Вы находитесь на странице: 1из 56

Contents

Articles
Poincaré conjecture 1
P versus NP problem 7
Hodge conjecture 17
Riemann hypothesis 21
Navier–Stokes existence and smoothness 44
Birch and Swinnerton-Dyer conjecture 48
Yang–Mills existence and mass gap 51

References
Article Sources and Contributors 53
Image Sources, Licenses and Contributors 54

Article Licenses
License 55
Poincaré conjecture 1

Poincaré conjecture
Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

In mathematics, the Poincaré


conjecture ([pwɛ̃kaʁe],[1]
English: /pwɛn.kɑˈreɪ/ pwen-kar-ay) is a
theorem about the characterization of
the three-dimensional sphere
For compact 2-dimensional surfaces without boundary, if every loop can be continuously
(3-sphere), which is the hypersphere tightened to a point, then the surface is topologically homeomorphic to a 2-sphere
that bounds the unit ball in (usually just called a sphere). The Poincaré conjecture asserts that the same is true for
four-dimensional space. The 3-dimensional surfaces.

conjecture states:

Every simply connected, closed 3-manifold is homeomorphic to the 3-sphere.


An equivalent form of the conjecture involves a coarser form of equivalence than homeomorphism called homotopy
equivalence: if a 3-manifold is homotopy equivalent to the 3-sphere, then it is necessarily homeomorphic to it.
Originally conjectured by Henri Poincaré, the theorem concerns a space that locally looks like ordinary
three-dimensional space but is connected, finite in size, and lacks any boundary (a closed 3-manifold). The Poincaré
conjecture claims 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. An analogous result has been known in higher
dimensions for some time.
After nearly a century of effort by mathematicians, Grigori Perelman presented a proof of the conjecture in three
papers made available in 2002 and 2003 on arXiv. The proof followed the program of Richard Hamilton. Several
high-profile teams of mathematicians have verified that Perelman's proof is correct.
The Poincaré conjecture, before being proven, was one of the most important open questions in topology. It is one of
the seven Millennium Prize Problems, for which the Clay Mathematics Institute offered a $1,000,000 prize for the
first correct solution. Perelman's work survived review and was confirmed in 2006, leading to his being offered a
Fields Medal, which he declined. Perelman was awarded the Millennium Prize on March 18, 2010.[2] On July 1,
2010, he turned down the prize saying that he believes his contribution in proving the Poincaré conjecture was no
greater than that of U.S. mathematician Richard Hamilton (who first suggested a program for the solution).[3] [4] The
Poincaré conjecture is the first and, as of April 2011, the only solved Millennium problem.
On December 22, 2006, the journal Science honored Perelman's proof of the Poincaré conjecture as the scientific
"Breakthrough of the Year", the first time this had been bestowed in the area of mathematics.[5]
Poincaré conjecture 2

History

Poincaré's question
At the beginning of the 20th century, Henri Poincaré was working on the foundations of topology—what would later
be called combinatorial topology and then algebraic topology. He was particularly interested in what topological
properties characterized a sphere.
Poincaré claimed in 1900 that homology, a tool he had devised based on prior work by Enrico Betti, was sufficient to
tell if a 3-manifold was a 3-sphere. However, in a 1904 paper he described a counterexample to this claim, a space
now called the Poincaré homology sphere. The Poincaré sphere was the first example of a homology sphere, a
manifold that had the same homology as a sphere, of which many others have since been constructed. To establish
that the Poincaré sphere was different from the 3-sphere, Poincaré introduced a new topological invariant, the
fundamental group, and showed that the Poincaré sphere had a fundamental group of order 120, while the 3-sphere
had a trivial fundamental group. In this way he was able to conclude that these two spaces were, indeed, different.
In the same paper, Poincaré wondered whether a 3-manifold with the homology of a 3-sphere and also trivial
fundamental group had to be a 3-sphere. Poincaré's new condition—i.e., "trivial fundamental group"—can be
restated as "every loop can be shrunk to a point."
The original phrasing was as follows:
Consider a compact 3-dimensional manifold V without boundary. Is it possible that the fundamental group of
V could be trivial, even though V is not homeomorphic to the 3-dimensional sphere?
Poincaré never declared whether he believed this additional condition would characterize the 3-sphere, but
nonetheless, the statement that it does is known as the Poincaré conjecture. Here is the standard form of the
conjecture:
Every simply connected, closed 3-manifold is homeomorphic to the 3-sphere.

Attempted solutions
This problem seems to have lain dormant for a time, until J. H. C. Whitehead revived interest in the conjecture, when
in the 1930s he first claimed a proof, and then retracted it. In the process, he discovered some interesting examples of
simply connected non-compact 3-manifolds not homeomorphic to R3, the prototype of which is now called the
Whitehead manifold.
In the 1950s and 1960s, other mathematicians were to claim proofs only to discover a flaw. Influential
mathematicians such as Bing, Haken, Moise, and Papakyriakopoulos attacked the conjecture. In 1958 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.[6] Bing also described some of the pitfalls in trying to
prove the Poincaré conjecture.[7]
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."[8] 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).[9] [10]
An exposition of attempts to prove this conjecture can be found in the non-technical book Poincaré's Prize by
George Szpiro.[11]
Poincaré conjecture 3

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 is necessary; in dimensions four and higher there are
simply-connected manifolds which are not homeomorphic 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 dimension four. 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:[12]
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 Perelman's solution


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.[13] 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 using ideas from
Hamilton's work.

In late 2002 and 2003 Perelman posted three papers on the arXiv.[14] [15] [16]
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 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 the arXiv in May 2006
which filled in the details of Perelman's proof of the geometrization
conjecture.[17]
• Huai-Dong Cao and Xi-Ping Zhu published a paper in the June 2006 issue
of the Asian Journal of Mathematics giving a complete proof of the Several stages of the Ricci flow on a
Poincaré and geometrization conjectures.[18] They initially claimed the two-dimensional manifold.

proof as their own achievement based on the "Hamilton-Perelman theory",


but later retracted the original version of their paper, and posted a revised version, in which they referred to their
work as the more modest "exposition of Hamilton–Perelman's proof".[19] They were also forced to publish an
erratum disclosing that they had failed to cite properly the previous work of Kleiner and Lott published in 2003.
In the same issue, the AJM editorial board issued an apology for what it called "incautions" in the Cao–Zhu
Poincaré conjecture 4

paper.
• John Morgan and Gang Tian posted a paper on the arXiv in July 2006 which gave a detailed proof of just the
Poincaré Conjecture (which is somewhat easier than the full geometrization conjecture)[20] and expanded this to a
book.[21]
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 conjecture, but Perelman
refused the medal.[22] [23] [24] John Morgan spoke at the ICM on the Poincaré conjecture on August 24, 2006,
declaring that "in 2003, Perelman solved the Poincaré Conjecture."[25]
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 idea is to try to improve this metric; for example, if the metric can be
improved enough so that it has constant curvature, then it must be the 3-sphere. The metric is improved using the
Ricci flow equations;

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, if the manifold has positive Ricci curvature
everywhere he showed that the manifold becomes extinct in finite time under Ricci flow without any other
singularities. (In other words, the manifold collapses to a point in finite time; it is easy to describe the structure just
before the manifold collapses.) This easily implies the Poincaré conjecture in the case of positive Ricci curvature.
However in general the Ricci flow equations lead to singularities of the metric after a finite time. Perelman showed
how to continue past these singularities: very roughly, 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.
A special case of Perelman's theorems about Ricci flow with surgery is given as follows.
The Ricci flow with surgery on a closed oriented 3-manifold is well defined for all time. If the fundamental
group is a free product of finite groups and cyclic groups then the Ricci flow with surgery becomes extinct in
finite time, and at all times all components of the manifold are connected sums of S2 bundles over S1 and
quotients of S3.
This result implies the Poincaré conjecture because it is easy to check it for the possible manifolds listed in the
conclusion.
The condition on the fundamental group turns out to be necessary (and sufficient) for finite time extinction, and in
particular includes the case of trivial fundamental group. 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. By studying the limit of the
manifold for large time, Perelman proved Thurston's geometrization conjecture for any fundamental group: 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, but this extra complication is not necessary for proving just the Poincaré conjecture.[26]
Poincaré conjecture 5

Notes
[1] "Poincaré, Jules Henri" (http:/ / www. bartleby. com/ 61/ 3/ P0400300. html). The American Heritage Dictionary of the English Language
(fourth edition ed.). Boston: Houghton Mifflin Company. 2000. ISBN 0-395-82517-2. . Retrieved 2007-05-05..
[2] Clay Mathematics Institute (March 18, 2010). "Prize for Resolution of the Poincaré Conjecture Awarded to Dr. Grigoriy Perelman" (http:/ /
www. claymath. org/ poincare/ millenniumPrizeFull. pdf) (PDF). Press release. . Retrieved March 18, 2010. "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."
[3] Последнее "нет" доктора Перельмана (http:/ / www. interfax. ru/ society/ txt. asp?id=143603), Interfax 1 July 2010
[4] Ritter, Malcolm (1 July 2010). "Russian mathematician rejects million prize" (http:/ / www. boston. com/ news/ science/ articles/ 2010/ 07/
01/ russian_mathematician_rejects_1_million_prize/ ?p1=Well_MostPop_Emailed1). The Boston Globe. .
[5] Mackenzie, Dana (2006-12-22). "The Poincaré Conjecture--Proved" (http:/ / www. sciencemag. org/ cgi/ content/ full/ 314/ 5807/ 1848).
Science (American Association for the Advancement of Science) 314 (5807): 1848–1849. doi:10.1126/science.314.5807.1848.
PMID 17185565. ISSN: 0036-8075. .
[6] Bing, RH (1958). "Necessary and sufficient conditions that a 3-manifold be S3". The Annals of Mathematics, 2nd Ser. 68 (1): 17–37.
doi:10.2307/1970041. JSTOR 1970041.
[7] Bing, RH (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.
[8] Milnor, John (2004). "The Poincaré Conjecture 99 Years Later: A Progress Report" (http:/ / www. math. sunysb. edu/ ~jack/ PREPRINTS/
poiproof. pdf) (PDF). . Retrieved 2007-05-05.
[9] Taubes, Gary (July 1987). "What happens when hubris meets nemesis". Discover 8: 66–77.
[10] Matthews, Robert (9 April 2002). "$1 million mathematical mystery "solved"" (http:/ / www. newscientist. com/ article. ns?id=dn2143).
NewScientist.com. . Retrieved 2007-05-05.
[11] Szpiro, George (July 29, 2008). Poincaré's Prize: The Hundred-Year Quest to Solve One of Math's Greatest Puzzles. Plume.
ISBN 978-0-452-28964-2.
[12] 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
[13] Hamilton, Richard (1982). "Three-manifolds with positive Ricci curvature". Journal of Differential Geometry 17: 255–306. Reprinted in:
Cao, H.D.; et al. (Editors) (2003). Collected Papers on Ricci Flow. International Press. ISBN 978-1571461100.
[14] Perelman, Grigori (2002). The entropy formula for the Ricci flow and its geometric applications. arXiv:math.DG/0211159.
[15] Perelman, Grigori (2003). Ricci flow with surgery on three-manifolds. arXiv:math.DG/0303109.
[16] Perelman, Grigori (2003). Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv:math.DG/0307245.
[17] Kleiner, Bruce; John W. Lott (2006). Notes on Perelman's Papers. arXiv:math.DG/0605667.
[18] 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" (http:/ / www. intlpress. com/ AJM/ p/ 2006/ 10_2/ AJM-10-2-165-492. pdf) (PDF). Asian
Journal of Mathematics 10 (2). .
[19] Cao, Huai-Dong and Zhu, Xi-Ping (December 3, 2006). "Hamilton–Perelman's Proof of the Poincaré Conjecture and the Geometrization
Conjecture". arXiv.org. arXiv:math.DG/0612069.
[20] Morgan, John; Gang Tian (2006). Ricci Flow and the Poincaré Conjecture. arXiv:math.DG/0607607.
[21] Morgan, John; Gang Tian (2007). Ricci Flow and the Poincaré Conjecture. Clay Mathematics Institute. ISBN 0821843281.
[22] Nasar, Sylvia; David Gruber (August 28, 2006). "Manifold destiny". The New Yorker: pp. 44–57. On-line version at the New Yorker website
(http:/ / www. newyorker. com/ fact/ content/ articles/ 060828fa_fact2).
[23] Chang, Kenneth (August 22, 2006). "Highest Honor in Mathematics Is Refused" (http:/ / www. nytimes. com/ 2006/ 08/ 22/ science/
22cnd-math. html?hp& ex=1156305600& en=aa3a9d418768062c& ei=5094& partner=homepage). New York Times. .
[24] "Reclusive Russian solves 100-year-old maths problem" (http:/ / www. chinadaily. com. cn/ cndy/ 2006-08/ 23/ content_671442. htm).
China Daily: p. 7. 23 August 2006. .
[25] A Report on the Poincaré Conjecture. Special lecture by John Morgan.
[26] Terence Tao wrote an exposition of Ricci flow with surgery in: Tao, Terence (2006). Perelman's proof of the Poincaré conjecture: a
nonlinear PDE perspective. arXiv:math.DG/0610903.
Poincaré conjecture 6

