Poincaré conjecture

Theorem in geometric topology
  • Geometrization conjecture
  • Zeeman conjecture[1]
GeneralizationsGeneralized Poincaré conjecture
Millennium Prize Problems
  • v
  • t
  • e

In the mathematical field of geometric topology, the Poincaré conjecture (UK: /ˈpwæ̃kær/,[2] US: /ˌpwæ̃kɑːˈr/,[3][4] French: [pwɛ̃kaʁe]) is a theorem about the characterization of the 3-sphere, which is the hypersphere that bounds the unit ball in four-dimensional space.

Originally conjectured by Henri Poincaré in 1904, the theorem concerns spaces that locally look like ordinary three-dimensional space but which are finite in extent. Poincaré hypothesized that if such a space has the additional property that each loop in the space can be continuously tightened to a point, then it is necessarily a three-dimensional sphere. Attempts to resolve the conjecture drove much progress in the field of geometric topology during the 20th century.

The eventual proof built upon Richard S. Hamilton's program of using the Ricci flow to solve the problem. By developing a number of new techniques and results in the theory of Ricci flow, Grigori Perelman was able to modify and complete Hamilton's program. In papers posted to the arXiv repository in 2002 and 2003, Perelman presented his work proving the Poincaré conjecture (and the more powerful geometrization conjecture of William Thurston). Over the next several years, several mathematicians studied his papers and produced detailed formulations of his work.

Hamilton and Perelman's work on the conjecture is widely recognized as a milestone of mathematical research. Hamilton was recognized with the Shaw Prize and the Leroy P. Steele Prize for Seminal Contribution to Research. The journal Science marked Perelman's proof of the Poincaré conjecture as the scientific Breakthrough of the Year in 2006.[5] The Clay Mathematics Institute, having included the Poincaré conjecture in their well-known Millennium Prize Problem list, offered Perelman their prize of US$1 million for the conjecture's resolution.[6] He declined the award, saying that Hamilton's contribution had been equal to his own.[7][8]

Overview

Neither of the two colored loops on this torus can be continuously tightened to a point. A torus is not homeomorphic to a sphere.

The Poincaré conjecture was a mathematical problem in the field of geometric topology. In terms of the vocabulary of that field, it says the following:

Poincaré conjecture.
Every three-dimensional topological manifold which is closed, connected, and has trivial fundamental group is homeomorphic to the three-dimensional sphere.

Familiar shapes, such as the surface of a ball (which is known in mathematics as the two-dimensional sphere) or of a torus, are two-dimensional. The surface of a ball has trivial fundamental group, meaning that any loop drawn on the surface can be continuously deformed to a single point. By contrast, the surface of a torus has nontrivial fundamental group, as there are loops on the surface which cannot be so deformed. Both are topological manifolds which are closed (meaning that they have no boundary and take up a finite region of space) and connected (meaning that they consist of a single piece). Two closed manifolds are said to be homeomorphic when it is possible for the points of one can be reallocated to the other in a continuous way. Because the (non)triviality of the fundamental group is known to be invariant under homeomorphism, it follows that the two-dimensional sphere and torus are not homeomorphic.

The two-dimensional analogue of the Poincaré conjecture says that any two-dimensional topological manifold which is closed and connected but non-homeomorphic to the two-dimensional sphere must possess a loop which cannot be continuously contracted to a point. (This is illustrated by the example of the torus, as above.) This analogue is known to be true via the classification of closed and connected two-dimensional topological manifolds, which was understood in various forms since the 1860s. In higher dimensions, the closed and connected topological manifolds do not have a straightforward classification, precluding an easy resolution of the Poincaré conjecture.

