Triangulation (topology)

In mathematics, triangulation describes the replacement of topological spaces by piecewise linear spaces, i.e. the choice of a homeomorphism in a suitable simplicial complex. Spaces being homeomorphic to a simplicial complex are called triangulable. Triangulation has various uses in different branches of mathematics, for instance in algebraic topology, in complex analysis or in modeling.

A triangulated torus
Another triangulation of the torus
A triangulated dolphin shape

Motivation edit

On the one hand, it is sometimes useful to forget about superfluous information of topological spaces: The replacement of the original spaces with simplicial complexes may help to recognize crucial properties and to gain a better understanding of the considered object.

On the other hand, simplicial complexes are objects of combinatorial character and therefore one can assign them quantities rising from their combinatorial pattern, for instance, the Euler characteristic. Triangulation allows now to assign such quantities to topological spaces.

Investigations concerning the existence and uniqueness of triangulations established a new branch in topology, namely the piecewise-linear-topology (short PL-topology). Its main purpose is topological properties of simplicial complexes and its generalization, cell-complexes.

Simplicial complexes edit

Abstract simplicial complexes edit

An abstract simplicial complex above a set   is a system   of non-empty subsets such that:

  •   for each  ;
  • if   and      .

The elements of   are called simplices, the elements of   are called vertices. A simplex with   vertices has dimension   by definition. The dimension of an abstract simplicial complex is defined as  .[1]

Abstract simplicial complexes can be thought of as geometrical objects too. This requires the term of geometric simplex.

 
Geometric simplices in dimension 1, 2 and 3

Geometric simplices edit

Let   be   affinely independent points in  , i.e. the vectors  are linearly independent. The set   is said to be the simplex spanned by  . It has dimension   by definition. The points   are called the vertices of  , the simplices spanned by   of the   vertices are called faces and the boundary   is defined to be the union of its faces.

The  -dimensional standard-simplex is the simplex spanned by the unit vectors  [2]

Geometric simplicial complexes edit

A geometric simplicial complex   is a collection of geometric simplices such that

  • If   is a simplex in  , then all its faces are in  .
  • If   are two distinct simplices in  , their interiors are disjoint.

The union of all the simplices in   gives the set of points of  , denoted   This set   is endowed with a topology by choosing the closed sets to be   is closed for all  . Note that, that in general, this topology is not the same as the subspace topology that   inherits from  . The topologies do coincide in the case that each point in the complex lies only in finitely many simplices.[2]

Each geometric complex can be associated with an abstract complex by choosing as a ground set   the set of vertices that appear in any simplex of   and as system of subsets the subsets of   which correspond to vertex sets of simplices in  .

A natural question is if vice versa, any abstract simplicial complex corresponds to a geometric complex. In general, the geometric construction as mentioned here is not flexible enough: consider for instance an abstract simplicial complex of infinite dimension. However, the following more abstract construction provides a topological space for any kind of abstract simplicial complex:

Let   be an abstract simplicial complex above a set  . Choose a union of simplices  , but each in   of dimension sufficiently large, such that the geometric simplex   is of dimension   if the abstract geometric simplex   has dimension  . If  ,  can be identified with a face of   and the resulting topological space is the gluing   Effectuating the gluing for each inclusion, one ends up with the desired topological space.

 
A 2-dimensional geometric simplicial complex with vertex V, link(V), and star(V) are highlighted in red and pink.

As in the previous construction, by the topology induced by gluing, the closed sets in this space are the subsets that are closed in the subspace topology of every simplex   in the complex.

The simplicial complex   which consists of all simplices   of dimension   is called the  -th skeleton of  .

A natural neighbourhood of a vertex   in a simplicial complex   is considered to be given by the star   of a simplex, whose boundary is the link  .

Simplicial maps edit

The maps considered in this category are simplicial maps: Let  ,   be abstract simplicial complexes above sets  ,  . A simplicial map is a function   which maps each simplex in   onto a simplex in  . By affine-linear extension on the simplices,   induces a map between the geometric realizations of the complexes.[2]

Examples edit

  • Let   and let  . The associated geometric complex is a star with center  .
  • Let   and let  . Its geometric realization   is a tetrahedron.
  • Let   as above and let  . The geometric simplicial complex is the boundary of a tetrahedron  .

Definition edit

A triangulation of a topological space   is a homeomorphism   where   is a simplicial complex. Topological spaces do not necessarily admit a triangulation and if they do, it is not necessarily unique.