External links
• The Poincaré conjecture described (http://www.claymath.org/prizeproblems/poincare.htm) by the Clay
Mathematics Institute.
• The Poincaré Conjecture (video) (http://www.youtube.com/watch?v=AUoaTrQTM5o) Brief visual overview
of the Poincaré Conjecture, background and solution.
• The Geometry of 3-Manifolds(video) (http://athome.harvard.edu/threemanifolds/) A public lecture on the
Poincaré and geometrization conjectures, given by C. McMullen at Harvard in 2006.
• Bruce Kleiner (Yale) and John W. Lott (University of Michigan): "Notes & commentary on Perelman's Ricci flow
papers" (http://www.math.lsa.umich.edu/~lott/ricciflow/perelman.html).
• Stephen Ornes, What is The Poincaré Conjecture? (http://www.seedmagazine.com/news/2006/08/
what_is_the_poincar_conjecture.php), Seed Magazine, 25 August 2006.
• The slides (http://www.mcm.ac.cn/Active/yau_new.pdf) used by Yau in a popular talk on the Poincaré
conjecture.
• "The Poincaré Conjecture" (http://www.bbc.co.uk/radio4/history/inourtime/inourtime_20061102.shtml) –
BBC Radio 4 programme In Our Time, 2 November 2006. Contributors June Barrow-Green, Lecturer in the
History of Mathematics at the Open University, Ian Stewart, Professor of Mathematics at the University of
Warwick, Marcus du Sautoy, Professor of Mathematics at the University of Oxford, and presenter Melvyn Bragg.
• "Solving an Old Math Problem Nets Award, Trouble" (http://www.npr.org/templates/story/story.
php?storyId=6682439) – NPR segment, December 26, 2006.
• Nasar, Sylvia; and Gruber, David (21 August 2006). "Manifold Destiny: A legendary problem and the battle over
who solved it." (http://www.newyorker.com/fact/content/articles/060828fa_fact2). The New Yorker.
Retrieved 2006-08-24.
P versus NP problem 7

P versus NP problem
Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

The P versus NP problem is a major unsolved


problem in computer science. Informally, it asks
whether every problem whose solution can be
efficiently checked by a computer can also be
efficiently solved by a computer. It was introduced in
1971 by Stephen Cook in his paper "The complexity of
theorem proving procedures"[2] and is considered by
many to be the most important open problem in the
field.[3] It is one of the seven Millennium Prize
Problems selected by the Clay Mathematics Institute to
carry a US$ 1,000,000 prize for the first correct Diagram of complexity classes provided that P ≠ NP. The existence
of problems outside both P and NP-complete in this case was
solution. [1]
established by Ladner's theorem.

In essence, the question P = NP? asks:


Suppose that solutions to a problem can be verified quickly. Then, can the solutions themselves also be
computed quickly?
The theoretical notion of quick used here is an algorithm that runs in polynomial time. The general class of questions
for which some algorithm can provide an answer in polynomial time is called "class P" or just "P".
For some questions, there is no known way to find an answer quickly, but if one is provided with information
showing what the answer is, it may be possible to verify the answer quickly. The class of questions for which an
answer can be verified in polynomial time is called NP.
Consider the subset sum problem, an example of a problem that is easy to verify, but whose answer may be difficult
to compute. Given a set of integers, does some nonempty subset of them sum to 0? For instance, does a subset of the
set {−2, −3, 15, 14, 7, −10} add up to 0? The answer "yes, because {−2, −3, −10, 15} add up to zero" can be quickly
verified with three additions. However, there is no known algorithm to find such a subset in polynomial time (there
is, however, in exponential time, which consists of 2n-1 tries), and indeed such an algorithm cannot exist if the two
complexity classes are not the same; hence this problem is in NP (quickly checkable) but not necessarily in P
(quickly solvable).
An answer to the P = NP question would determine whether problems like the subset-sum problem that can be
verified in polynomial time can also be solved in polynomial time. If it turned out that P does not equal NP, it would
mean that there are problems in NP (such as NP-complete problems) that are harder to compute than to verify: they
could not be solved in polynomial time, but the answer could be verified in polynomial time.
P versus NP problem 8

Context
The relation between the complexity classes P and NP is studied in computational complexity theory, the part of the
theory of computation dealing with the resources required during computation to solve a given problem. The most
common resources are time (how many steps it takes to solve a problem) and space (how much memory it takes to
solve a problem).
In such analysis, a model of the computer for which time must be analyzed is required. Typically such models
assume that the computer is deterministic (given the computer's present state and any inputs, there is only one
possible action that the computer might take) and sequential (it performs actions one after the other).
In this theory, the class P consists of all those decision problems (defined below) that can be solved on a
deterministic sequential machine in an amount of time that is polynomial in the size of the input; the class NP
consists of all those decision problems whose positive solutions can be verified in polynomial time given the right
information, or equivalently, whose solution can be found in polynomial time on a non-deterministic machine.[4]
Clearly, P ⊆ NP. Arguably the biggest open question in theoretical computer science concerns the relationship
between those two classes:
Is P equal to NP?
In a 2002 poll of 100 researchers, 61 believed the answer to be no, 9 believed the answer is yes, and 22 were unsure;
8 believed the question may be independent of the currently accepted axioms and so impossible to prove or
disprove.[5]

NP-complete
To attack the P = NP question the concept
of NP-completeness is very useful.
Informally the NP-complete problems are
the "toughest" problems in NP in the sense
that they are the ones most likely not to be
in P. NP-complete problems are a set of
problems that any other NP-problem can be
reduced to in polynomial time, but retain the
ability to have their solution verified in
polynomial time. In comparison, NP-hard
problems are those at least as hard as
NP-complete problems, meaning all
NP-problems can be reduced to them, but
not all NP-hard problems are in NP, Euler diagram for P, NP, NP-complete, and NP-hard set of problems
meaning not all of them have solutions
verifiable in polynomial time.

For instance, the decision problem version of the travelling salesman problem is NP-complete, so any instance of
any problem in NP can be transformed mechanically into an instance of the traveling salesman problem, in
polynomial time. The traveling salesman problem is one of many such NP-complete problems. If any NP-complete
problem is in P, then it would follow that P = NP. Unfortunately, many important problems have been shown to be
NP-complete, and as of 2011 not a single fast algorithm for any of them is known.
Based on the definition alone it's not obvious that NP-complete problems exist. A trivial and contrived NP-complete
problem can be formulated as: given a description of a Turing machine M guaranteed to halt in polynomial time,
does there exist a polynomial-size input that M will accept?[6] It is in NP because (given an input) it is simple to
P versus NP problem 9

check whether or not M accepts the input by simulating M; it is NP-complete because the verifier for any particular
instance of a problem in NP can be encoded as a polynomial-time machine M that takes the solution to be verified as
input. Then the question of whether the instance is a yes or no instance is determined by whether a valid input exists.
The first natural problem proven to be NP-complete was the Boolean satisfiability problem. This result came to be
known as Cook–Levin theorem; its proof that satisfiability is NP-complete contains technical details about Turing
machines as they relate to the definition of NP. However, after this problem was proved to be NP-complete, proof by
reduction provided a simpler way to show that many other problems are in this class. Thus, a vast class of seemingly
unrelated problems are all reducible to one another, and are in a sense "the same problem".

Harder problems
Although it is unknown whether P = NP, problems outside of P are known. A number of succinct problems
(problems that operate not on normal input, but on a computational description of the input) are known to be
EXPTIME-complete. Because it can be shown that P EXPTIME, these problems are outside P, and so require
more than polynomial time. In fact, by the time hierarchy theorem, they cannot be solved in significantly less than
exponential time. Examples include finding a perfect strategy for chess (on an N×N board)[7] and some other board
games.[8]
The problem of deciding the truth of a statement in Presburger arithmetic requires even more time. Fischer and
Rabin proved in 1974 that every algorithm that decides the truth of Presburger statements has a runtime of at least
for some constant c. Here, n is the length of the Presburger statement. Hence, the problem is known to need
more than exponential run time. Even more difficult are the undecidable problems, such as the halting problem. They
cannot be completely solved by any algorithm, in the sense that for any particular algorithm there is at least one input
for which that algorithm will not produce the right answer; it will either produce the wrong answer, finish without
giving a conclusive answer, or otherwise run forever without producing any answer at all.

Problems in NP not known to be in P or NP-complete


It was shown by Ladner that if P ≠ NP then there exist problems in NP that are neither in P nor NP-complete.[1] Such
problems are called NP-intermediate problems. The graph isomorphism problem, the discrete logarithm problem and
the integer factorization problem are examples of problems believed to be NP-intermediate. They are some of the
very few NP problems not known to be in P or to be NP-complete.
The graph isomorphism problem is the computational problem of determining whether two finite graphs are
isomorphic. An important unsolved problem in complexity theory is whether the graph isomorphism problem is in P,
NP-complete, or NP-intermediate. The answer is not known, but it is believed that the problem is at least not
NP-complete.[9] If graph isomorphism is NP-complete, the polynomial time hierarchy collapses to its second
level.[10] Since it is widely believed that the polynomial hierarchy does not collapse to any finite level, it is believed
that graph isomorphism is not NP-complete. The best algorithm for this problem, due to Laszlo Babai and Eugene
Luks has run time 2O(√(n log n)) for graphs with n vertices.
The integer factorization problem is the computational problem of determining the prime factorization of a given
integer. Phrased as a decision problem, it is the problem of deciding whether the input has a factor less than k. No
efficient integer factorization algorithm is known, and this fact forms the basis of several modern cryptographic
systems, such as the RSA algorithm. The integer factorization problem is in NP and in co-NP (and even in UP and
co-UP[11] ). If the problem is NP-complete, the polynomial time hierarchy will collapse to its first level (i.e., NP will
equal co-NP). The best known algorithm for integer factorization is the general number field sieve, which takes
expected time O(e(64/9)1/3(n.log 2)1/3(log (n.log 2))2/3) to factor an n-bit integer. However, the best known quantum
algorithm for this problem, Shor's algorithm, does run in polynomial time. Unfortunately, this fact doesn't say much
about where the problem lies with respect to non-quantum complexity classes.
P versus NP problem 10

Does P mean "easy"?


All of the above discussion has assumed
that P means "easy" and "not in P" means
"hard". This assumption, known as
Cobham's thesis, though a common and
reasonably accurate assumption in
complexity theory, is not always true in
practice; the size of constant factors or
exponents may have practical importance,
or there may be solutions that work for
situations encountered in practice despite
having poor worst-case performance in
theory (this is the case for instance for the
The graph shows time (average of 100 instances in msec using a 933 MHz Pentium
simplex algorithm in linear III) vs.problem size for knapsack problems for a state-of-the-art specialized
programming). Other solutions violate the algorithm. Quadratic fit suggests that empirical algorithmic complexity for instances
[12]
Turing machine model on which P and with 50–10,000 variables is O((log n)2).

NP are defined by introducing concepts


like randomness and quantum computation.

Because of these factors, even if a problem is shown to be NP-complete, and even if P ≠ NP, there may still be
effective approaches to tackling the problem in practice. There are algorithms for many NP-complete problems, such
as the knapsack problem, the travelling salesman problem and the boolean satisfiability problem, that can solve to
optimality many real-world instances in reasonable time. The empirical average-case complexity (time vs. problem
size) of such algorithms can be surprisingly low.

Reasons to believe P ≠ NP
According to a poll,[5] many computer scientists believe that P ≠ NP. A key reason for this belief is that after
decades of studying these problems no one has been able to find a polynomial-time algorithm for any of more than
3000 important known NP-complete problems (see List of NP-complete problems). These algorithms were sought
long before the concept of NP-completeness was even defined (Karp's 21 NP-complete problems, among the first
found, were all well-known existing problems at the time they were shown to be NP-complete). Furthermore, the
result P = NP would imply many other startling results that are currently believed to be false, such as NP = co-NP
and P = PH.
It is also intuitively argued that the existence of problems that are hard to solve but for which the solutions are easy
to verify matches real-world experience.[13]
If P = NP, then the world would be a profoundly different place than we usually assume it to be. There would
be no special value in "creative leaps," no fundamental gap between solving a problem and recognizing the
solution once it's found. Everyone who could appreciate a symphony would be Mozart; everyone who could
follow a step-by-step argument would be Gauss...
— Scott Aaronson, MIT
On the other hand, some researchers believe that there is overconfidence in believing P ≠ NP and that researchers
should explore proofs of P = NP as well. For example, in 2002 these statements were made:[5]
The main argument in favor of P ≠ NP is the total lack of fundamental progress in the area of exhaustive
search. This is, in my opinion, a very weak argument. The space of algorithms is very large and we are only at
the beginning of its exploration. [. . .] The resolution of Fermat's Last Theorem also shows that very simple
P versus NP problem 11

questions may be settled only by very deep theories.