History

Poincaré's question

In the 1800s, Bernhard Riemann and Enrico Betti initiated the study of topological invariants of manifolds.[9][10] They introduced the Betti numbers, which associate to any manifold a list of nonnegative integers. Riemann had showed that a closed connected two-dimensional manifold is fully characterized by its Betti numbers. As part of his 1895 paper Analysis Situs (announced in 1892), Poincaré showed that Riemann's result does not extend to higher dimensions.[11][12][13] To do this he introduced the fundamental group as a novel topological invariant, and was able to exhibit examples of three-dimensional manifolds which have the same Betti numbers but distinct fundamental groups. He posed the question of whether the fundamental group is sufficient to topologically characterize a manifold (of given dimension), although he made no attempt to pursue the answer, saying only that it would "demand lengthy and difficult study."[12][13][14]

The primary purpose of Poincaré's paper was the interpretation of the Betti numbers in terms of his newly-introduced homology groups, along with the Poincaré duality theorem on the symmetry of Betti numbers. Following criticism of the completeness of his arguments, he released a number of subsequent "supplements" to enhance and correct his work. The closing remark of his second supplement, published in 1900, said:[15][13]

In order to avoid making this work too prolonged, I confine myself to stating the following theorem, the proof of which will require further developments:

Each polyhedron which has all its Betti numbers equal to 1 and all its tables Tq orientable is simply connected, i.e., homeomorphic to a hypersphere.

(In a modern language, taking note of the fact that Poincaré is using the terminology of simple-connectedness in an unusual way,[16] this says that a closed connected oriented manifold with the homology of a sphere must be homeomorphic to a sphere.[14]) This modified his negative generalization of Riemann's work in two ways. Firstly, he was now making use of the full homology groups and not only the Betti numbers. Secondly, he narrowed the scope of the problem from asking if an arbitrary manifold is characterized by topological invariants to asking whether the sphere can be so characterized.

However, after publication he found his announced theorem to be incorrect. In his fifth and final supplement, published in 1904, he proved this with the counterexample of the Poincaré homology sphere, which is a closed connected three-dimensional manifold which has the homology of the sphere but whose fundamental group has 120 elements. This example made it clear that homology is not powerful enough to characterize the topology of a manifold. In the closing remarks of the fifth supplement, Poincaré modified his erroneous theorem to use the fundamental group instead of homology:[17][13]

One question remains to be dealt with: is it possible for the fundamental group of V to reduce to the identity without V being simply connected? [...] However, this question would carry us too far away.

In this remark, as in the closing remark of the second supplement, Poincaré used the term "simply connected" in a way which is at odds with modern usage, as well as his own 1895 definition of the term.[12][16] (According to modern usage, Poincaré's question is a tautology, asking if it is possible for a manifold to be simply connected without being simply connected.) However, as can be inferred from context,[18] Poincaré was asking whether the triviality of the fundamental group uniquely characterizes the sphere.[14]

Throughout the work of Riemann, Betti, and Poincaré, the topological notions in question are not defined or used in a way that would be recognized as precise from a modern perspective. Even the key notion of a "manifold" was not used in a consistent way in Poincaré's own work, and there was frequent confusion between the notion of a topological manifold, a PL manifold, and a smooth manifold.[16][19] For this reason, it is not possible to read Poincaré's questions unambiguously. It is only through the formalization and vocabulary of topology as developed by later mathematicians that Poincaré's closing question has been understood as the "Poincaré conjecture" as stated in the preceding section.

However, despite its usual phrasing in the form of a conjecture, proposing that all manifolds of a certain type are homeomorphic to the sphere, Poincaré only posed an open-ended question, without venturing to conjecture one way or the other. Moreover, there is no evidence as to which way he believed his question would be answered.[14]

Solutions

In the 1930s, J. H. C. Whitehead claimed a proof but then retracted it. In the process, he discovered some examples of simply-connected (indeed contractible, i.e. homotopically equivalent to a point) non-compact 3-manifolds not homeomorphic to R 3 {\displaystyle \mathbb {R} ^{3}} , the prototype of which is now called the Whitehead manifold.

