p. 261]{Chou:84},
\cite[thm. 2.51]{Coxeter:67}):
\begin{quote}
Let $P$ be a point on the circle circumscribed to the triangle
$\Delta\,ABC$ and $X,Y,Z$ the pedal points of the perpendiculars from
$P$ onto the lines passing through pairs of vertices of the triangle.
These points are collinear.
\end{quote}

Take the center $M$ of the circumscribed circle as the origin and $r$
as its radius. The proof of the problem may be mechanized in the
following way:
\begin{code}\>\+
        M:=Point(0,0);\\
        A:=choose\_pc(M,r,u1);\\
        B:=choose\_pc(M,r,u2);\\
        C:=choose\_pc(M,r,u3);\\
        P:=choose\_pc(M,r,u4);\\
        X:=pedalpoint(P,pp\_line(A,B));\\
        Y:=pedalpoint(P,pp\_line(B,C));\\
        Z:=pedalpoint(P,pp\_line(A,C));\\[8pt]
 
        con:=collinear(X,Y,Z);
\end{code}
Since $con$ simplifies to zero this proves the theorem.

 
\subsection{Theorems of equational type}

An ``almost'' constructive proof of Simson's theorem may be obtained
in the following way:
\begin{code}\>\+
	A:=Point(0,0); B:=Point(u1,u2);\\
	C:=Point(u3,u4); P:=Point(u5,x1);\\
        X:=pedalpoint(P,pp\_line(A,B));\\
        Y:=pedalpoint(P,pp\_line(B,C));\\
        Z:=pedalpoint(P,pp\_line(A,C));\\[6pt]

	poly:=p4\_circle(A,B,C,P);\\[6pt]
 
        con:=collinear(X,Y,Z);
\end{code}
There is a single dependent variable bound by the quadratic condition
$poly$ that the given points are on a common circle. $con$ is a
rational expression with numerator equal to
\formel{poly\cdot {\rm collinear}(A,B,C)^2. }
Since the second factor may be considered as degeneracy condition this
also proves Simson's theorem. The factors of the denominator
\formel{{\rm den}(con)={\rm sqrdist}(A,B)\cdot {\rm sqrdist}(A,C)\cdot
{\rm sqrdist}(B,C) }
are exactly the non degeneracy conditions collected during the
computation. They may be printed with {\tt print\_ndg()}.
\medskip

One may also substitute the rational coordinate construction of
$X,Y,Z$ through {\tt pedalpoint} with additional dependent variables
and polynomial conditions:
\begin{code}\>\+
M:=Point(0,0); A:=Point(0,1); \\
B:=Point(u1,x1); C:=Point(u2,x2); P:=Point(u3,x3);\\
X:=varpoint(A,B,x4);\\ Y:=varpoint(B,C,x5);\\ Z:=varpoint(A,C,x6);
\end{code}
The polynomial conditions
\begin{code}
polys:=\{\>\> sqrdist(M,B)-1, sqrdist(M,C)-1, sqrdist(M,P)-1,\+\+\\
	orthogonal(pp\_line(A,B),pp\_line(P,X)),\\
	orthogonal(pp\_line(A,C),pp\_line(P,Z)),\\
	orthogonal(pp\_line(B,C),pp\_line(P,Y))\};
\end{code}
contain three quadratic polynomials in $x_1,x_2,x_3$ and three
polynomials linear in $x_4,x_5,x_6$. The quadratic polynomials
correspond to different points on the circle with given
$x$-coordinate.  The best variable order eliminates linear variables
first. Thus the following computations prove the theorem
\begin{code}
con:=collinear(X,Y,Z);\\[8pt]

vars:=\{x4,x5,x6,x1,x2,x3\};\\
setring(vars,\{\},lex);\\
setideal(polys,polys);\\
con mod gbasis polys;
\end{code}
since the conclusion polynomial reduces to zero.
\medskip