—Moshe Y. Vardi, Rice University
Being attached to a speculation is not a good guide to research planning. One should always try both directions
of every problem. Prejudice has caused famous mathematicians to fail to solve famous problems whose
solution was opposite to their expectations, even though they had developed all the methods required.
—Anil Nerode, Cornell University

Consequences of proof
One of the reasons the problem attracts so much attention is the consequences of the answer. A proof that P = NP
could have stunning practical consequences, if the proof leads to efficient methods for solving some of the important
problems in NP. It is also possible that a proof would not lead directly to efficient methods, perhaps if the proof is
non-constructive, or the size of the bounding polynomial is too big to be efficient in practice. The consequences,
both positive and negative, arise since various NP-complete problems are fundamental in many fields.
Cryptography, for example, relies on certain problems being difficult. A constructive and efficient solution to an
NP-complete problem such as 3-SAT would break most existing cryptosystems including public-key cryptography, a
foundation for many modern security applications such as secure economic transactions over the Internet, and
symmetric ciphers such as AES or 3DES, used for the encryption of communications data. These would need to be
modified or replaced by information-theoretically secure solutions.
On the other hand, there are enormous positive consequences that would follow from rendering tractable many
currently mathematically intractable problems. For instance, many problems in operations research are NP-complete,
such as some types of integer programming, and the travelling salesman problem, to name two of the most famous
examples. Efficient solutions to these problems would have enormous implications for logistics. Many other
important problems, such as some problems in protein structure prediction, are also NP-complete;[14] if these
problems were efficiently solvable it could spur considerable advances in biology.
But such changes may pale in significance compared to the revolution an efficient method for solving NP-complete
problems would cause in mathematics itself. According to Stephen Cook,[15]
...it would transform mathematics by allowing a computer to find a formal proof of any theorem which has a
proof of a reasonable length, since formal proofs can easily be recognized in polynomial time. Example
problems may well include all of the CMI prize problems.
Research mathematicians spend their careers trying to prove theorems, and some proofs have taken decades or even
centuries to find after problems have been stated—for instance, Fermat's Last Theorem took over three centuries to
prove. A method that is guaranteed to find proofs to theorems, should one exist of a "reasonable" size, would
essentially end this struggle.
A proof that showed that P ≠ NP would lack the practical computational benefits of a proof that P = NP, but would
nevertheless represent a very significant advance in computational complexity theory and provide guidance for
future research. It would allow one to show in a formal way that many common problems cannot be solved
efficiently, so that the attention of researchers can be focused on partial solutions or solutions to other problems. Due
to widespread belief in P ≠ NP, much of this focusing of research has already taken place.[16]
A "not equal" resolution to the P versus NP problem still leaves open the average-case complexity of hard problems
in NP. For example, it is possible that SAT requires exponential time in the worst case, but that almost all randomly
selected instances of it are efficiently solvable. Russell Impagliazzo has described five hypothetical "worlds" that
could result from different possible resolutions to the average-case complexity question.[17] These range from
"Algorithmica", where P=NP and problems like SAT can be solved efficiently in all instances, to "Cryptomania",
where P≠NP and generating hard instances of problems outside P is easy, with three intermediate possibilities
reflecting different possible distributions of difficulty over instances of NP-hard problems. The "world" where P≠NP
P versus NP problem 12

but all problems in NP are tractable in the average case is called "Heuristica" in the paper. A Princeton University
workshop in 2009 studied the status of the five worlds.[18]

Results about difficulty of proof


Although the P = NP? problem itself remains open, despite a million-dollar prize and a huge amount of dedicated
research, efforts to solve the problem have led to several new techniques. In particular, some of the most fruitful
research related to the P = NP problem has been in showing that existing proof techniques are not powerful enough
to answer the question, thus suggesting that novel technical approaches are required.
As additional evidence for the difficulty of the problem, essentially all known proof techniques in computational
complexity theory fall into one of the following classifications, each of which is known to be insufficient to prove
that P ≠ NP:

Classification Definition

Relativizing
Imagine a world where every algorithm is allowed to make queries to some fixed subroutine called an oracle, and the running time
proofs
of the oracle is not counted against the running time of the algorithm. Most proofs (especially classical ones) apply uniformly in a
world with oracles regardless of what the oracle does. These proofs are called relativizing. In 1975, Baker, Gill, and Solovay
[19]
showed that P = NP with respect to some oracles, while P ≠ NP for other oracles. Since relativizing proofs can only prove
statements that are uniformly true with respect to all possible oracles, this showed that relativizing techniques cannot resolve P =
NP.

Natural proofs In 1993, Alexander Razborov and Steven Rudich defined a general class of proof techniques for circuit complexity lower bounds,
called natural proofs. At the time all previously known circuit lower bounds were natural, and circuit complexity was considered a
very promising approach for resolving P = NP. However, Razborov and Rudich showed that, if one-way functions exist, then no
natural proof method can distinguish between P and NP. Although one-way functions have never been formally proven to exist,
most mathematicians believe that they do, and a proof or disproof of their existence would be a much stronger statement than the
quantification of P relative to NP. Thus it is unlikely that natural proofs alone can resolve P = NP.

Algebrizing After the Baker-Gill-Solovay result, new non-relativizing proof techniques were successfully used to prove that IP = PSPACE.
proofs However, in 2008, Scott Aaronson and Avi Wigderson showed that the main technical tool used in the IP = PSPACE proof,
[20]
known as arithmetization, was also insufficient to resolve P = NP.

These barriers are another reason why NP-complete problems are useful: if a polynomial-time algorithm can be
demonstrated for an NP-complete problem, this would solve the P = NP problem in a way not excluded by the above
results.
These barriers have also led some computer scientists to suggest that the P versus NP problem may be independent
of standard axiom systems like ZFC (cannot be proved or disproved within them). The interpretation of an
independence result could be that either no polynomial-time algorithm exists for any NP-complete problem, and
such a proof cannot be constructed in (e.g.) ZFC, or that polynomial-time algorithms for NP-complete problems may
exist, but it's impossible to prove in ZFC that such algorithms are correct.[21] However, if it can be shown, using
techniques of the sort that are currently known to be applicable, that the problem cannot be decided even with much
weaker assumptions extending the Peano axioms (PA) for integer arithmetic, then there would necessarily exist
nearly-polynomial-time algorithms for every problem in NP.[22] Therefore, if one believes (as most complexity
theorists do) that not all problems in NP have efficient algorithms, it would follow that proofs of independence using
those techniques cannot be possible. Additionally, this result implies that proving independence from PA or ZFC
using currently known techniques is no easier than proving the existence of efficient algorithms for all problems in
NP.
P versus NP problem 13

Claimed solutions
While the P versus NP problem is generally considered unsolved,[23] many amateur and some professional
researchers have claimed solutions. Woeginger (2010) has a comprehensive list.[24] An August 2010 claim of proof
that P ≠ NP, by Vinay Deolalikar, researcher at HP Labs, Palo Alto, received heavy Internet and press attention after
being initially described as "seem[ing] to be a relatively serious attempt" by two leading specialists.[25] The proof has
been reviewed publicly by academics,[26] [27] and Neil Immerman, an expert in the field, had pointed out two
possibly fatal errors in the proof.[28] As of September 15, 2010, Deolalikar was reported to be working on a detailed
expansion of his attempted proof.[29] However, the general consensus amongst theoretical computer scientists is now
that the attempted proof is not correct, or even a significant advancement in our understanding of the problem.

Logical characterizations
The P = NP problem can be restated in terms of expressible certain classes of logical statements, as a result of work
in descriptive complexity. All languages (of finite structures with a fixed signature including a linear order relation)
in P can be expressed in first-order logic with the addition of a suitable least fixed point combinator (effectively, this,
in combination with the order, allows the definition of recursive functions); indeed, (as long as the signature contains
at least one predicate or function in addition to the distinguished order relation [so that the amount of space taken to
store such finite structures is actually polynomial in the number of elements in the structure]), this precisely
characterizes P. Similarly, NP is the set of languages expressible in existential second-order logic—that is,
second-order logic restricted to exclude universal quantification over relations, functions, and subsets. The languages
in the polynomial hierarchy, PH, correspond to all of second-order logic. Thus, the question "is P a proper subset of
NP" can be reformulated as "is existential second-order logic able to describe languages (of finite linearly ordered
structures with nontrivial signature) that first-order logic with least fixed point cannot?". The word "existential" can
even be dropped from the previous characterization, since P = NP if and only if P = PH (as the former would
establish that NP = co-NP, which in turn implies that NP = PH). PSPACE = NPSPACE as established Savitch's
theorem, this follows directly from the fact that the square of a polynomial function is still a polynomial function.
However, it is believed, but not proven, a similar relationship may not exist between the polynomial time complexity
classes, P and NP so the question is still open.

Polynomial-time algorithms
No algorithm for any NP-complete problem is known to run in polynomial time. However, there are algorithms for
NP-complete problems with the property that if P = NP, then the algorithm runs in polynomial time (although with
enormous constants, making the algorithm impractical). The following algorithm, due to Levin, is such an example.
It correctly accepts the NP-complete language SUBSET-SUM, and runs in polynomial time if and only if P = NP:

// Algorithm that accepts the NP-complete language SUBSET-SUM.


//
// This is a polynomial-time algorithm if and only if P=NP.
//
// "Polynomial-time" means it returns "yes" in polynomial time when
// the answer should be "yes", and runs forever when it is "no".
//
// Input: S = a finite set of integers
// Output: "yes" if any subset of S adds up to 0.
// Runs forever with no output otherwise.
// Note: "Program number P" is the program obtained by
// writing the integer P in binary, then
P versus NP problem 14

// considering that string of bits to be a


// program. Every possible program can be
// generated this way, though most do nothing
// because of syntax errors.

FOR N = 1...infinity
FOR P = 1...N
Run program number P for N steps with input S
IF the program outputs a list of distinct integers
AND the integers are all in S
AND the integers sum to 0

THEN
OUTPUT "yes" and HALT

If, and only if, P = NP, then this is a polynomial-time algorithm accepting an NP-complete language. "Accepting"
means it gives "yes" answers in polynomial time, but is allowed to run forever when the answer is "no".
This algorithm is enormously impractical, even if P = NP. If the shortest program that can solve SUBSET-SUM in
polynomial time is b bits long, the above algorithm will try at least 2b−1 other programs first.

Formal definitions for P and NP


Conceptually a decision problem is a problem that takes as input some string w over an alphabet , and outputs
"yes" or "no". If there is an algorithm (say a Turing machine, or a computer program with unbounded memory) that
can produce the correct answer for any input string of length n in at most steps, where k and c are
constants independent of the input string, then we say that the problem can be solved in polynomial time and we
place it in the class P. Formally, P is defined as the set of all languages that can be decided by a deterministic
polynomial-time Turing machine. That is,
P=
where
and a deterministic polynomial-time Turing machine is a deterministic Turing machine M that satisfies the following
two conditions:
1. ; and
2. there exists such that (where O refers to the big O notation),
where
and
NP can be defined similarly using nondeterministic Turing machines (the traditional way). However, a modern
approach to define NP is to use the concept of certificate and verifier. Formally, NP is defined as the set of
languages over a finite alphabet that have a verifier that runs in polynomial time, where the notion of "verifier" is
defined as follows.
Let L be a language over a finite alphabet, .
if, and only if, there exists a binary relation and a positive integer k such that the
following two conditions are satisfied:
1. For all , such that and ; and
2. the language over is decidable by a Turing machine in polynomial
time.
P versus NP problem 15

A Turing machine that decides is called a verifier for L and a y such that is called a certificate of
membership of x in L.
In general, a verifier does not have to be polynomial-time. However, for L to be in NP, there must be a verifier that
runs in polynomial time.

Example
Let and

Clearly, the question of whether a given x is a composite is equivalent to the question of whether x is a member of
. It can be shown that by verifying that satisfies
the above definition (if we identify natural numbers with their binary representations).
also happens to be in P.[30] [31]

Formal definition for NP-completeness


There are many equivalent ways of describing NP-completeness.
Let be a language over a finite alphabet .
is NP-complete if, and only if, the following two conditions are satisfied:
1. ; and
2. any is polynomial-time-reducible to (written as ), where if, and only if, the
following two conditions are satisfied:
1. There exists such that ; and
2. there exists a polynomial-time Turing machine that halts with on its tape on any input .

Notes
[1] R. E. Ladner "On the structure of polynomial time reducibility," J.ACM, 22, pp. 151–171, 1975. Corollary 1.1. ACM site (http:/ / portal. acm.
org/ citation. cfm?id=321877& dl=ACM& coll=& CFID=15151515& CFTOKEN=6184618).
[2] Cook, Stephen (1971). "The complexity of theorem proving procedures" (http:/ / portal. acm. org/ citation. cfm?coll=GUIDE& dl=GUIDE&
id=805047). Proceedings of the Third Annual ACM Symposium on Theory of Computing. pp. 151–158. .
[3] Lance Fortnow, The status of the P versus NP problem (http:/ / www. cs. uchicago. edu/ ~fortnow/ papers/ pnp-cacm. pdf), Communications
of the ACM 52 (2009), no. 9, pp. 78–86. doi:10.1145/1562164.1562186
[4] Sipser, Michael: Introduction to the Theory of Computation, Second Edition, International Edition, page 270. Thomson Course Technology,
2006. Definition 7.19 and Theorem 7.20.
[5] William I. Gasarch (June 2002). "The P=?NP poll." (http:/ / www. cs. umd. edu/ ~gasarch/ papers/ poll. pdf) (PDF). SIGACT News 33 (2):
34–47. doi:10.1145/1052796.1052804. . Retrieved 2008-12-29.
[6] Scott Aaronson. "PHYS771 Lecture 6: P, NP, and Friends" (http:/ / www. scottaaronson. com/ democritus/ lec6. html). . Retrieved
2007-08-27.
[7] Aviezri Fraenkel and D. Lichtenstein (1981). "Computing a perfect strategy for n×n chess requires time exponential in n". J. Comb. Th. A
(31): 199–214.
[8] David Eppstein. "Computational Complexity of Games and Puzzles" (http:/ / www. ics. uci. edu/ ~eppstein/ cgt/ hard. html). .
[9] Arvind, Vikraman; Kurur, Piyush P. (2006). "Graph isomorphism is in SPP". Information and Computation 204 (5): 835–852.
doi:10.1016/j.ic.2006.02.002.
[10] Uwe Schöning, "Graph isomorphism is in the low hierarchy", Proceedings of the 4th Annual Symposium on Theoretical Aspects of
Computer Science, 1987, 114–124; also: Journal of Computer and System Sciences, vol. 37 (1988), 312–323
[11] Lance Fortnow. Computational Complexity Blog: Complexity Class of the Week: Factoring. September 13, 2002. http:/ / weblog. fortnow.
com/ 2002/ 09/ complexity-class-of-week-factoring. html
[12] Pisinger, D. 2003. "Where are the hard knapsack problems?" Technical Report 2003/08, Department of Computer Science, University of
Copenhagen, Copenhagen, Denmark
[13] Scott Aaronson. "Reasons to believe" (http:/ / scottaaronson. com/ blog/ ?p=122). ., point 9.
[14] Berger B, Leighton T (1998). "Protein folding in the hydrophobic-hydrophilic (HP) model is NP-complete". J. Comput. Biol. 5 (1): 27–40.
doi:10.1089/cmb.1998.5.27. PMID 9541869.
P versus NP problem 16