In the 1950s and 1960s, other mathematicians attempted proofs of the conjecture only to discover that they contained flaws. Influential mathematicians such as Georges de Rham, R. H. Bing, Wolfgang Haken, Edwin E. Moise, and Christos Papakyriakopoulos attempted to prove the conjecture. In 1958, R. H. Bing proved a weak version of the Poincaré conjecture: if every simple closed curve of a compact 3-manifold is contained in a 3-ball, then the manifold is homeomorphic to the 3-sphere.[20] Bing also described some of the pitfalls in trying to prove the Poincaré conjecture.[21]

Włodzimierz Jakobsche showed in 1978 that, if the Bing–Borsuk conjecture is true in dimension 3, then the Poincaré conjecture must also be true.[22]

Over time, the conjecture gained the reputation of being particularly tricky to tackle. John Milnor commented that sometimes the errors in false proofs can be "rather subtle and difficult to detect."[23] Work on the conjecture improved understanding of 3-manifolds. Experts in the field were often reluctant to announce proofs and tended to view any such announcement with skepticism. The 1980s and 1990s witnessed some well-publicized fallacious proofs (which were not actually published in peer-reviewed form).[24][25]

An exposition of attempts to prove this conjecture can be found in the non-technical book Poincaré's Prize by George Szpiro.[26]

Dimensions

The classification of closed surfaces gives an affirmative answer to the analogous question in two dimensions. For dimensions greater than three, one can pose the Generalized Poincaré conjecture: is a homotopy n-sphere homeomorphic to the n-sphere? A stronger assumption than simply-connectedness is necessary; in dimensions four and higher there are simply-connected, closed manifolds which are not homotopy equivalent to an n-sphere.

Historically, while the conjecture in dimension three seemed plausible, the generalized conjecture was thought to be false. In 1961, Stephen Smale shocked mathematicians by proving the Generalized Poincaré conjecture for dimensions greater than four and extended his techniques to prove the fundamental h-cobordism theorem. In 1982, Michael Freedman proved the Poincaré conjecture in four dimensions. Freedman's work left open the possibility that there is a smooth four-manifold homeomorphic to the four-sphere which is not diffeomorphic to the four-sphere. This so-called smooth Poincaré conjecture, in dimension four, remains open and is thought to be very difficult. Milnor's exotic spheres show that the smooth Poincaré conjecture is false in dimension seven, for example.

These earlier successes in higher dimensions left the case of three dimensions in limbo. The Poincaré conjecture was essentially true in both dimension four and all higher dimensions for substantially different reasons. In dimension three, the conjecture had an uncertain reputation until the geometrization conjecture put it into a framework governing all 3-manifolds. John Morgan wrote:[27]

It is my view that before Thurston's work on hyperbolic 3-manifolds and … the Geometrization conjecture there was no consensus among the experts as to whether the Poincaré conjecture was true or false. After Thurston's work, notwithstanding the fact that it had no direct bearing on the Poincaré conjecture, a consensus developed that the Poincaré conjecture (and the Geometrization conjecture) were true.

Hamilton's program and solution

Several stages of the Ricci flow on a two-dimensional manifold

Hamilton's program was started in his 1982 paper in which he introduced the Ricci flow on a manifold and showed how to use it to prove some special cases of the Poincaré conjecture.[28] In the following years, he extended this work but was unable to prove the conjecture. The actual solution was not found until Grigori Perelman published his papers.

In late 2002 and 2003, Perelman posted three papers on arXiv.[29][30][31] In these papers, he sketched a proof of the Poincaré conjecture and a more general conjecture, Thurston's geometrization conjecture, completing the Ricci flow program outlined earlier by Richard S. Hamilton.

From May to July 2006, several groups presented papers that filled in the details of Perelman's proof of the Poincaré conjecture, as follows:

  • Bruce Kleiner and John W. Lott posted a paper on arXiv in May 2006 which filled in the details of Perelman's proof of the geometrization conjecture, following partial versions which had been publicly available since 2003.[32] Their manuscript was published in the journal "Geometry and Topology" in 2008. A small number of corrections were made in 2011 and 2013; for instance, the first version of their published paper made use of an incorrect version of Hamilton's compactness theorem for Ricci flow.
  • Huai-Dong Cao and Xi-Ping Zhu published a paper in the June 2006 issue of the Asian Journal of Mathematics with an exposition of the complete proof of the Poincaré and geometrization conjectures.[33] The opening paragraph of their paper stated

In this paper, we shall present the Hamilton-Perelman theory of Ricci flow. Based on it, we shall give the first written account of a complete proof of the Poincaré conjecture and the geometrization conjecture of Thurston. While the complete work is an accumulated efforts of many geometric analysts, the major contributors are unquestionably Hamilton and Perelman.