\noindent 12) The Butterfly Theorem (\cite[p. 269]{Chou:84},
\cite[thm. 2.81]{Coxeter:67}) :
\begin{quote}
Let $A,B,C,D$ be four points on a circle with center $O$, $P$ the
intersection point of $AC$ and $BD$ and $F$ resp. $G$ the intersection
point of the line through $P$ perpendicular to $OP$ with $AB$
resp. $CD$. Then $P$ is the midpoint of $FG$.
\end{quote}
Taking $P$ as the origin and the lines $g(FG)$ and $g(OP)$ as axes we
get the following coordinatization:
\begin{code}\>\+
P:=Point(0,0); O:=Point(u1,0);\\
A:=Point(u2,u3); B:=Point(u4,x1);\\
C:=Point(x2,x3); D:=Point(x4,x5); \\
F:=Point(0,x6); G:=Point(0,x7);\-\\[6pt]

polys:=\{\>\> sqrdist(O,B)-sqrdist(O,A),\+\+\\
	sqrdist(O,C)-sqrdist(O,A), \\
	sqrdist(O,D)-sqrdist(O,A),\\
	point\_on\_line(P,pp\_line(A,C)),\\
	point\_on\_line(P,pp\_line(B,D)),\\
	point\_on\_line(F,pp\_line(A,D)),\\
	point\_on\_line(G,pp\_line(B,C))\};\-\-\\[6pt]

con:=num sqrdist(P,midpoint(F,G));
\end{code}
Note that the formulation of the theorem includes $A\neq C$ and $B\neq
D$. Hence the conclusion may (and will) fail on some of the components
of $Z(polys)$. This can be avoided supplying appropriate constraints
to the \gr\ factorizer:
\begin{code}\>\+
vars:=\{x6,x7,x3,x5,x1,x2,x4\};\\
setring(vars,\{\},lex);\\[6pt]

sol:=groebfactor(polys,\{sqrdist(A,C),sqrdist(B,D)\});\\[6pt]

for each u in sol collect con mod u;
\end{code}
$sol$ contains a single solution that reduces the conclusion $con$ to
zero. Hence the \gr\ factorizer could split the components and remove
the auxiliary ones.



Note that there is also a constructive proof of the Butterfly theorem,
see {\tt geometry.tst}.
\medskip


\noindent 13) Let's prove another property of Feuerbach's circle
(\cite[thm. 5.61]{Coxeter:67}): 
\begin{quote}
For an arbitrary triangle $\Delta\,ABC$ Feuerbach's circle is tangent
to its in- and excircles (tangent circles for short).
\end{quote}
Take the same coordinates as in example 5 and construct the
coordinates of the center $N$ of Feuerbach's circle $c_1$ as in
example 4:
\begin{code}\>\+
A:=Point(0,0); B:=Point(2,0); C:=Point(u1,u2);\\
 
M:=intersection\_point(mp(A,B),mp(B,C));\\
H:=intersection\_point(altitude(A,B,C),altitude(B,C,A));\\
N:=midpoint(M,H);\\[6pt]	

c1:=c1\_circle(N,sqrdist(N,midpoint(A,B)));
\end{code}
The coordinates of the center {\tt P:=Point(x1,x2)} of one of the
tangent circles are bound by the conditions
\begin{code}
polys:=\{point\_on\_bisector(P,A,B,C), point\_on\_bisector(P,B,C,A)\};
\end{code}
Due to the choice of the coordinates $x_2$ is the radius of this
circle.  Hence the conclusion may be expressed as
\begin{code}\>
con:=cc\_tangent(c1\_circle(P,x2\^{}2),c1);
\end{code}
The polynomial conditions $polys$ have four generic solutions, the
centers of the four tangent circles, as derived in example 5. Since
\begin{code}\>\+
vars:=\{x1,x2\};\\
setring(vars,\{\},lex);\\
setideal(polys,polys);\\
num con mod gbasis polys;
\end{code}
yields zero this proves that all four circles are tangent to
Feuerbach's circle. \cite[ch.5,\S 6]{Coxeter:67} points out that
Feuerbach's circle of $\Delta\,ABC$ coincides with Feuerbach's circle
of each of the triangles $\Delta\,ABH,\, \Delta\,ACH$ and
$\Delta\,BCH$. Hence there are another 12 circles tangent to
$c_1$. This may be proved




Note that the proof in \cite{Coxeter:67} uses inversion geometry. The
author doesn't know about a really ``elementary'' proof of this
theorem.