[15] Cook, Stephen (April 2000). The P versus NP Problem (http:/ / www. claymath. org/ millennium/ P_vs_NP/ Official_Problem_Description.
pdf). Clay Mathematics Institute. . Retrieved 2006-10-18.
[16] L. R. Foulds (October 1983). "The Heuristic Problem-Solving Approach" (http:/ / www. jstor. org/ pss/ 2580891). The Journal of the
Operational Research Society 34 (10): 927–934. doi:10.2307/2580891. .
[17] R. Impagliazzo, "A personal view of average-case complexity," (http:/ / cseweb. ucsd. edu/ ~russell/ average. ps) sct, pp.134, 10th Annual
Structure in Complexity Theory Conference (SCT'95), 1995
[18] http:/ / intractability. princeton. edu/ blog/ 2009/ 05/ program-for-workshop-on-impagliazzos-worlds/
[19] T. P. Baker, J. Gill, R. Solovay. Relativizations of the P =? NP Question. SIAM Journal on Computing, 4(4): 431–442 (1975)
[20] S. Aaronson and A. Wigderson. Algebrization: A New Barrier in Complexity Theory, in Proceedings of ACM STOC'2008, pp. 731–740.
[21] Aaronson, Scott. "Is P Versus NP Formally Independent?" (http:/ / www. scottaaronson. com/ papers/ pnp. pdf). .
[22] Ben-David, Shai; Halevi, Shai (1992). On the independence of P versus NP (http:/ / www. cs. technion. ac. il/ ~shai/ ph. ps. gz). Technical
Report. 714. Technion. .
[23] John Markoff (8 October 2009). "Prizes Aside, the P-NP Puzzler Has Consequences" (http:/ / www. nytimes. com/ 2009/ 10/ 08/ science/
Wpolynom. html). The New York Times. .
[24] Gerhard J. Woeginger (2010-08-09). "The P-versus-NP page" (http:/ / www. win. tue. nl/ ~gwoegi/ P-versus-NP. htm). . Retrieved
2010-08-12.
[25] Markoff, John (16 August 2010). "Step 1: Post Elusive Proof. Step 2: Watch Fireworks." (http:/ / www. nytimes. com/ 2010/ 08/ 17/ science/
17proof. html?_r=1). The New York Times. . Retrieved 20 September 2010.
[26] Polymath project wiki. "Deolalikar's P vs NP paper" (http:/ / michaelnielsen. org/ polymath1/ index. php?title=Deolalikar_P_vs_NP_paper).
.
[27] Science News, "Crowdsourcing peer review" (http:/ / www. sciencenews. org/ index/ generic/ activity/ view/ id/ 63252/ title/
Crowdsourcing_peer_review)
[28] Dick Lipton (12 August 2010). "Fatal Flaws in Deolalikar's Proof?" (http:/ / rjlipton. wordpress. com/ 2010/ 08/ 12/
fatal-flaws-in-deolalikars-proof/ ). .
[29] Dick Lipton (15 September 2010). "An Update on Vinay Deolalikar's Proof" (http:/ / rjlipton. wordpress. com/ 2010/ 09/ 15/
an-update-on-vinay-deolalikars-proof/ ). . Retrieved December 31, 2010.
[30] M. Agrawal, N. Kayal, N. Saxena. "Primes is in P" (http:/ / www. cse. iitk. ac. in/ users/ manindra/ algebra/ primality_v6. pdf) (PDF). .
Retrieved 2008-12-29.
[31] AKS primality test

Further reading
• Fraenkel, A. S.; Lichtenstein, D.. Computing a Perfect Strategy for n*n Chess Requires Time Exponential in N.
(http://www.pubzone.org/dblp/conf/icalp/FraenkelL81). doi:10.1007/3-540-10843-2+23.
• Garey, Michael (1979). Computers and Intractability. San Francisco: W.H. Freeman. ISBN 0716710455.
• Goldreich, Oded (2010). P, Np, and Np-Completeness. Cambridge: Cambridge University Press.
ISBN 9780521122542.
• Immerman, N. (1983). Languages which capture complexity classes. pp. 347. doi:10.1145/800061.808765.
• Cormen, Thomas (2001). Introduction to Algorithms. Cambridge: MIT Press. ISBN 0262032937.
• Papadimitriou, Christos (1994). Computational Complexity. Boston: Addison-Wesley. ISBN 0201530821.
• Fortnow, L. (2009). "The status of the P versus NP problem". Communications of the ACM 52 (9): 78.
doi:10.1145/1562164.1562186.

External links
• The Clay Mathematics Institute Millennium Prize Problems (http://www.claymath.org/millennium/)
• The Clay Math Institute Official Problem Description (http://www.claymath.org/millennium/P_vs_NP/
Official_Problem_Description.pdf)PDF (118 KB)
• Ian Stewart on Minesweeper as NP-complete at The Clay Math Institute (http://www.claymath.org/
Popular_Lectures/Minesweeper/)
• Gerhard J. Woeginger. The P-versus-NP page (http://www.win.tue.nl/~gwoegi/P-versus-NP.htm). A list of
links to a number of purported solutions to the problem. Some of these links state that P equals NP, some of them
state the opposite. It is probable that all these alleged solutions are incorrect.
• Computational Complexity of Games and Puzzles (http://www.ics.uci.edu/~eppstein/cgt/hard.html)
P versus NP problem 17

• Complexity Zoo: Class P (http://qwiki.stanford.edu/index.php/Complexity_Zoo:P#p), Complexity Zoo: Class


NP (http://qwiki.stanford.edu/index.php/Complexity_Zoo:N#np)
• Scott Aaronson 's Shtetl Optimized blog: Reasons to believe (http://scottaaronson.com/blog/?p=122), a list of
justifications for the belief that P ≠ NP

Hodge conjecture
The Hodge conjecture is a major unsolved problem in algebraic geometry which relates the algebraic topology of a
non-singular complex algebraic variety and the subvarieties of that variety. More specifically, the conjecture says
that certain de Rham cohomology classes are algebraic, that is, they are sums of Poincaré duals of the homology
classes of subvarieties. The Hodge conjecture is one of the Clay Mathematics Institute's Millennium Prize Problems.

Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

Motivation
Let X be a compact complex manifold of complex dimension n. Then X is an orientable smooth manifold of real
dimension 2n, so its cohomology groups lie in degrees zero through 2n. X is a Kähler manifold, so that there is a
decomposition on its cohomology with complex coefficients:

where is the subgroup of cohomology classes which are represented by harmonic forms of type (p, q).
That is, these are the cohomology classes represented by differential forms which, in some choice of local
coordinates , can be written as a harmonic function times .
(See Hodge theory for more details.) Taking wedge products of these harmonic representatives corresponds to the
cup product in cohomology, so the cup product is compatible with the Hodge decomposition:

Since X is a compact oriented manifold, X has a fundamental class.


Let Z be a complex submanifold of X of dimension k, and let i : Z → X be the inclusion map. Choose a differential
form of type (p, q). We can integrate over Z:

To evaluate this integral, choose a point of Z and call it 0. Around 0, we can choose local coordinates
on X such that Z is just . If p > k, then must contain some where pulls back to
zero on Z. The same is true if q > k. Consequently, this integral is zero if (p, q) ≠ (k, k).
More abstractly, the integral can be written as the cap product of the homology class of Z and the cohomology class
represented by . By Poincaré duality, the homology class of Z is dual to a cohomology class which we will call
Hodge conjecture 18

[Z], and the cap product can be computed by taking the cup product of [Z] and and capping with the fundamental
class of X. Because [Z] is a cohomology class, it has a Hodge decomposition. By the computation we did above, if
we cup this class with any class of type (p, q) ≠ (k, k), then we get zero. Because , we conclude that [Z]
Loosely speaking, the Hodge conjecture asks:
Which cohomology classes in come from complex subvarieties Z?

Statement of the Hodge conjecture


Let:

We call this the group of Hodge classes of degree 2k on X.


The modern statement of the Hodge conjecture is:
Hodge conjecture. Let X be a projective complex manifold. Then every Hodge class on X is a linear
combination with rational coefficients of the cohomology classes of complex subvarieties of X.
A projective complex manifold is a complex manifold which can be embedded in complex projective space. Because
projective space carries a Kähler metric, the Fubini-Study metric, such a manifold is always a Kähler manifold. By
Chow's theorem, a projective complex manifold is also a smooth projective algebraic variety, that is, it is the zero set
of a collection of homogenous polynomials.

Reformulation in terms of algebraic cycles


Another way of phrasing the Hodge conjecture involves the idea of an algebraic cycle. An algebraic cycle on X is a
formal combination of subvarieties of X, that is, it is something of the form:

The coefficients are usually taken to be integral or rational. We define the cohomology class of an algebraic cycle to
be the sum of the cohomology classes of its components. This is an example of the cycle class map of de Rham
cohomology, see Weil cohomology. For example, the cohomology class of the above cycle would be:

Such a cohomology class is called algebraic. With this notation, the Hodge conjecture becomes:
Let X be a projective complex manifold. Then every Hodge class on X is algebraic.

Known cases of the Hodge conjecture

Low dimension and codimension


The first result on the Hodge conjecture is due to Solomon Lefschetz. In fact, it predates the conjecture and provided
some of Hodge's motivation.
Theorem (Lefschetz theorem on (1,1)-classes) Any element of is the cohomology
class of a divisor on X. In particular, the Hodge conjecture is true for .
A very quick proof can be given using sheaf cohomology and the exponential exact sequence. (The cohomology
class of a divisor turns out to equal to its first Chern class.) Lefschetz's original proof proceeded by normal functions,
which were introduced by Henri Poincaré. However, Griffiths's transversality theorem shows that this approach
cannot prove the Hodge conjecture for higher codimensional subvarieties.
By the Hard Lefschetz theorem, one can prove:
Hodge conjecture 19

Theorem. If the Hodge conjecture holds for Hodge classes of degree p, p < n, then the Hodge conjecture holds
for Hodge classes of degree 2n − p.
Combining the above two theorems implies that Hodge conjecture is true for Hodge classes of degree 2n − 2. This
proves the Hodge conjecture when X has dimension at most three.
The Lefschetz theorem on (1,1)-classes also implies that if all Hodge classes are generated by the Hodge classes of
divisors, then the Hodge conjecture is true:
Corollary. If the algebra

is generated by Hdg1(X), then the Hodge conjecture holds for X.

Abelian varieties
For most abelian varieties, the algebra is generated in degree one, so the Hodge conjecture holds. In
particular, the Hodge conjecture holds for sufficiently general abelian varieties, for products of elliptic curves, and
for simple abelian varieties. However, David Mumford constructed an example of an abelian variety where
is not generated by products of divisor classes. André Weil generalized this example by showing that
whenever the variety has complex multiplication by an imaginary quadratic field, then is not generated
by products of divisor classes. Moonen and Zarhin proved that in dimension less than 5, either is
generated in degree one, or the variety has complex multiplication by an imaginary quadratic field. In the latter case,
the Hodge conjecture is only known in special cases.

Generalizations

The integral Hodge conjecture


Hodge's original conjecture was:
Integral Hodge conjecture. Let X be a projective complex manifold. Then every cohomology class in
is the cohomology class of an algebraic cycle with integral coefficients on X.
This is now known to be false. The first counterexample was constructed by Michael Atiyah and Friedrich
Hirzebruch. Using K-theory, they constructed an example of a torsion Hodge class, that is, a Hodge class such
that for some positive integer n, . Such a cohomology class cannot be the class of a cycle. Burt Totaro
reinterpreted their result in the framework of cobordism and found many examples of torsion classes.
The simplest adjustment of the integral Hodge conjecture is:
Integral Hodge conjecture modulo torsion. Let X be a projective complex manifold. Then every non-torsion
cohomology class in is the cohomology class of an algebraic cycle with integral
coefficients on X.
This is also false. János Kollár found an example of a Hodge class which is not algebraic, but which has an
integral multiple which is algebraic.
Hodge conjecture 20

The Hodge conjecture for Kähler varieties


A natural generalization of the Hodge conjecture would ask:
Hodge conjecture for Kähler varieties, naive version. Let X be a complex Kähler manifold. Then every
Hodge class on X is a linear combination with rational coefficients of the cohomology classes of complex
subvarieties of X.
This is too optimistic, because there are not enough subvarieties to make this work. A possible substitute is to ask
instead one of the two following questions:
Hodge conjecture for Kähler varieties, vector bundle version. Let X be a complex Kähler manifold. Then
every Hodge class on X is a linear combination with rational coefficients of Chern classes of vector bundles on
X.
Hodge conjecture for Kähler varieties, coherent sheaf version. Let X be a complex Kähler manifold. Then
every Hodge class on X is a linear combination with rational coefficients of Chern classes of coherent sheaves
on X.
Claire Voisin proved that the Chern classes of coherent sheaves give strictly more Hodge classes than the Chern
classes of vector bundles and that the Chern classes of coherent sheaves are insufficient to generate all the Hodge
classes. Consequently, the only known formulations of the Hodge conjecture for Kähler varieties are false.