Some observers interpreted Cao and Zhu as taking credit for Perelman's work. They later posted a revised version, with new wording, on arXiv.[34] In addition, a page of their exposition was essentially identical to a page in one of Kleiner and Lott's early publicly available drafts; this was also amended in the revised version, together with an apology by the journal's editorial board.
  • John Morgan and Gang Tian posted a paper on arXiv in July 2006 which gave a detailed proof of just the Poincaré Conjecture (which is somewhat easier than the full geometrization conjecture)[35] and expanded this to a book.[36][37]

All three groups found that the gaps in Perelman's papers were minor and could be filled in using his own techniques.

On August 22, 2006, the ICM awarded Perelman the Fields Medal for his work on the Ricci flow, but Perelman refused the medal.[38][39] John Morgan spoke at the ICM on the Poincaré conjecture on August 24, 2006, declaring that "in 2003, Perelman solved the Poincaré Conjecture."[40]

In December 2006, the journal Science honored the proof of Poincaré conjecture as the Breakthrough of the Year and featured it on its cover.[5]

Ricci flow with surgery

Hamilton's program for proving the Poincaré conjecture involves first putting a Riemannian metric on the unknown simply connected closed 3-manifold. The basic idea is to try to "improve" this metric; for example, if the metric can be improved enough so that it has constant positive curvature, then according to classical results in Riemannian geometry, it must be the 3-sphere. Hamilton prescribed the "Ricci flow equations" for improving the metric;

t g i j = 2 R i j {\displaystyle \partial _{t}g_{ij}=-2R_{ij}}

where g is the metric and R its Ricci curvature, and one hopes that, as the time t increases, the manifold becomes easier to understand. Ricci flow expands the negative curvature part of the manifold and contracts the positive curvature part.

In some cases, Hamilton was able to show that this works; for example, his original breakthrough was to show that if the Riemannian manifold has positive Ricci curvature everywhere, then the above procedure can only be followed for a bounded interval of parameter values, t [ 0 , T ) {\displaystyle t\in [0,T)} with T < {\displaystyle T<\infty } , and more significantly, that there are numbers c t {\displaystyle c_{t}} such that as t T {\displaystyle t\nearrow T} , the Riemannian metrics c t g ( t ) {\displaystyle c_{t}g(t)} smoothly converge to one of constant positive curvature. According to classical Riemannian geometry, the only simply-connected compact manifold which can support a Riemannian metric of constant positive curvature is the sphere. So, in effect, Hamilton showed a special case of the Poincaré conjecture: if a compact simply-connected 3-manifold supports a Riemannian metric of positive Ricci curvature, then it must be diffeomorphic to the 3-sphere.

If, instead, one only has an arbitrary Riemannian metric, the Ricci flow equations must lead to more complicated singularities. Perelman's major achievement was to show that, if one takes a certain perspective, if they appear in finite time, these singularities can only look like shrinking spheres or cylinders. With a quantitative understanding of this phenomenon, he cuts the manifold along the singularities, splitting the manifold into several pieces and then continues with the Ricci flow on each of these pieces. This procedure is known as Ricci flow with surgery.

Perelman provided a separate argument based on curve shortening flow to show that, on a simply-connected compact 3-manifold, any solution of the Ricci flow with surgery becomes extinct in finite time. An alternative argument, based on the min-max theory of minimal surfaces and geometric measure theory, was provided by Tobias Colding and William Minicozzi. Hence, in the simply-connected context, the above finite-time phenomena of Ricci flow with surgery is all that is relevant. In fact, this is even true if the fundamental group is a free product of finite groups and cyclic groups.