Examples edit

  • Simplicial complexes can be triangulated by identity.
  • Let   be as in the examples seen above. The closed unit ball   is homeomorphic to a tetrahedron so it admits a triangulation, namely the homeomorphism  . Restricting   to   yields a homeomorphism  .
 
The 2-dimensional sphere and a triangulation
  • The torus   admits a triangulation. To see this, consider the torus as a square where the parallel faces are glued together. This square can be triangulated as shown below:
     
    A two dimensional torus, represented as the gluing of a square via the map g, identifying its opposite sites
  • The projective plane   admits a triangulation (see CW-complexes)
  • One can show that differentiable manifolds admit triangulations.[3]

Invariants edit

Triangulations of spaces allow assigning combinatorial invariants rising from their dedicated simplicial complexes to spaces. These are characteristics that equal for complexes that are isomorphic via a simplicial map and thus have the same combinatorial pattern.

This data might be useful to classify topological spaces up to homeomorphism but only given that the characteristics are also topological invariants, meaning, they do not depend on the chosen triangulation. For the data listed here, this is the case.[4] For details and the link to singular homology, see topological invariance.

Homology edit

Via triangulation, one can assign a chain complex to topological spaces that arise from its simplicial complex and compute its simplicial homology. Compact spaces always admit finite triangulations and therefore their homology groups are finitely generated and only finitely many of them do not vanish. Other data as Betti-numbers or Euler characteristic can be derived from homology.

Betti-numbers and Euler-characteristics edit

Let   be a finite simplicial complex. The  -th Betti-number   is defined to be the rank of the  -th simplicial homology group of the spaces. These numbers encode geometric properties of the spaces: The Betti-number   for instance represents the number of connected components. For a triangulated, closed orientable surfaces  ,   holds where   denotes the genus of the surface: Therefore its first Betti-number represents the doubled number of handles of the surface.[5]

With the comments above, for compact spaces all Betti-numbers are finite and almost all are zero. Therefore, one can form their alternating sum

 

which is called the Euler charakteristic of the complex, a catchy topological invariant.

Topological invariance edit

To use these invariants for the classification of topological spaces up to homeomorphism one needs invariance of the characteristics regarding homeomorphism.

A famous approach to the question was at the beginning of the 20th century the attempt to show that any two triangulations of the same topological space admit a common subdivision. This assumption is known as Hauptvermutung ( German: Main assumption). Let   be a simplicial complex. A complex   is said to be a subdivision of   iff:

  • every simplex of   is contained in a simplex of   and
  • every simplex of   is a finite union of simplices in   .[2]

Those conditions ensure that subdivisions does not change the simplicial complex as a set or as a topological space. A map   between simplicial complexes is said to be piecewise linear if there is a refinement   of   such that   is piecewise linear on each simplex of  . Two complexes that correspond to another via piecewise linear bijection are said to be combinatorial isomorphic. In particular, two complexes that have a common refinement are combinatorially equivalent. Homology groups are invariant to combinatorial equivalence and therefore the Hauptvermutung would give the topological invariance of simplicial homology groups. In 1918, Alexander introduced the concept of singular homology. Henceforth, most of the invariants arising from triangulation were replaced by invariants arising from singular homology. For those new invariants, it can be shown that they were invariant regarding homeomorphism and even regarding homotopy equivalence.[6] Furthermore it was shown that singular and simplicial homology groups coincide.[6] This workaround has shown the invariance of the data to homeomorphism. Hauptvermutung lost in importance but it was initial for a new branch in topology: The piecewise linear topology (short PL-topology).[7]

Hauptvermutung edit

The Hauptvermutung (German for main conjecture) states that two triangulations always admit a common subdivision. Originally, its purpose was to prove invariance of combinatorial invariants regarding homeomorphisms. The assumption that such subdivisions exist in general is intuitive, as subdivision are easy to construct for simple spaces, for instance for low dimensional manifolds. Indeed the assumption was proven for manifolds of dimension   and for differentiable manifolds but it was disproved in general:[8] An important tool to show that triangulations do not admit a common subdivision. i. e their underlying complexes are not combinatorially isomorphic is the combinatorial invariant of Reidemeister torsion.