The generalized Hodge conjecture


Hodge made an additional, stronger conjecture than the integral Hodge conjecture. Say that a cohomology class on X
is of level c if it is the pushforward of a cohomology class on a c-codimensional subvariety of X. The cohomology
classes of level at least c filter the cohomology of X, and it is easy to see that the cth step of the filtration
satisfies

Hodge's original statement was:


Generalized Hodge conjecture, Hodge's version.

Grothendieck observed that this cannot be true, even with rational coefficients, because the right-hand side is not
always a Hodge structure. His corrected form of the Hodge conjecture is:
Generalized Hodge conjecture. is the largest sub-Hodge structure of contained
in
This version is open.

Algebraicity of Hodge loci


The strongest evidence in favor of the Hodge conjecture is the algebraicity result of Cattani, Deligne and Kaplan.
Suppose that we vary the complex structure of X over a simply connected base. Then the topological cohomology of
X does not change, but the Hodge decomposition does change. It is known that if the Hodge conjecture is true, then
the locus of all points on the base where the cohomology of a fiber is a Hodge class is in fact an algebraic subset, that
is, it is cut out by polynomial equations. Cattani, Deligne, and Kaplan proved that this is always true, without
assuming the Hodge conjecture.
Hodge conjecture 21

References
• Cattani, Eduardo; Deligne, Pierre; Kaplan, Aroldo (1995), "On the locus of Hodge classes" [1], Journal of the
American Mathematical Society (American Mathematical Society) 8 (2): 483–506, doi:10.2307/2152824,
ISSN 0894-0347, MR1273413
• Hodge, W. V. D. (1950), "The topological invariants of algebraic varieties", Proceedings of the International
Congress of Mathematicians (Cambridge, MA) 1: 181–192.
• Grothendieck, A (1969), "Hodge's general conjecture is false for trivial reasons", Topology 8: 299–303,
doi:10.1016/0040-9383(69)90016-0.

External links
• The Clay Math Institute Official Problem Description (pdf) [2]
• Popular lecture on Hodge Conjecture by Dan Freed (University of Texas) (Real Video) [3] (Slides) [4]
• Indranil Biswas; Kapil Paranjape. The Hodge Conjecture for general Prym varieties [5]

References
[1] http:/ / jstor. org/ stable/ 2152824
[2] http:/ / www. claymath. org/ millennium/ Hodge_Conjecture/ Official_Problem_Description. pdf
[3] http:/ / claymath. msri. org/ hodgeconjecture. mov
[4] http:/ / www. ma. utexas. edu/ users/ dafr/ HodgeConjecture/ netscape_noframes. html
[5] http:/ / arxiv. org/ abs/ math/ 0007192v1

Riemann hypothesis

The real part (red) and imaginary part (blue) of


the Riemann zeta function along the critical line
Re(s) = 1/2. The first non-trivial zeros can be
seen at Im(s) = ±14.135, ±21.022 and ±25.011.
Riemann hypothesis 22

Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

In mathematics, the Riemann hypothesis, proposed by Bernhard Riemann (1859), is a conjecture about the
distribution of the zeros of the Riemann zeta function which states that all non-trivial zeros have real part 1/2. The
name is also used for some closely related analogues, such as the Riemann hypothesis for curves over finite fields.
The Riemann hypothesis implies results about the distribution of prime numbers that are in some ways as good as
possible. Along with suitable generalizations, it is considered by some mathematicians to be the most important
unresolved problem in pure mathematics (Bombieri 2000). The Riemann hypothesis is part of Problem 8, along with
the Goldbach conjecture, in Hilbert's list of 23 unsolved problems, and is also one of the Clay Mathematics Institute
Millennium Prize Problems. Since it was formulated, it has withstood concentrated efforts from many outstanding
mathematicians. In 1973, Pierre Deligne proved an analogue of the Riemann Hypothesis for zeta functions of
varieties defined over finite fields. The full version of the hypothesis remains unsolved, although modern computer
calculations have shown that the first 10 trillion zeros lie on the critical line.
The Riemann zeta function ζ(s) is defined for all complex numbers s ≠ 1. It has zeros at the negative even integers
(i.e. at s = −2, −4, −6, ...). These are called the trivial zeros. The Riemann hypothesis is concerned with the
non-trivial zeros, and states that:
The real part of any non-trivial zero of the Riemann zeta function is 1/2.
Thus the non-trivial zeros should lie on the critical line, 1/2 + it, where t is a real number and i is the imaginary unit.
There are several popular books on the Riemann hypothesis, such as Derbyshire (2003), Rockmore (2005), Sabbagh
(2003), du Sautoy (2003). The books Edwards (1974), Patterson (1988) and Borwein et al. (2008) give mathematical
introductions, while Titchmarsh (1986), Ivić (1985) and Karatsuba & Voronin (1992) are advanced monographs.

The Riemann zeta function


The Riemann zeta function is defined for complex s with real part greater than 1 by the absolutely convergent
infinite series

Leonhard Euler showed that this series equals the Euler product

where the infinite product extends over all prime numbers p, and again converges for complex s with real part
greater than 1. The convergence of the Euler product shows that ζ(s) has no zeros in this region, as none of the
factors have zeros.
The Riemann hypothesis discusses zeros outside the region of convergence of this series, so it needs to be
analytically continued to all complex s. This can be done by expressing it in terms of the Dirichlet eta function as
follows. If s is greater than one, then the zeta function satisfies
Riemann hypothesis 23

However, the series on the right converges not just when s is greater than one, but more generally whenever s has
positive real part. Thus, this alternative series extends the zeta function from Re(s) > 1 to the larger domain Re(s) >
0.
In the strip 0 < Re(s) < 1 the zeta function also satisfies the functional equation

One may then define ζ(s) for all remaining nonzero complex numbers s by assuming that this equation holds outside
the strip as well, and letting ζ(s) equal the right-hand side of the equation whenever s has non-positive real part. If s
is a negative even integer then ζ(s) = 0 because the factor sin(πs/2) vanishes; these are the trivial zeros of the zeta
function. (If s is a positive even integer this argument does not apply because the zeros of sin are cancelled by the
poles of the gamma function.) The value ζ(0) = −1/2 is not determined by the functional equation, but is the limiting
value of ζ(s) as s approaches zero. The functional equation also implies that the zeta function has no zeros with
negative real part other than the trivial zeros, so all non-trivial zeros lie in the critical strip where s has real part
between 0 and 1.

History
"…es ist sehr wahrscheinlich, dass alle Wurzeln reell sind. Hiervon wäre allerdings ein strenger Beweis zu wünschen; ich habe
indess die Aufsuchung desselben nach einigen flüchtigen vergeblichen Versuchen vorläufig bei Seite gelassen, da er für den
nächsten Zweck meiner Untersuchung entbehrlich schien."
"…it is very probable that all roots are real. Of course one would wish for a rigorous proof here; I have for the time being, after some
fleeting vain attempts, provisionally put aside the search for this, as it appears dispensable for the next objective of my
investigation."

Riemann's statement of the Riemann hypothesis, from (Riemann 1859). (He was discussing a version of the zeta function, modified
so that its roots are real rather than on the critical line.)

In his 1859 paper On the Number of Primes Less Than a Given Magnitude Riemann found an explicit formula for the
number of primes π(x) less than a given number x. His formula was given in terms of the related function

which counts primes where a prime power pn counts as 1/n of a prime. The number of primes can be recovered from
this function by

where μ is the Möbius function. Riemann's formula is then

where the sum is over the nontrivial zeros of the zeta function and where Π0 is a slightly modified version of Π that
replaces its value at its points of discontinuity by the average of its upper and lower limits:

The summation in Riemann's formula is not absolutely convergent, but may be evaluated by taking the zeros ρ in
order of the absolute value of their imaginary part. The function Li occurring in the first term is the (unoffset)
logarithmic integral function given by the Cauchy principal value of the divergent integral
Riemann hypothesis 24

The terms Li(xρ) involving the zeros of the zeta function need some care in their definition as Li has branch points at
0 and 1, and are defined (for x > 1) by analytic continuation in the complex variable ρ in the region Re(ρ) > 0, i.e.
they should be considered as Ei(ρ ln x). The other terms also correspond to zeros: the dominant term Li(x) comes
from the pole at s = 1, considered as a zero of multiplicity −1, and the remaining small terms come from the trivial
zeros. For some graphs of the sums of the first few terms of this series see Riesel & Göhl (1970) or Zagier (1977).
This formula says that the zeros of the Riemann zeta function control the oscillations of primes around their
"expected" positions. Riemann knew that the non-trivial zeros of the zeta function were symmetrically distributed
about the line s = 1/2 + it, and he knew that all of its non-trivial zeros must lie in the range 0 ≤ Re(s) ≤ 1. He checked
that a few of the zeros lay on the critical line with real part 1/2 and suggested that they all do; this is the Riemann
hypothesis.

Consequences of the Riemann hypothesis


The practical uses of the Riemann hypothesis include many propositions which are known to be true under the
Riemann hypothesis, and some which can be shown to be equivalent to the Riemann hypothesis.

Distribution of prime numbers


Riemann's explicit formula for the number of primes less than a given number in terms of a sum over the zeros of the
Riemann zeta function says that the magnitude of the oscillations of primes around their expected position is
controlled by the real parts of the zeros of the zeta function. In particular the error term in the prime number theorem
is closely related to the position of the zeros: for example, the supremum of real parts of the zeros is the infimum of
numbers β such that the error is O(xβ) (Ingham 1932).
Von Koch (1901) proved that the Riemann hypothesis is equivalent to the "best possible" bound for the error of the
prime number theorem.
A precise version of Koch's result, due to Schoenfeld (1976), says that the Riemann hypothesis is equivalent to

Growth of arithmetic functions


The Riemann hypothesis implies strong bounds on the growth of many other arithmetic functions, in addition to the
primes counting function above.
One example involves the Möbius function μ. The statement that the equation

is valid for every s with real part greater than 1/2, with the sum on the right hand side converging, is equivalent to the
Riemann hypothesis. From this we can also conclude that if the Mertens function is defined by

then the claim that

for every positive ε is equivalent to the Riemann hypothesis (Titchmarsh 1986). (For the meaning of these symbols,
see Big O notation.) The determinant of the order n Redheffer matrix is equal to M(n), so the Riemann hypothesis
can also be stated as a condition on the growth of these determinants. The Riemann hypothesis puts a rather tight
Riemann hypothesis 25

bound on the growth of M, since Odlyzko & te Riele (1985) disproved the slightly stronger Mertens conjecture

The Riemann hypothesis is equivalent to many other conjectures about the rate of growth of other arithmetic
functions aside from μ(n). A typical example is Robin's theorem (Robin 1984), which states that if σ(n) is the divisor
function, given by

then

for all n > 5040 if and only if the Riemann hypothesis is true, where γ is the Euler–Mascheroni constant.
Another example was found by Franel & Landau (1924) showing that the Riemann hypothesis is equivalent to a
statement that the terms of the Farey sequence are fairly regular. More precisely, if Fn is the Farey sequence of order
n, beginning with 1/n and up to 1/1, then the claim that for all ε > 0

is equivalent to the Riemann hypothesis. Here is the number of terms in the Farey sequence of order

n.
For an example from group theory, if g(n) is Landau's function given by the maximal order of elements of the
symmetric group Sn of degree n, then Massias, Nicolas & Robin (1988) showed that the Riemann hypothesis is
equivalent to the bound

for all sufficiently large n.

Lindelöf hypothesis and growth of the zeta function


The Riemann hypothesis has various weaker consequences as well; one is the Lindelöf hypothesis on the rate of
growth of the zeta function on the critical line, which says that, for any ε > 0,

as t tends to infinity.
The Riemann hypothesis also implies quite sharp bounds for the growth rate of the zeta function in other regions of
the critical strip. For example, it implies that

so the growth rate of ζ(1+it) and its inverse would be known up to a factor of 2 (Titchmarsh 1986).

Large prime gap conjecture


The prime number theorem implies that on average, the gap between the prime p and its successor is log p. However,
some gaps between primes may be much larger than the average. Cramér proved that, assuming the Riemann
hypothesis, every gap is O(√p log p). This is a case when even the best bound that can currently be proved using the
Riemann hypothesis is far weaker than what seems to be true: Cramér's conjecture implies that every gap is
O((log p)2) which, while larger than the average gap, is far smaller than the bound implied by the Riemann
Riemann hypothesis 26

hypothesis. Numerical evidence supports Cramér's conjecture (Nicely 1999).

Criteria equivalent to the Riemann hypothesis


Many statements equivalent to the Riemann hypothesis have been found, though so far none of them have led to
much progress in solving it. Some typical examples are as follows.
The Riesz criterion was given by Riesz (1916), to the effect that the bound

holds for all if and only if the Riemann hypothesis holds.