This condition on the fundamental group turns out to be necessary and sufficient for finite time extinction. It is equivalent to saying that the prime decomposition of the manifold has no acyclic components and turns out to be equivalent to the condition that all geometric pieces of the manifold have geometries based on the two Thurston geometries S2×R and S3. In the context that one makes no assumption about the fundamental group whatsoever, Perelman made a further technical study of the limit of the manifold for infinitely large times, and in so doing, proved Thurston's geometrization conjecture: at large times, the manifold has a thick-thin decomposition, whose thick piece has a hyperbolic structure, and whose thin piece is a graph manifold. Due to Perelman's and Colding and Minicozzi's results, however, these further results are unnecessary in order to prove the Poincaré conjecture.

Solution

Grigori Perelman

On November 13, 2002, Russian mathematician Grigori Perelman posted the first of a series of three eprints on arXiv outlining a solution of the Poincaré conjecture. Perelman's proof uses a modified version of a Ricci flow program developed by Richard S. Hamilton. In August 2006, Perelman was awarded, but declined, the Fields Medal (worth $15,000 CAD) for his work on the Ricci flow. On March 18, 2010, the Clay Mathematics Institute awarded Perelman the $1 million Millennium Prize in recognition of his proof.[41][42] Perelman rejected that prize as well.[7][43]

Perelman proved the conjecture by deforming the manifold using the Ricci flow (which behaves similarly to the heat equation that describes the diffusion of heat through an object). The Ricci flow usually deforms the manifold towards a rounder shape, except for some cases where it stretches the manifold apart from itself towards what are known as singularities. Perelman and Hamilton then chop the manifold at the singularities (a process called "surgery"), causing the separate pieces to form into ball-like shapes. Major steps in the proof involve showing how manifolds behave when they are deformed by the Ricci flow, examining what sort of singularities develop, determining whether this surgery process can be completed, and establishing that the surgery need not be repeated infinitely many times.

The first step is to deform the manifold using the Ricci flow. The Ricci flow was defined by Richard S. Hamilton as a way to deform manifolds. The formula for the Ricci flow is an imitation of the heat equation, which describes the way heat flows in a solid. Like the heat flow, Ricci flow tends towards uniform behavior. Unlike the heat flow, the Ricci flow could run into singularities and stop functioning. A singularity in a manifold is a place where it is not differentiable: like a corner or a cusp or a pinching. The Ricci flow was only defined for smooth differentiable manifolds. Hamilton used the Ricci flow to prove that some compact manifolds were diffeomorphic to spheres, and he hoped to apply it to prove the Poincaré Conjecture. He needed to understand the singularities.[citation needed]

Hamilton created a list of possible singularities that could form, but he was concerned that some singularities might lead to difficulties. He wanted to cut the manifold at the singularities and paste in caps and then run the Ricci flow again, so he needed to understand the singularities and show that certain kinds of singularities do not occur. Perelman discovered the singularities were all very simple: consider that a cylinder is formed by 'stretching' a circle along a line in another dimension, repeating that process with spheres instead of circles essentially gives the form of the singularities. Perelman proved this using something called the "Reduced Volume", which is closely related to an eigenvalue of a certain elliptic equation.

Sometimes, an otherwise complicated operation reduces to multiplication by a scalar (a number). Such numbers are called eigenvalues of that operation. Eigenvalues are closely related to vibration frequencies and are used in analyzing a famous problem: can you hear the shape of a drum? Essentially, an eigenvalue is like a note being played by the manifold. Perelman proved this note goes up as the manifold is deformed by the Ricci flow. This helped him eliminate some of the more troublesome singularities that had concerned Hamilton, particularly the cigar soliton solution, which looked like a strand sticking out of a manifold with nothing on the other side. In essence, Perelman showed that all the strands that form can be cut and capped and none stick out on one side only.

Completing the proof, Perelman takes any compact, simply connected, three-dimensional manifold without boundary and starts to run the Ricci flow. This deforms the manifold into round pieces with strands running between them. He cuts the strands and continues deforming the manifold until, eventually, he is left with a collection of round three-dimensional spheres. Then, he rebuilds the original manifold by connecting the spheres together with three-dimensional cylinders, morphs them into a round shape, and sees that, despite all the initial confusion, the manifold was, in fact, homeomorphic to a sphere.