Reidemeister-torsion edit

To disprove the Hauptvermutung it is helpful to use combinatorial invariants which are not topological invariants. A famous example is Reidemeister-torsion. It can be assigned to a tuple   of CW-complexes: If   this characteristic will be a topological invariant but if   in general not. An approach to Hauptvermutung was to find homeomorphic spaces with different values of Reidemeister-torsion. This invariant was used initially to classify lens-spaces and first counterexamples to the Hauptvermutung were built based on lens-spaces:[8]

Classification of lens-spaces edit

In its original formulation, lens spaces are 3-manifolds, constructed as quotient spaces of the 3-sphere: Let   be natural numbers, such that   are coprime. The lens space   is defined to be the orbit space of the free group action

 
 .

For different tuples  , lens spaces will be homotopy-equivalent but not homeomorphic. Therefore they can't be distinguished with the help of classical invariants as the fundamental group but by the use of Reidemeister-torsion.

Two lens spaces   are homeomorphic, if and only if  .[9] This is the case iff two lens spaces are simple-homotopy-equivalent. The fact can be used to construct counterexamples for the Hauptvermutung as follows. Suppose there are spaces   derived from non-homeomorphic lens spaces  having different Reidemeister torsion. Suppose further that the modification into   does not affect Reidemeister torsion but such that after modification   and   are homeomorphic. The resulting spaces will disprove the Hauptvermutung.

Existence of triangulation edit

Besides the question of concrete triangulations for computational issues, there are statements about spaces that are easier to prove given that they are simplicial complexes. Especially manifolds are of interest. Topological manifolds of dimension   are always triangulable[10][11][1] but there are non-triangulable manifolds for dimension  , for   arbitrary but greater than three.[12][13] Further, differentiable manifolds always admit triangulations.[3]

Piecewise linear structures edit

Manifolds are an important class of spaces. It is natural to require them not only to be triangulable but moreover to admit a piecewise linear atlas, a PL-structure:

Let   be a simplicial complex such that every point admits an open neighborhood   such that there is a triangulation of   and a piecewise linear homeomorphism  . Then   is said to be a piecewise linear (PL) manifold of dimension   and the triangulation together with the PL-atlas is said to be a PL-structure on  .

An important lemma is the following:

Let   be a topological space. It is equivalent

  1.   is an  -dimensional manifold and admits a PL-structure.
  2. There is a triangulation of   such that the link of each vertex is an   sphere.
  3. For each triangulation of   the link of each vertex is an   sphere.

The equivalence of the second and the third statement is because that the link of a vertex is independent of the chosen triangulation up to combinatorial isomorphism.[14] One can show that differentiable manifolds admit a PL-structure as well as manifolds of dimension  .[15] Counterexamples for the triangulation conjecture are counterexamples for the conjecture of the existence of PL-structure of course.

Moreover, there are examples for triangulated spaces which do not admit a PL-structure. Consider an  -dimensional PL-homology-sphere  . The double suspension   is a topological  -sphere. Choosing a triangulation   obtained via the suspension operation on triangulations the resulting simplicial complex is not a PL-manifold, because there is a vertex   such that   is not a   sphere.[16]

A question arising with the definition is if PL-structures are always unique: Given two PL-structures for the same space  , is there a there a homeomorphism   which is piecewise linear with respect to both PL-structures? The assumption is similar to the Hauptvermutung and indeed there are spaces which have different PL-structures which are not equivalent. Triangulation of PL-equivalent spaces can be transformed into one another via Pachner moves:

Pachner Moves edit

 
One Pachner-move replaces two tetrahedra by three tetrahedra

Pachner moves are a way to manipulate triangulations: Let   be a simplicial complex. For two simplices   the Join

  are the points lying on straights between points in   and in  . Choose   such that   for any   lying not in  . A new complex  , can be obtained by replacing   by  . This replacement is called a Pachner move. The theorem of Pachner states that whenever two triangulated manifolds are PL-equivalent, there is a series of Pachner moves transforming both into another.[17]

Cellular complexes edit

 
The real projective plane as a simplicial complex and as CW-complex. As CW-complex it can be obtained by gluing first   and   to get the 1-sphere and then attaching the disc   by the map  .

A similar but more flexible construction than simplicial complexes is the one of cellular complexes (or CW-complexes). Its construction is as follows:

An  -cell is the closed  -dimensional unit-ball  , an open  -cell is its inner  . Let   be a topological space, let   be a continuous map. The gluing   is said to be obtained by gluing on an  -cell.

A cell complex is a union   of topological spaces such that

  •   is a discrete set
  • each   is obtained from   by gluing on a family of  -cells.