Nyman (1950) proved that the Riemann Hypothesis is true if and only if the space of functions of the form

where ρ(z) is the fractional part of z, 0 ≤ θν ≤ 1, and

is dense in the Hilbert space L2(0,1) of square-integrable functions on the unit interval. Beurling (1955) extended this
by showing that the zeta function has no zeros with real part greater than 1/p if and only if this function space is
dense in Lp(0,1)
Salem (1953) showed that the Riemann hypothesis is true if and only if the integral equation

has no non-trivial bounded solutions φ for 1/2<σ<1.


Weil's criterion is the statement that the positivity of a certain function is equivalent to the Riemann hypothesis.
Related is Li's criterion, a statement that the positivity of a certain sequence of numbers is equivalent to the Riemann
hypothesis.
Speiser (1934) proved that the Riemann hypothesis is equivalent to the statement that , the derivative of
, has no zeros in the strip

That ζ has only simple zeros on the critical line is equivalent to its derivative having no zeros on the critical line.

Consequences of the generalized Riemann hypothesis


Several applications use the generalized Riemann hypothesis for Dirichlet L-series or zeta functions of number fields
rather than just the Riemann hypothesis. Many basic properties of the Riemann zeta function can easily be
generalized to all Dirichlet L-series, so it is plausible that a method that proves the Riemann hypothesis for the
Riemann zeta function would also work for the generalized Riemann hypothesis for Dirichlet L-functions. Several
results first proved using the generalized Riemann hypothesis were later given unconditional proofs without using it,
though these were usually much harder. Many of the consequences on the following list are taken from Conrad
(2010).
• In 1913, Gronwall showed that the generalized Riemann hypothesis implies that Gauss's list of imaginary
quadratic fields with class number 1 is complete, though Baker, Stark and Heegner later gave unconditional
proofs of this without using the generalized Riemann hypothesis.
• In 1917, Hardy and Littlewood showed that the generalized Riemann hypothesis implies a conjecture of
Chebyshev that
Riemann hypothesis 27

which says that in some sense primes 3 mod 4 are more common than primes 1 mod 4.
• In 1923 Hardy and Littlewood showed that the generalized Riemann hypothesis implies a weak form of the
Goldbach conjecture for odd numbers: that every sufficiently large odd number is the sum of 3 primes, though in
1937 Vinogradov gave an unconditional proof. In 1997 Deshouillers, Effinger, te Riele, and Zinoviev showed that
the generalized Riemann hypothesis implies that every odd number greater than 5 is the sum of 3 primes.
• In 1934, Chowla showed that the generalized Riemann hypothesis implies that the first prime in the arithmetic
progression a mod m is at most Km2log(m)2 for some fixed constant K.
• In 1967, Hooley showed that the generalized Riemann hypothesis implies Artin's conjecture on primitive roots.
• In 1973, Weinberger showed that the generalized Riemann hypothesis implies that Euler's list of idoneal numbers
is complete.
• Weinberger (1973) showed that the generalized Riemann hypothesis for the zeta functions of all algebraic number
fields implies that any number field with class number 1 is either Euclidean or an imaginary quadratic number
field of discriminant −19, −43, −67, or −163.
• In 1976, G. Miller showed that the generalized Riemann hypothesis implies that one can test if a number is prime
in polynomial times. In 2002, Manindra Agrawal, Neeraj Kayal and Nitin Saxena proved this result
unconditionally using the AKS primality test.
• Odlyzko (1990) discussed how the generalized Riemann hypothesis can be used to give sharper estimates for
discriminants and class numbers of number fields.
• Ono & Soundararajan (1997) showed that the generalized Riemann hypothesis implies that Ramanujan's integral
quadratic form x2 +y2 + 10z2 represents all integers that it represents locally, with exactly 18 exceptions.

Generalizations and analogues of the Riemann hypothesis

Dirichlet L-series and other number fields


The Riemann hypothesis can be generalized by replacing the Riemann zeta function by the formally similar, but
much more general, global L-functions. In this broader setting, one expects the non-trivial zeros of the global
L-functions to have real part 1/2. It is these conjectures, rather than the classical Riemann hypothesis only for the
single Riemann zeta function, which accounts for the true importance of the Riemann hypothesis in mathematics.
The generalized Riemann hypothesis extends the Riemann hypothesis to all Dirichlet L-functions. In particular it
implies the conjecture that Siegel zeros (zeros of L functions between 1/2 and 1) do not exist.
The extended Riemann hypothesis extends the Riemann hypothesis to all Dedekind zeta functions of algebraic
number fields. The extended Riemann hypothesis for abelian extension of the rationals is equivalent to the
generalized Riemann hypothesis. The Riemann hypothesis can also be extended to the L-functions of Hecke
characters of number fields.
The grand Riemann hypothesis extends it to all automorphic zeta functions, such as Mellin transforms of Hecke
eigenforms.
Riemann hypothesis 28

Function fields and zeta functions of varieties over finite fields


Artin (1924) introduced global zeta functions of (quadratic) function fields and conjectured an analogue of the
Riemann hypothesis for them, which has been proven by Hasse in the genus 1 case and by Weil (1948) in general.
For instance, the fact that the Gauss sum, of the quadratic character of a finite field of size q (with q odd), has
absolute value

is actually an instance of the Riemann hypothesis in the function field setting. This led Weil (1949) to conjecture a
similar statement for all algebraic varieties; the resulting Weil conjectures were proven by Pierre Deligne (1974,
1980).

Selberg zeta functions


Selberg (1956) introduced the Selberg zeta function of a Riemann surface. These are similar to the Riemann zeta
function: they have a functional equation, and an infinite product similar to the Euler product but taken over closed
geodesics rather than primes. The Selberg trace formula is the analogue for these functions of the explicit formulas in
prime number theory. Selberg proved that the Selberg zeta functions satisfy the analogue of the Riemann hypothesis,
with the imaginary parts of their zeros related to the eigenvalues of the Laplacian operator of the Riemann surface.

Ihara zeta functions


The Ihara zeta function of a finite graph is an analogue of the Selberg zeta function introduced by Yasutaka Ihara. A
regular finite graph is a Ramanujan graph, a mathematical model of efficient communication networks, if and only if
its Ihara zeta function satisfies the analogue of the Riemann hypothesis as was pointed out by T. Sunada.

Montgomery's pair correlation conjecture


Montgomery (1973) suggested the pair correlation conjecture that the correlation functions of the (suitably
normalized) zeros of the zeta function should be the same as those of the eigenvalues of a random hermitian matrix.
Odlyzko (1987) showed that this is supported by large scale numerical calculations of these correlation functions.
Montgomery showed that (assuming the Riemann hypothesis) at least 2/3 of all zeros are simple, and a related
conjecture is that all zeros of the zeta function are simple (or more generally have no non-trivial integer linear
relations between their imaginary parts). Dedekind zeta functions of algebraic number fields, which generalize the
Riemann zeta function, often do have multiple complex zeros. This is because the Dedekind zeta functions factorize
as a product of powers of Artin L-functions, so zeros of Artin L-functions sometimes give rise to multiple zeros of
Dedekind zeta functions. Other examples of zeta functions with multiple zeros are the L-functions of some elliptic
curves: these can have multiple zeros at the real point of their critical line; the Birch-Swinnerton-Dyer conjecture
predicts that the multiplicity of this zero is the rank of the elliptic curve.

Other zeta functions


There are many other examples of zeta functions with analogues of the Riemann hypothesis, some of which have
been proved. Goss zeta functions of function fields have a Riemann hypothesis, proved by Sheats (1998). The main
conjecture of Iwasawa theory, proved by Barry Mazur and Andrew Wiles for cyclotomic fields, and Wiles for totally
real fields, identifies the zeros of a p-adic L-function with the eigenvalues of an operator, so can be thought of as an
analogue of the Hilbert–Pólya conjecture for p-adic L-functions (Wiles 2000).
Riemann hypothesis 29

Attempts to prove the Riemann hypothesis


Several mathematicians have addressed the Riemann hypothesis, but none of their attempts have yet been accepted
as correct solutions. Watkins (2007) lists some incorrect solutions, and more are frequently announced [1].

Operator theory
Hilbert and Polya suggested that one way to derive the Riemann hypothesis would be to find a self-adjoint operator,
from the existence of which the statement on the real parts of the zeros of ζ(s) would follow when one applies the
criterion on real eigenvalues. Some support for this idea comes from several analogues of the Riemann zeta
functions whose zeros correspond to eigenvalues of some operator: the zeros of a zeta function of a variety over a
finite field correspond to eigenvalues of a Frobenius element on an etale cohomology group, the zeros of a Selberg
zeta function are eigenvalues of a Laplacian operator of a Riemann surface, and the zeros of a p-adic zeta function
correspond to eigenvectors of a Galois action on ideal class groups.
Odlyzko (1987) showed that the distribution of the zeros of the Riemann zeta function shares some statistical
properties with the eigenvalues of random matrices drawn from the Gaussian unitary ensemble. This gives some
support to the Hilbert–Pólya conjecture.
In 1999, Michael Berry and Jon Keating conjectured that there is some unknown quantization of the classical
Hamiltonian so that

and even more strongly, that the Riemann zeros coincide with the spectrum of the operator . This is to
be contrasted to canonical quantization which leads to the Heisenberg uncertainty principle and the
natural numbers as spectrum of the quantum harmonic oscillator. The crucial point is that the Hamiltonian should be
a self-adjoint operator so that the quantization would be a realization of the Hilbert–Pólya program. In a connection
with this quantum mechanical problem Berry and Connes had proposed that the inverse of the potential of the
Hamiltonian is connected to the half-derivative of the function then, in

Berry-Connes approach (Connes 1999). This yields to a Hamiltonian whose

eigenvalues are the square of the imaginary part of the Riemann zeros, also the functional determinant of this
Hamiltonian operator is just the Riemann Xi-function
The analogy with the Riemann hypothesis over finite fields suggests that the Hilbert space containing eigenvectors
corresponding to the zeros might be some sort of first cohomology group of the spectrum Spec(Z) of the integers.
Deninger (1998) described some of the attempts to find such a cohomology theory.
Zagier (1983) constructed a natural space of invariant functions on the upper half plane which has eigenvalues under
the Laplacian operator corresponding to zeros of the Riemann zeta function, and remarked that in the unlikely event
that one could show the existence of a suitable positive definite inner product on this space the Riemann hypothesis
would follow. Cartier (1982) discussed a related example, where due to a bizarre bug a computer program listed
zeros of the Riemann zeta function as eigenvalues of the same Laplacian operator.
Schumayer & Hutchinson (2011) surveyed some of the attempts to construct a suitable physical model related to the
Riemann zeta function.
Riemann hypothesis 30

Lee–Yang theorem
The Lee–Yang theorem states that the zeros of certain partition functions in statistical mechanics all lie on a "critical
line" with real part 0, and this has led to some speculation about a relationship with the Riemann hypothesis (Knauf
1999).

Turán's result
Pál Turán (1948) showed that if the functions

have no zeros when the real part of s is greater than one then

for all x > 0,

where λ(n) is the Liouville function given by (−1)r if n has r prime factors. He showed that this in turn would imply
that the Riemann hypothesis is true. However Haselgrove (1958) proved that T(x) is negative for infinitely many x
(and also disproved the closely related Polya conjecture), and Borwein, Ferguson & Mossinghoff (2008) showed that
the smallest such x is 72185376951205. Spira (1968) showed by numerical calculation that the finite Dirichlet series
above for N=19 has a zero with real part greater than 1. Turán also showed that a somewhat weaker assumption, the
nonexistence of zeros with real part greater than 1+N−1/2+ε for large N in the finite Dirichlet series above, would also
imply the Riemann hypothesis, but Montgomery (1983) showed that for all sufficiently large N these series have
zeros with real part greater than 1 + (log log N)/(4 log N). Therefore, Turán's result is vacuously true and cannot be
used to help prove the Riemann hypothesis.

Noncommutative geometry
Connes (1999, 2000) has described a relationship between the Riemann hypothesis and noncommutative geometry,
and shows that a suitable analogue of the Selberg trace formula for the action of the idèle class group on the adèle
class space would imply the Riemann hypothesis. Some of these ideas are elaborated in Lapidus (2008).

Hilbert spaces of entire functions


Louis de Branges (1992) showed that the Riemann hypothesis would follow from a positivity condition on a certain
Hilbert space of entire functions. However Conrey & Li (2000) showed that the necessary positivity conditions are
not satisfied.

Quasicrystals
The Riemann hypothesis implies that the zeros of the zeta function form a quasicrystal, meaning a distribution with
discrete support whose Fourier transform also has discrete support. Dyson (2009) suggested trying to prove the
Riemann hypothesis by classifying, or at least studying, 1-dimensional quasicrystals.

Multiple zeta functions


Deligne's proof of the Riemann hypothesis over finite fields used the zeta functions of product varieties, whose zeros
and poles correspond to sums of zeros and poles of the original zeta function, in order to bound the real parts of the
zeros of the original zeta function. By analogy, Kurokawa (1992) introduced multiple zeta functions whose zeros and
poles correspond to sums of zeros and poles of the Riemann zeta function. To make the series converge he restricted
to sums of zeros or poles all with non-negative imaginary part. So far, the known bounds on the zeros and poles of
the multiple zeta functions are not strong enough to give useful estimates for the zeros of the Riemann zeta function.
Riemann hypothesis 31

Location of the zeros

Number of zeros
The functional equation combined with the argument principle implies that the number of zeros of the zeta function
with imaginary part between 0 and T is given by

for s=1/2+iT, where the argument is defined by varying it continuously along the line with Im(s)=T, starting with
argument 0 at ∞+iT. This is the sum of a large but well understood term

and a small but rather mysterious term

So the density of zeros with imaginary part near T is about log(T)/2π, and the function S describes the small
deviations from this. The function S(t) jumps by 1 at each zero of the zeta function, and for t ≥ 8 it decreases
monotonically between zeros with derivative close to −log t.
Karatsuba (1996) proved that every interval for contains at least

points where the function changes sign.