One immediate question posed was how one could be sure that infinitely many cuts are not necessary. This was raised due to the cutting potentially progressing forever. Perelman proved this cannot happen by using minimal surfaces on the manifold. A minimal surface is one on which any local deformation increases area; a familiar example is a soap film spanning a bent loop of wire. Hamilton had shown that the area of a minimal surface decreases as the manifold undergoes Ricci flow. Perelman verified what happened to the area of the minimal surface when the manifold was sliced. He proved that, eventually, the area is so small that any cut after the area is that small can only be chopping off three-dimensional spheres and not more complicated pieces. This is described as a battle with a Hydra by Sormani in Szpiro's book cited below. This last part of the proof appeared in Perelman's third and final paper on the subject.

References

  1. ^ Matveev, Sergei (2007). "1.3.4 Zeeman's Collapsing Conjecture". Algorithmic Topology and Classification of 3-Manifolds. Algorithms and Computation in Mathematics. Vol. 9. Springer. pp. 46–58. ISBN 978-3540458999.
  2. ^ "Poincaré, Jules-Henri". Lexico UK English Dictionary. Oxford University Press. Archived from the original on 2022-09-02.
  3. ^ "Poincaré". The American Heritage Dictionary of the English Language (5th ed.). HarperCollins. Retrieved 9 August 2019.
  4. ^ "Poincaré". Merriam-Webster.com Dictionary. Retrieved 9 August 2019.
  5. ^ a b Mackenzie, Dana (2006-12-22). "The Poincaré Conjecture – Proved". Science. 314 (5807): 1848–1849. doi:10.1126/science.314.5807.1848. PMID 17185565. S2CID 121869167.
  6. ^ "Prize for Resolution of the Poincaré Conjecture Awarded to Dr. Grigoriy Perelman" (Press release). Clay Mathematics Institute. March 18, 2010. Archived from the original (PDF) on March 22, 2010. Retrieved November 13, 2015. The Clay Mathematics Institute (CMI) announces today that Dr. Grigoriy Perelman of St. Petersburg, Russia, is the recipient of the Millennium Prize for resolution of the Poincaré conjecture.
  7. ^ a b "Последнее 'нет' доктора Перельмана" [The last "no" Dr. Perelman]. Interfax (in Russian). July 1, 2010. Retrieved 5 April 2016. Google Translated archived link at [1] (archived 2014-04-20)
  8. ^ Ritter, Malcolm (1 July 2010). "Russian mathematician rejects million prize". The Boston Globe.
  9. ^ Riemann, Bernhard (1851). Grundlagen für eine allgemeine Theorie der Functionen (Thesis). University of Göttingen. English translation: Riemann, Bernhard (2004). "Foundations for a general theory of functions of a complex variable". Collected Papers: Bernhard Riemann. Translated by Baker, Roger; Christenson, Charles; Orde, Henry. Heber City, UT: Kendrick Press. pp. 1–41. ISBN 0-9740427-2-2. MR 2121437. Zbl 1101.01013.
  10. ^ Betti, Enrico (1870). "Sopra gli spazi di un numero qualunque di dimensioni". Annali di Matematica Pura ed Applicata. 4: 140–158. doi:10.1007/BF02420029. JFM 03.0301.01.
  11. ^ Poincaré, H. (1892). "Sur l'Analysis situs". Comptes Rendus des Séances de l'Académie des Sciences. JFM 24.0506.02.
  12. ^ a b c Poincaré, H. (1895). "Analysis situs". Journal de l'École Polytechnique. 2e Série. 1: 1–121. JFM 26.0541.07.
  13. ^ a b c d Poincaré, Henri (2010). Papers on Topology: Analysis Situs and Its Five Supplements. History of Mathematics. Vol. 37. Translated by Stillwell, John. American Mathematical Society and London Mathematical Society. doi:10.1090/hmath/037. ISBN 978-0-8218-5234-7. MR 2723194. Zbl 1204.55002.
  14. ^ a b c d Gray, Jeremy (2013). Henri Poincaré: A Scientific Biography. Princeton, NJ: Princeton University Press. ISBN 978-0-691-15271-4. JSTOR j.ctt1r2fwt. MR 2986502. Zbl 1263.01002.
  15. ^ Poincaré, H. (1900). "Second complément à l'analysis situs". Proceedings of the London Mathematical Society. 32 (1): 277–308. doi:10.1112/plms/s1-32.1.277. JFM 31.0477.10. MR 1576227.
  16. ^ a b c cf. Stillwell's commentary in Poincaré (2010)
  17. ^ Poincaré, H. (1904). "Cinquième complément à l'analysis situs". Rendiconti del Circolo Matematico di Palermo. 18: 45–110. doi:10.1007/bf03014091. JFM 35.0504.13.