\section{Exercises}

\begin{itemize}
\item[1.] (\cite[p. 267]{Chou:84}) Let $ABCD$ be a square and $P$ a
point on the line parallel to $BD$ through $C$ such that
$l(BD)=l(BP)$, where $l(BD)$ denotes the distance between $B$ and
$D$. Let $Q$ be the intersection point of $BF$ and $CD$. Show that
$l(DP)=l(DQ)$.

\item[2.] The altitudes' pedal points theorem: Let $P,Q,R$ be the
altitudes' pedal points in the triangle $\Delta\,ABC$. Show that the
altitude through $Q$ bisects $\angle\, PQR$.

\item[3.] Let $\Delta\,ABC$ be an arbitrary triangle. Consider the
three altitude pedal points and the pedal points of the perpendiculars
from these points onto the the opposite sides of the triangle. Show
that these 6 points are on a common circle, the {\em Taylor circle}.  

\item[4.] Prove the formula \[F(\Delta\,ABC) = \frac{a\,b\,c}{4\,R},\]
for the area of the triangle $\Delta\,ABC$, if $a,b,c$ are the lengths
of its sides and $R$ the radius of its circumscribed circle.

\item[5.] (\cite[p. 283]{Chou:84}) Let $k$ be a circle, $A$ the contact
point of the tangent line from a point $B$ to $k$, $M$ the midpoint of
$AB$ and $D$ a point on $k$. Let $C$ be the second intersection point
of $DM$ with $k$, $E$ the second intersection point of $BD$ with $k$
and $F$ the second intersection point of $BC$ with $k$. Show that $EF$
is parallel to $AB$.

\item[6.] (35th IMO 1995, Toronto, problem 1, \cite{IMO}) Let
$A,B,C,D$ be four distinct points on a line, in that order. The
circles with diameters $AC$ and $BD$ intersect at the points $X$ and
$Y$. The line $XY$ meets $BC$ at the point $Z$. Let $P$ be a point on
the line $XY$ different from $Z$. The line $CP$ intersects the circle
with diameter $AC$ at the points $C$ and $M$, and the line $BP$
intersects the circle with diameter $BD$ at the points $B$ and
$N$. Prove that the lines $AM, DN$ and $XY$ are concurrent.

\item[7.] (34th IMO 1994, Hong Kong, problem 2, \cite{IMO}) $ABC$ is
an isosceles triangle with $AB = AC$. Suppose that
\begin{enumerate}
\item[(i)] $M$ is the midpoint of $BC$ and $O$ is the point on the
line $AM$ such that $OB$ is perpendicular to $AB$;
\item[(ii)] $Q$ is an arbitrary point on the segment $BC$ different
from $B$ and $C$;
\item[(iii)] $E$ lies on the line $AB$ and $F$ lies on the line $AC$
such that $E, Q$ and $F$ are distinct and collinear.
\end{enumerate}
\noindent
Prove that $OQ$ is perpendicular to $EF$ if and only if $QE = QF$.

\item[8.] (4th IMO 1959, Czechia, problem 6, \cite{Morozova:68}) Show
that the distance $d$ between the centers of the inscribed and the
circumscribed circles of a triangle $\Delta\,ABC$ satisfies
$d^2=r^2-2r\rho$, where $r$ is the radius of the circumscribed circle
and $\rho$ the radius of the inscribed circle.

\item[9.] (1th IMO 1959, Roumania, problem 5, \cite{Morozova:68}) Let
$M$ be a point on AB, $AMCD$ and $MBEF$ squares to the same side of
$g(AB)$ and $N$ the intersection point of their circumscribed circles,
different from $M$.
\begin{enumerate}
\item[(i)] Show that $g(AF)$ and $g(BC)$ intersect at $N$.
\item[(ii)] Show that all lines $g(MN)$ for various $M$ meet at a
common point. 
\end{enumerate}

\end{itemize}


\bibliographystyle{plain} \bibliography{geometry}

\end{document}


REDUCE Historical
REDUCE Sourceforge Project | Historical SVN Repository | GitHub Mirror | SourceHut Mirror | NotABug Mirror | Chisel Mirror | Chisel RSS ]