Each simplicial complex is a CW-complex, the inverse is not true. The construction of CW-complexes can be used to define cellular homology and one can show that cellular homology and simplicial homology coincide.[18] For computational issues, it is sometimes easier to assume spaces to be CW-complexes and determine their homology via cellular decomposition, an example is the projective plane  : Its construction as a CW-complex needs three cells, whereas its simplicial complex consists of 54 simplices.

Other applications edit

Classification of manifolds edit

By triangulating 1-dimensional manifolds, one can show that they are always homeomorphic to disjoint copies of the real line and the unit sphere  . Moreover, surfaces, i.e. 2-manifolds, can be classified completely: Let   be a compact surface.

  • If   is orientable, it is homeomorphic to a 2-sphere with   tori of dimension   attached, for some  .
  • If   is not orientable, it is homeomorphic to a Klein Bottle with   tori of dimension   attached, for some  .

To prove this theorem one constructs a fundamental polygon of the surface: This can be done by using the simplicial structure obtained by the triangulation.[19]

Maps on simplicial complexes edit

Giving spaces the structure of a simplicial structure might help to understand maps defined on the spaces. The maps can often be assumed to be simplicial maps via the simplicial approximation theorem:

Simplicial approximation edit

Let  ,   be abstract simplicial complexes above sets  ,  . A simplicial map is a function   which maps each simplex in   onto a simplex in  . By affin-linear extension on the simplices,   induces a map between the geometric realizations of the complexes. Each point in a geometric complex lies in the inner of exactly one simplex, its support. Consider now a continuous map  . A simplicial map   is said to be a simplicial approximation of   if and only if each   is mapped by   onto the support of   in  . If such an approximation exists, one can construct a homotopy   transforming   into   by defining it on each simplex; there it always exists, because simplices are contractible.

The simplicial approximation theorem guarantees for every continuous function   the existence of a simplicial approximation at least after refinement of  , for instance by replacing   by its iterated barycentric subdivision.[2] The theorem plays an important role for certain statements in algebraic topology in order to reduce the behavior of continuous maps on those of simplicial maps, for instance in Lefschetz's fixed-point theorem.

Lefschetz's fixed-point theorem edit

The Lefschetz number is a useful tool to find out whether a continuous function admits fixed-points. This data is computed as follows: Suppose that   and   are topological spaces that admit finite triangulations. A continuous map   induces homomorphisms   between its simplicial homology groups with coefficients in a field  . These are linear maps between  -vector spaces, so their trace   can be determined and their alternating sum

 

is called the Lefschetz number of  . If  , this number is the Euler characteristic of  . The fixpoint theorem states that whenever  ,   has a fixed-point. In the proof this is first shown only for simplicial maps and then generalized for any continuous functions via the approximation theorem. Brouwer's fixpoint theorem treats the case where   is an endomorphism of the unit-ball. For   all its homology groups   vanishes, and   is always the identity, so  , so   has a fixpoint.[20]

Formula of Riemann-Hurwitz edit

The Riemann-Hurwitz formula allows to determine the genus of a compact, connected Riemann surface   without using explicit triangulation. The proof needs the existence of triangulations for surfaces in an abstract sense: Let   be a non-constant holomorphic function on a surface with known genus. The relation between the genus   of the surfaces   and   is

 

where   denotes the degree of the map. The sum is well defined as it counts only the ramifying points of the function.

The background of this formula is that holomorphic functions on Riemann surfaces are ramified coverings. The formula can be found by examining the image of the simplicial structure near to ramifiying points.[21]