  18. ^ The opening paragraphs of Poincaré (1904) refer to "simply connected in the true sense of the word" as the condition of being homeomorphic to a sphere.
  19. ^ Dieudonné, Jean (1989). A History of Algebraic and Differential Topology, 1900–1960. Boston, MA: Birkhäuser Boston, Inc. doi:10.1007/978-0-8176-4907-4. ISBN 0-8176-3388-X. MR 0995842. Zbl 0673.55002.
  20. ^ Bing, R. H. (1958). "Necessary and sufficient conditions that a 3-manifold be S3". Annals of Mathematics. Second Series. 68 (1): 17–37. doi:10.2307/1970041. JSTOR 1970041.
  21. ^ Bing, R. H. (1964). "Some aspects of the topology of 3-manifolds related to the Poincaré conjecture". Lectures on Modern Mathematics. Vol. II. New York: Wiley. pp. 93–128.
  22. ^ M., Halverson, Denise; Dušan, Repovš (23 December 2008). "The Bing–Borsuk and the Busemann conjectures". Mathematical Communications. 13 (2). arXiv:0811.0886.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  23. ^ Milnor, John (2004). "The Poincaré Conjecture 99 Years Later: A Progress Report" (PDF). Retrieved 2007-05-05.
  24. ^ Taubes, Gary (July 1987). "What happens when hubris meets nemesis". Discover. 8: 66–77.
  25. ^ Matthews, Robert (9 April 2002). "$1 million mathematical mystery "solved"". NewScientist.com. Retrieved 2007-05-05.
  26. ^ Szpiro, George (2008). Poincaré's Prize: The Hundred-Year Quest to Solve One of Math's Greatest Puzzles. Plume. ISBN 978-0-452-28964-2.
  27. ^ Morgan, John W., Recent progress on the Poincaré conjecture and the classification of 3-manifolds. Bull. Amer. Math. Soc. (N.S.) 42 (2005), no. 1, 57–78
  28. ^ Hamilton, Richard (1982). "Three-manifolds with positive Ricci curvature". Journal of Differential Geometry. 17 (2): 255–306. doi:10.4310/jdg/1214436922. MR 0664497. Zbl 0504.53034. Reprinted in: Cao, H. D.; Chow, B.; Chu, S. C.; Yau, S.-T., eds. (2003). Collected Papers on Ricci Flow. Series in Geometry and Topology. Vol. 37. Somerville, MA: International Press. pp. 119–162. ISBN 1-57146-110-8.
  29. ^ Perelman, Grigori (2002). "The entropy formula for the Ricci flow and its geometric applications". arXiv:math.DG/0211159.
  30. ^ Perelman, Grigori (2003). "Ricci flow with surgery on three-manifolds". arXiv:math.DG/0303109.
  31. ^ Perelman, Grigori (2003). "Finite extinction time for the solutions to the Ricci flow on certain three-manifolds". arXiv:math.DG/0307245.
  32. ^ Kleiner, Bruce; John W. Lott (2008). "Notes on Perelman's Papers". Geometry and Topology. 12 (5): 2587–2855. arXiv:math.DG/0605667. doi:10.2140/gt.2008.12.2587. S2CID 119133773.
  33. ^ Cao, Huai-Dong; Xi-Ping Zhu (June 2006). "A Complete Proof of the Poincaré and Geometrization Conjectures – application of the Hamilton-Perelman theory of the Ricci flow" (PDF). Asian Journal of Mathematics. 10 (2). Archived from the original (PDF) on 2012-05-14.
  34. ^ Cao, Huai-Dong & Zhu, Xi-Ping (December 3, 2006). "Hamilton–Perelman's Proof of the Poincaré Conjecture and the Geometrization Conjecture". arXiv:math.DG/0612069.
  35. ^ Morgan, John; Gang Tian (2006). "Ricci Flow and the Poincaré Conjecture". arXiv:math.DG/0607607.
  36. ^ Morgan, John; Gang Tian (2007). Ricci Flow and the Poincaré Conjecture. Clay Mathematics Institute. ISBN 978-0-8218-4328-4.
  37. ^ Morgan, John; Tian, Gang (2015). "Correction to Section 19.2 of Ricci Flow and the Poincare Conjecture". arXiv:1512.00699 [math.DG].
  38. ^ Nasar, Sylvia; David Gruber (August 28, 2006). "Manifold destiny". The New Yorker. pp. 44–57. On-line version at the New Yorker website.
  39. ^ Chang, Kenneth (August 22, 2006). "Highest Honor in Mathematics Is Refused". The New York Times.
  40. ^ A Report on the Poincaré Conjecture. Special lecture by John Morgan.
  41. ^ "Prize for Resolution of the Poincaré Conjecture Awarded to Dr. Grigoriy Perelman". Clay Mathematics Institute. March 18, 2010. Archived from the original on 2010-03-22.
  42. ^ "Poincaré Conjecture". Clay Mathematics Institute. Retrieved 2018-10-04.
  43. ^ Malcolm Ritter (2010-07-01). "Russian mathematician rejects $1 million prize". Phys.Org. Retrieved 2011-05-15.