Selberg (1946) showed that the average moments of even powers of S are given by

This suggests that S(T)/(log log T)1/2 resembles a Gaussian random variable with mean 0 and variance 2π2 (Ghosh
(1983) proved this fact). In particular |S(T)| is usually somewhere around (log log T)1/2, but occasionally much
larger. The exact order of growth of S(T) is not known. There has been no unconditional improvement to Riemann's
original bound S(T)=O(log T), though the Riemann hypothesis implies the slightly smaller bound S(T)=O(log T/log
log T) (Titchmarsh 1985). The true order of magnitude may be somewhat less than this, as random functions with the
same distribution as S(T) tend to have growth of order about log(T)1/2. In the other direction it cannot be too small:
Selberg (1946) showed that S(T) ≠ o((log T)1/3/(log log T)7/3), and assuming the Riemann hypothesis Montgomery
showed that S(T) ≠ o((log T)1/2/(log log T)1/2).
Numerical calculations confirm that S grows very slowly: |S(T)| < 1 for T < 280, |S(T)| < 2 for T < 6800000, and the
largest value of |S(T)| found so far is not much larger than 3 (Odlyzko 2002).
Riemann's estimate S(T) = O(log T) implies that the gaps between zeros are bounded, and Littlewood improved this
slightly, showing that the gaps between their imaginary parts tends to 0.
Riemann hypothesis 32

The theorem of Hadamard and de la Vallée-Poussin


Hadamard (1896) and de la Vallée-Poussin (1896) independently proved that no zeros could lie on the line Re(s) = 1.
Together with the functional equation and the fact that there are no zeros with real part greater than 1, this showed
that all non-trivial zeros must lie in the interior of the critical strip 0 < Re(s) < 1. This was a key step in their first
proofs of the prime number theorem.
Both the original proofs that the zeta function has no zeros with real part 1 are similar, and depend on showing that if
ζ(1+it) vanishes, then ζ(1+2it) is singular, which is not possible. One way of doing this is by using the inequality
for σ>1, t real,
and looking at the limit as σ tends to 1. This inequality follows by taking the real part of the log of the Euler product
to see that

(where the sum is over all prime powers pn) so that

which is at least 1 because all the terms in the sum are positive, due to the inequality

Zero-free regions
De la Vallée-Poussin (1899-1900) proved that if σ+it is a zero of the Riemann zeta function, then 1-σ ≥ C/log(t) for
some positive constant C. In other words zeros cannot be too close to the line σ=1: there is a zero-free region close to
this line. This zero-free region has been enlarged by several authors. Ford (2002) gave a version with explicit
numerical constants: ζ(σ + it) ≠ 0 whenever |t| ≥ 3 and

Zeros on the critical line


Hardy (1914) and Hardy & Littlewood (1921) showed there are infinitely many zeros on the critical line, by
considering moments of certain functions related to the zeta function. Selberg (1942) proved that at least a (small)
positive proportion of zeros lie on the line. Levinson (1974) improved this to one-third of the zeros by relating the
zeros of the zeta function to those of its derivative, and Conrey (1989) improved this further to two-fifths.
Most zeros lie close to the critical line. More precisely, Bohr & Landau (1914) showed that for any positive ε, all but
an infinitely small proportion of zeros lie within a distance ε of the critical line. Ivić (1985) gives several more
precise versions of this result, called zero density estimates, which bound the number of zeros in regions with
imaginary part at most T and real part at least 1/2+ε.
Riemann hypothesis 33

The Hardy-Littlewood conjectures


In 1914 Godfrey Harold Hardy proved that has infinitely many real zeros.
Let be the total number of real zeros, be the total number of zeros of odd order of the function
, lying on the interval .
The next two conjectures of Hardy and John Edensor Littlewood on the distance between real zeros of
and on the density of zeros of on intervals for sufficiently great ,
and with as less as possible value of , where is an arbitrarily small number, open two new directions
in the investigation of the Riemann zeta function:
1. for any there exists such that for and the interval
contains a zero of odd order of the function .
2. for any there exist and , such that for and the
inequality is true.

The Selberg conjecture


Atle Selberg (1942) investigated the problem of Hardy-Littlewood 2 and proved that for any there exists such
and , such that for and the inequality
is true. Selberg conjectured that this could be tightened to . A.
A. Karatsuba (1984a, 1984b, 1985) proved that for a fixed satisfying the condition , a
sufficiently large and , , the interval contains at least
real zeros of the Riemann zeta function and therefore confirmed the Selberg conjecture. The estimates

of Selberg and Karatsuba can not be improved in respect of the order of growth as .
Karatsuba (1992) proved that an analog of the Selberg conjecture holds for almost all intervals ,
, where is an arbitrarily small fixed positive number. The Karatsuba method permits to investigate
zeros of the Riemann zeta-function on "supershort" intervals of the critical line, that is, on the intervals
, the length of which grows slower than any, even arbitrarily small degree . In particular, he
proved that for any given numbers , satisfying the conditions almost all intervals
for contain at least zeros of the function . This
estimate is quite close to the one that follows from the Riemann hypothesis.
Riemann hypothesis 34

Numerical calculations
The function

Absolute value of the ζ-function

has the same zeros as the zeta function in the critical strip, and is real on the critical line because of the functional
equation, so one can prove the existence of zeros exactly on the real line between two points by checking
numerically that the function has opposite signs at these points. Usually one writes

where Hardy's function Z and the Riemann-Siegel theta function θ are uniquely defined by this and the condition that
they are smooth real functions with θ(0)=0. By finding many intervals where the function Z changes sign one can
show that there are many zeros on the critical line. To verify the Riemann hypothesis up to a given imaginary part T
of the zeros, one also has to check that there are no further zeros off the line in this region. This can be done by
calculating the total number of zeros in the region and checking that it is the same as the number of zeros found on
the line. This allows one to verify the Riemann hypothesis computationally up to any desired value of T (provided all
the zeros of the zeta function in this region are simple and on the critical line).
Some calculations of zeros of the zeta function are listed below. So far all zeros that have been checked are on the
critical line and are simple. (A multiple zero would cause problems for the zero finding algorithms, which depend on
finding sign changes between zeros.) For tables of the zeros, see Haselgrove & Miller (1960) or Odlyzko.

Year Number of zeros Author

1859? 3 B. Riemann used the Riemann-Siegel formula (unpublished, but reported in Siegel 1932).

1903 15 J. P. Gram (1903) used Euler–Maclaurin summation and discovered Gram's law. He showed that all 10 zeros
with imaginary part at most 50 range lie on the critical line with real part 1/2 by computing the sum of the
inverse 10th powers of the roots he found.

1914 79 (γn ≤ 200) R. J. Backlund (1914) introduced a better method of checking all the zeros up to that point are on the line, by
studying the argument S(T) of the zeta function.

1925 138 (γn ≤ 300) J. I. Hutchinson (1925) found the first failure of Gram's law, at the Gram point g126.

1935 195 E. C. Titchmarsh (1935) used the recently rediscovered Riemann-Siegel formula, which is much faster than
Euler–Maclaurin summation.It takes about O(T3/2+ε) steps to check zeros with imaginary part less than T, while
the Euler–Maclaurin method takes about O(T2+ε) steps.
Riemann hypothesis 35

1936 1041 E. C. Titchmarsh (1936) and L. J. Comrie were the last to find zeros by hand.

1953 1104 A. M. Turing (1953) found a more efficient way to check that all zeros up to some point are accounted for by the
zeros on the line, by checking that Z has the correct sign at several consecutive Gram points and using the fact
that S(T) has average value 0. This requires almost no extra work because the sign of Z at Gram points is already
known from finding the zeros, and is still the usual method used. This was the first use of a digital computer to
calculate the zeros.

1956 15000 D. H. Lehmer (1956) discovered a few cases where the zeta function has zeros that are "only just" on the line:
two zeros of the zeta function are so close together that it is unusually difficult to find a sign change between
them. This is called "Lehmer's phenomenon", and first occurs at the zeros with imaginary parts 7005.063 and
7005.101, which differ by only .04 while the average gap between other zeros near this point is about 1.

1956 25000 D. H. Lehmer

1958 35337 N. A. Meller

1966 250000 R. S. Lehman

1968 3500000 Rosser, Yohe & Schoenfeld (1969) stated Rosser's rule (described below).

1977 40000000 R. P. Brent

1979 81000001 R. P. Brent

1982 200000001 R. P. Brent, J. van de Lune, H. J. J. te Riele, D. T. Winter

1983 300000001 J. van de Lune, H. J. J. te Riele

1986 1500000001 van de Lune, te Riele & Winter (1986) gave some statistical data about the zeros and give several graphs of Z at
places where it has unusual behavior.

1987 A few of large (~1012) A. M. Odlyzko (1987) computed smaller numbers of zeros of much larger height, around 1012, to high precision
height to check Montgomery's pair correlation conjecture.

1992 A few of large (~1020) A. M. Odlyzko (1992) computed a 175 million zeroes of heights around 1020 and a few more of heights around
height 2×1020, and gave an extensive discussion of the results.

1998 10000 of large (~1021) A. M. Odlyzko (1998) computed some zeros of height about 1021
height

2001 10000000000 J. van de Lune (unpublished)

2004 900000000000 S. Wedeniwski (ZetaGrid distributed computing)

2004 10000000000000 and a X. Gourdon (2004) and Patrick Demichel used the Odlyzko–Schönhage algorithm. They also checked two
few of large (up to billion zeros around heights 1013, 1014, ... , 1024.
~1024) heights

Gram points
A Gram point is a value of t such that ζ(1/2 + it) = Z(t)e − iθ(t) is a non-zero real; these are easy to find because they
are the points where the Euler factor at infinity π−s/2Γ(s/2) is real at s = 1/2 + it, or equivalently θ(t) is a multiple nπ
of π. They are usually numbered as gn for n = −1, 0, 1, ..., where gn is the unique solution of θ(t) = nπ with t ≥ 8 (θ is
increasing beyond this point; there is a second point with θ(t) = −π near 3.4, and θ(0) = 0). Gram observed that there
was often exactly one zero of the zeta function between any two Gram points; Hutchinson called this observation
Gram's law. There are several other closely related statements that are also sometimes called Gram's law: for
example, (−1)nZ(gn) is usually positive, or Z(t) usually has opposite sign at consecutive Gram points. The imaginary
parts γn of the first few zeros (in blue) and the first few Gram points gn are given in the following table

g−1 γ1 g0 γ2 g1 γ3 g2 γ4 g3 γ5 g4 γ6 g5

0 3.4 9.667 14.135 17.846 21.022 23.170 25.011 27.670 30.425 31.718 32.935 35.467 37.586 38.999
Riemann hypothesis 36

The first failure of Gram's law occurs at the 127'th zero and the Gram
point g126, which are in the "wrong" order.

This shows the values of ζ(1/2+it) in the complex


plane for 0 ≤ t ≤ 34. (For t=0, ζ(1/2) ≈ -1.460
corresponds to the leftmost point of the red
curve.) Gram's law states that the curve usually
crosses the real axis once between zeros.

g124 γ126 g125 g126 γ127 γ128 g127 γ129 g128

279.148 279.229 280.802 282.455 282.465 283.211 284.104 284.836 285.752

A Gram point t is called good if the zeta function is positive at 1/2 + it. The indices of the "bad" Gram points where
Z has the "wrong" sign are 126, 134, 195, 211,... (sequence A114856 [2] in OEIS). A Gram block is an interval
bounded by two good Gram points such that all the Gram points between them are bad. A refinement of Gram's law
called Rosser's rule due to Rosser, Yohe & Schoenfeld (1969) says that Gram blocks often have the expected number
of zeros in them (the same as the number of Gram intervals), even though some of the individual Gram intervals in
the block may not have exactly one zero in them. For example, the interval bounded by g125 and g127 is a Gram
block containing a unique bad Gram point g126, and contains the expected number 2 of zeros although neither of its
two Gram intervals contains a unique zero. Rosser et al. checked that there were no exceptions to Rosser's rule in the
first 3 million zeros, although there are infinitely many exceptions to Rosser's rule over the entire zeta function.
Gram's rule and Rosser's rule both say that in some sense zeros do not stray too far from their expected positions.
The distance of a zero from its expected position is controlled by the function S defined above, which grows
extremely slowly: its average value is of the order of (log log T)1/2, which only reaches 2 for T around 1024. This
means that both rules hold most of the time for small T but eventually break down often.