Citations edit

  1. ^ a b John M. Lee (2000), Springer Verlag (ed.), Introduction to Topological manifolds (in German), New York/Berlin/Heidelberg: Springer Verlag, p. 92, ISBN 0-387-98759-2
  2. ^ a b c d e James R. Munkres (1984), Elements of algebraic topology (in German), vol. 1984, Menlo Park, California: Addison Wesley, p. 83, ISBN 0-201-04586-9
  3. ^ a b J. H. C. Whitehead (1940), "On C1-Complexes", Annals of Mathematics (in German), vol. 41, no. 4, pp. 809–824, doi:10.2307/1968861, ISSN 0003-486X, JSTOR 1968861
  4. ^ J. W. Alexander (1926), "Combinatorial Analysis Situs", Transactions of the American Mathematical Society (in German), vol. 28, no. 2, pp. 301–329, doi:10.1090/S0002-9947-1926-1501346-5, ISSN 0002-9947, JSTOR 1989117
  5. ^ R. Stöcker, H. Zieschang (1994), Algebraische Topologie (in German) (2. überarbeitete ed.), Stuttgart: B.G.Teubner, p. 270, ISBN 3-519-12226-X
  6. ^ a b Allen Hatcher (2006), Algebraic Topologie (in German), Cambridge/New York/Melbourne: Cambridge University Press, p. 110, ISBN 0-521-79160--X
  7. ^ A.A.Ranicki. "One the Hauptvermutung" (PDF). The Hauptvermutung Book.
  8. ^ a b John Milnor (1961), "Two Complexes Which are Homeomorphic But Combinatorially Distinct", The Annals of Mathematics (in German), vol. 74, no. 3, p. 575, doi:10.2307/1970299, ISSN 0003-486X, JSTOR 1970299
  9. ^ Marshall M. Cohen (1973), "A Course in Simple-Homotopy Theory", Graduate Texts in Mathematics, Graduate Texts in Mathematics (in German), vol. 10, doi:10.1007/978-1-4684-9372-6, ISBN 978-0-387-90055-1, ISSN 0072-5285
  10. ^ Edwin Moise (1977), Geometric Topology in Dimensions 2 and 3 (in German), New York: Springer Verlag
  11. ^ Tibor Rado. "Über den Begriff der Riemannschen Fläche" (PDF).
  12. ^ R. C. Kirby, L. C. Siebenmann (1977-12-31), "Annex B. On The Triangulation of Manifolds and the Hauptvermutung", Foundational Essays on Topological Manifolds, Smoothings, and Triangulations. (AM-88) (in German), Princeton University Press, pp. 299–306
  13. ^ "Chapter IV; Casson's Invariant for Oriented Homology 3-spheres", Casson's Invariant for Oriented Homology Three-Spheres (in German), Princeton University Press, pp. 63–79, 1990-12-31
  14. ^ Toenniessen, Fridtjof (2017), Topologie (PDF) (in German), doi:10.1007/978-3-662-54964-3, ISBN 978-3-662-54963-6, retrieved 2022-04-20
  15. ^ Edwin E. Moise (1952), "Affine Structures in 3-Manifolds: V. The Triangulation Theorem and Hauptvermutung", The Annals of Mathematics (in German), vol. 56, no. 1, p. 96, doi:10.2307/1969769, ISSN 0003-486X, JSTOR 1969769
  16. ^ Robert D. Edwards (2006-10-18), "Suspensions of homology spheres", arXiv:math/0610573 (in German), arXiv:math/0610573, Bibcode:2006math.....10573E
  17. ^ W B R Lickorish (1999-11-20), "Simplicial moves on complexes and manifolds", Proceedings of the Kirbyfest (in German), Mathematical Sciences Publishers, arXiv:math/9911256, doi:10.2140/gtm.1999.2.299, S2CID 9765634
  18. ^ Toenniessen, Fridtjof (2017), Topologie (PDF) (in German), p. 315, doi:10.1007/978-3-662-54964-3, ISBN 978-3-662-54963-6, retrieved 2022-04-20
  19. ^ Seifert, H. (Herbert), 1907-1996. (2003), Lehrbuch der Topologie (in German), AMS Chelsea Pub., ISBN 0-8218-3595-5{{citation}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  20. ^ Bredon, Glen E. (1993), Springer Verlag (ed.), Topology and Geometry (in German), Berlin/ Heidelberg/ New York, pp. 254 f, ISBN 3-540-97926-3{{citation}}: CS1 maint: location missing publisher (link)
  21. ^ Otto Forster (1977), "Kompakte Riemannsche Flächen", Heidelberger Taschenbücher (in German), Berlin, Heidelberg: Springer Berlin Heidelberg, pp. 88–154, ISBN 978-3-540-08034-3

See also edit

Literature edit

  • Allen Hatcher: Algebraic Topology, Cambridge University Press, Cambridge/New York/Melbourne 2006, ISBN 0-521-79160-X
  • James R. Munkres: . Band 1984. Addison Wesley, Menlo Park, California 1984, ISBN 0-201-04586-9
  • Marshall M. Cohen: A course in Simple-Homotopy Theory . In: Graduate Texts in Mathematics. 1973, ISSN 0072-5285, doi:10.1007/978-1-4684-9372-6.