Further reading

  • Kleiner, Bruce; Lott, John (2008). "Notes on Perelman's papers". Geometry & Topology. 12 (5): 2587–2855. arXiv:math/0605667. doi:10.2140/gt.2008.12.2587. MR 2460872. S2CID 119133773.
  • Huai-Dong Cao; Xi-Ping Zhu (December 3, 2006). "Hamilton-Perelman's Proof of the Poincaré Conjecture and the Geometrization Conjecture". arXiv:math.DG/0612069.
  • Morgan, John W.; Tian, Gang (2007). Ricci Flow and the Poincaré Conjecture. Clay Mathematics Monographs. Vol. 3. Providence, RI: American Mathematical Society. arXiv:math/0607607. ISBN 978-0-8218-4328-4. MR 2334563.
  • O'Shea, Donal (2007). The Poincaré Conjecture: In Search of the Shape of the Universe. Walker & Company. ISBN 978-0-8027-1654-5.
  • Perelman, Grisha (November 11, 2002). "The entropy formula for the Ricci flow and its geometric applications". arXiv:math.DG/0211159.
  • Perelman, Grisha (March 10, 2003). "Ricci flow with surgery on three-manifolds". arXiv:math.DG/0303109.
  • Perelman, Grisha (July 17, 2003). "Finite extinction time for the solutions to the Ricci flow on certain three-manifolds". arXiv:math.DG/0307245.
  • Szpiro, George (2008). Poincaré's Prize: The Hundred-Year Quest to Solve One of Math's Greatest Puzzles. Plume. ISBN 978-0-452-28964-2.
  • Stillwell, John (2012). "Poincaré and the early history of 3-manifolds". Bulletin of the American Mathematical Society. 49 (4): 555–576. doi:10.1090/S0273-0979-2012-01385-X. MR 2958930.
  • Yau, Shing-Tung; Nadis, Steve (2019). The Shape of a Life: One Mathematician's Search for the Universe's Hidden Geometry. New Haven, CT: Yale University Press. ISBN 978-0-300-23590-6. MR 3930611.

External links

Wikiquote has quotations related to Poincaré conjecture.
  • v
  • t
  • e
Science Breakthroughs of the Year
Science
journal
Authority control databases: National Edit this at Wikidata
  • Israel
  • United States
  • Japan