Arguments for and against the Riemann hypothesis


Mathematical papers about the Riemann hypothesis tend to be cautiously noncommittal about its truth. Of authors
who express an opinion, most of them, such as Riemann (1859) or Bombieri (2000), imply that they expect (or at
least hope) that it is true. The few authors who express serious doubt about it include Ivić (2008) who lists some
reasons for being skeptical, and Littlewood (1962) who flatly states that he believes it to be false, and that there is no
evidence whatever for it and no imaginable reason for it to be true. The consensus of the survey articles (Bombieri
2000, Conrey 2003, and Sarnak 2008) is that the evidence for it is strong but not overwhelming, so that while it is
probably true there is some reasonable doubt about it.
Some of the arguments for (or against) the Riemann hypothesis are listed by Sarnak (2008), Conrey (2003), and Ivić
(2008), and include the following reasons.
• Several analogues of the Riemann hypothesis have already been proved. The proof of the Riemann hypothesis for
varieties over finite fields by Deligne (1974) is possibly the single strongest theoretical reason in favor of the
Riemann hypothesis. This provides some evidence for the more general conjecture that all zeta functions
Riemann hypothesis 37

associated with automorphic forms satisfy a Riemann hypothesis, which includes the classical Riemann
hypothesis as a special case. Similarly Selberg zeta functions satisfy the analogue of the Riemann hypothesis, and
are in some ways similar to the Riemann zeta function, having a functional equation and an infinite product
expansion analogous to the Euler product expansion. However there are also some major differences; for example
they are not given by Dirichlet series. The Riemann hypothesis for the Goss zeta function was proved by Sheats
(1998). In contrast to these positive examples, however, some Epstein zeta functions do not satisfy the Riemann
hypothesis, even though they have an infinite number of zeros on the critical line (Titchmarsh 1986). These
functions are quite similar to the Riemann zeta function, and have a Dirichlet series expansion and a functional
equation, but the ones known to fail the Riemann hypothesis do not have an Euler product and are not directly
related to automorphic representations.
• The numerical verification that many zeros lie on the line seems at first sight to be strong evidence for it.
However analytic number theory has had many conjectures supported by large amounts of numerical evidence
that turn out to be false. See Skewes number for a notorious example, where the first exception to a plausible
conjecture related to the Riemann hypothesis probably occurs around 10316; a counterexample to the Riemann
hypothesis with imaginary part this size would be far beyond anything that can currently be computed. The
problem is that the behavior is often influenced by very slowly increasing functions such as log log T, that tend to
infinity, but do so so slowly that this cannot be detected by computation. Such functions occur in the theory of the
zeta function controlling the behavior of its zeros; for example the function S(T) above has average size around
(log log T)1/2 . As S(T) jumps by at least 2 at any counterexample to the Riemann hypothesis, one might expect
any counterexamples to the Riemann hypothesis to start appearing only when S(T) becomes large. It is never
much more than 3 as far as it has been calculated, but is known to be unbounded, suggesting that calculations may
not have yet reached the region of typical behavior of the zeta function.
• Denjoy's probabilistic argument for the Riemann hypothesis (Edwards 1974) is based on the observation that If
μ(x) is a random sequence of "1"s and "−1"s then, for every ε > 0, the partial sums

(the values of which are positions in a simple random walk) satisfy the bound

with probability 1. The Riemann hypothesis is equivalent to this bound for the Möbius function μ and the
Mertens function M derived in the same way from it. In other words, the Riemann hypothesis is in some sense
equivalent to saying that μ(x) behaves like a random sequence of coin tosses. When μ(x) is non-zero its sign
gives the parity of the number of prime factors of x, so informally the Riemann hypothesis says that the parity
of the number of prime factors of an integer behaves randomly. Such probabilistic arguments in number theory
often give the right answer, but tend to be very hard to make rigorous, and occasionally give the wrong answer
for some results, such as Maier's theorem.
• The calculations in Odlyzko (1987) show that the zeros of the zeta function behave very much like the
eigenvalues of a random Hermitian matrix, suggesting that they are the eigenvalues of some self-adjoint operator,
which would imply the Riemann hypothesis. However all attempts to find such an operator have failed.
• There are several theorems, such as Goldbach's conjecture for sufficiently large odd numbers, that were first
proved using the generalized Riemann hypothesis, and later shown to be true unconditionally. This could be
considered as weak evidence for the generalized Riemann hypothesis, as several of its "predictions" turned out to
be true.
• Lehmer's phenomenon (Lehmer 1956) where two zeros are sometimes very close is sometimes given as a reason
to disbelieve in the Riemann hypothesis. However one would expect this to happen occasionally just by chance
even if the Riemann hypothesis were true, and Odlyzko's calculations suggest that nearby pairs of zeros occur just
as often as predicted by Montgomery's conjecture.
Riemann hypothesis 38

• Patterson (1988) suggests that the most compelling reason for the Riemann hypothesis for most mathematicians is
the hope that primes are distributed as regularly as possible.

References
• Artin, Emil (1924), "Quadratische Körper im Gebiete der höheren Kongruenzen. II. Analytischer Teil",
Mathematische Zeitschrift 19 (1): 207–246, doi:10.1007/BF01181075
• Beurling, Arne (1955), "A closure problem related to the Riemann zeta-function", Proceedings of the National
Academy of Sciences of the United States of America 41 (5): 312–314, doi:10.1073/pnas.41.5.312, MR0070655
• Bohr, H.; Landau, E. (1914), "Ein Satz über Dirichletsche Reihen mit Anwendung auf die ζ-Funktion und die
L-Funktionen", Rendiconti del Circolo Matematico di Palermo 37 (1): 269–272, doi:10.1007/BF03014823
• Bombieri, Enrico (2000) (PDF), The Riemann Hypothesis - official problem description [3], Clay Mathematics
Institute, retrieved 2008-10-25 Reprinted in (Borwein et al. 2008).
• Borwein, Peter; Choi, Stephen; Rooney, Brendan et al., eds. (2008), The Riemann Hypothesis: A Resource for the
Afficionado and Virtuoso Alike, CMS Books in Mathematics, New York: Springer,
doi:10.1007/978-0-387-72126-2, ISBN 978-0387721255
• Borwein, Peter; Ferguson, Ron; Mossinghoff, Michael J. (2008), "Sign changes in sums of the Liouville
function", Mathematics of Computation 77 (263): 1681–1694, doi:10.1090/S0025-5718-08-02036-X,
MR2398787
• de Branges, Louis (1992), "The convergence of Euler products", Journal of Functional Analysis 107 (1):
122–210, doi:10.1016/0022-1236(92)90103-P, MR1165869
• Cartier, P. (1982), "Comment l'hypothèse de Riemann ne fut pas prouvée", Seminar on Number Theory, Paris
1980-81 (Paris, 1980/1981), Progr. Math., 22, Boston, MA: Birkhäuser Boston, pp. 35–48, MR693308
• Connes, Alain (1999), "Trace formula in noncommutative geometry and the zeros of the Riemann zeta function",
Selecta Mathematica. New Series 5 (1): 29–106, arXiv:math/9811068, doi:10.1007/s000290050042, MR1694895
• Connes, Alain (2000), "Noncommutative geometry and the Riemann zeta function", Mathematics: frontiers and
perspectives, Providence, R.I.: American Mathematical Society, pp. 35–54, MR1754766
• Conrey, J. B. (1989), "More than two fifths of the zeros of the Riemann zeta function are on the critical line" [4],
J. Reine angew. Math. 399: 1–16, MR1004130
• Conrey, J. Brian (2003), "The Riemann Hypothesis" [5] (PDF), Notices of the American Mathematical Society:
341–353 Reprinted in (Borwein et al. 2008).
• Conrey, J. B.; Li, Xian-Jin (2000), "A note on some positivity conditions related to zeta and L-functions",
International Mathematics Research Notices 2000 (18): 929–940, arXiv:math/9812166,
doi:10.1155/S1073792800000489, MR1792282
• Deligne, Pierre (1974), "La conjecture de Weil. I" [6], Publications Mathématiques de l'IHÉS 43: 273–307,
doi:10.1007/BF02684373, MR0340258
• Deligne, Pierre (1980), "La conjecture de Weil : II" [7], Publications Mathématiques de l'IHÉS 52: 137–252,
doi:10.1007/BF02684780
• Deninger, Christopher (1998), Some analogies between number theory and dynamical systems on foliated spaces
[8]
, "Proceedings of the International Congress of Mathematicians, Vol. I (Berlin, 1998)", Documenta
Mathematica: 163–186, MR1648030
• Derbyshire, John (2003), Prime Obsession, Joseph Henry Press, Washington, DC, ISBN 978-0-309-08549-6,
MR1968857
• Dyson, Freeman (2009), "Birds and frogs" [9], Notices of the American Mathematical Society 56 (2): 212–223,
MR2483565
• Edwards, H. M. (1974), Riemann's Zeta Function, New York: Dover Publications, ISBN 978-0-486-41740-0,
MR0466039
Riemann hypothesis 39

• Ford, Kevin (2002), "Vinogradov's integral and bounds for the Riemann zeta function", Proceedings of the
London Mathematical Society. Third Series 85 (3): 565–633, doi:10.1112/S0024611502013655, MR1936814
• Franel, J.; Landau, E. (1924), "Les suites de Farey et le problème des nombres premiers", Göttinger Nachr.:
198–206
• Ghosh, Amit (1983), "On the Riemann zeta function---mean value theorems and the distribution of |S(T)|", J.
Number Theory 17: 93–102, doi:10.1016/0022-314X(83)90010-0
• Gourdon, Xavier (2004) (PDF), The 1013 first zeros of the Riemann Zeta function, and zeros computation at very
large height [10]
• Gram, J. P. (1903), "Note sur les zéros de la fonction ζ(s) de Riemann", Acta Mathematica 27: 289–304,
doi:10.1007/BF02421310
• Hadamard, Jacques (1896), "Sur la distribution des zéros de la fonction ζ(s) et ses conséquences arithmétiques"
[11]
, Bulletin Société Mathématique de France 14: 199–220 Reprinted in (Borwein et al. 2008).
• Hardy, G. H. (1914), "Sur les Zéros de la Fonction ζ(s) de Riemann" [12], C. R. Acad. Sci. Paris 158: 1012–1014,
JFM 45.0716.04 Reprinted in (Borwein et al. 2008).
• Hardy, G. H.; Littlewood, J. E. (1921), "The zeros of Riemann's zeta-function on the critical line", Math. Z. 10
(3–4): 283–317, doi:10.1007/BF01211614
• Haselgrove, C. B. (1958), "A disproof of a conjecture of Pólya", Mathematika 5 (02): 141–145,
doi:10.1112/S0025579300001480, MR0104638 Reprinted in (Borwein et al. 2008).
• Haselgrove, C. B.; Miller, J. C. P. (1960), Tables of the Riemann zeta function, Royal Society Mathematical
Tables, Vol. 6, Cambridge University Press, ISBN 978-0-521-06152-0, MR0117905 Review [13]
• Hutchinson, J. I. (1925), "On the Roots of the Riemann Zeta-Function", Transactions of the American
Mathematical Society 27 (1): 49–60, doi:10.2307/1989163, JSTOR 1989163
• Ingham, A.E. (1932), The Distribution of Prime Numbers, Cambridge Tracts in Mathematics and Mathematical
Physics, 30, Cambridge University Press. Reprinted 1990, ISBN 978-0-521-39789-6, MR1074573
• Ivić, A. (1985), The Riemann Zeta Function, New York: John Wiley & Sons, ISBN 978-0-471-80634-9,
MR0792089 (Reprinted by Dover 2003)
• Ivić, Aleksandar (2008), "On some reasons for doubting the Riemann hypothesis", in Borwein, Peter; Choi,
Stephen; Rooney, Brendan et al., The Riemann Hypothesis: A Resource for the Afficionado and Virtuoso Alike,
CMS Books in Mathematics, New York: Springer, pp. 131–160, arXiv:math.NT/0311162,
ISBN 978-0387721255
• Karatsuba, A. A. (1984a), "Zeros of the function ζ(s) on short intervals of the critical line" (in Russian), Izv. Akad.
Nauk SSSR, Ser. Mat. 48 (3): 569–584, MR0747251
• Karatsuba, A. A. (1984b), "Distribution of zeros of the function ζ(1/2+it)" (in Russian), Izv. Akad. Nauk SSSR,
Ser. Mat. 48 (6): 1214–1224, MR0772113
• Karatsuba, A. A. (1985), "Zeros of the Riemann zeta-function on the critical line" (in Russian), Trudy Mat. Inst.
Steklov. (167): 167–178, MR0804073
• Karatsuba, A. A. (1992), "On the number of zeros of the Riemann zeta-function lying in almost all short intervals
of the critical line" (in Russian), Izv. Ross. Akad. Nauk, Ser. Mat. 56 (2): 372–397, MR1180378
• Karatsuba, A. A.; Voronin, S. M. (1992), The Riemann zeta-function, de Gruyter Expositions in Mathematics, 5,
Berlin: Walter de Gruyter & Co., ISBN 978-3-11-013170-3, MR1183467
• Keating, Jonathan P.; Snaith, N. C. (2000), "Random matrix theory and ζ(1/2+it)", Communications in
Mathematical Physics 214 (1): 57–89, doi:10.1007/s002200000261, MR1794265
• Knauf, Andreas (1999), "Number theory, dynamical systems and statistical mechanics", Reviews in Mathematical
Physics. A Journal for Both Review and Original Research Papers in the Field of Mathematical Physics 11 (8):
1027–1060, doi:10.1142/S0129055X99000325, MR1714352
• von Koch, Helge (1901), "Sur la distribution des nombres premiers", Acta Mathematica 24: 159–182,
doi:10.1007/BF02403071
Riemann hypothesis 40

• Kurokawa, Nobushige (1992), "Multiple zeta functions: an example", Zeta functions in geometry (Tokyo, 1990),
Adv. Stud. Pure Math., 21, Tokyo: Kinokuniya, pp. 219–226, MR1210791
• Lapidus, Michel L. (2008), In search of the Riemann zeros, Providence, R.I.: American Mathematical Society,
ISBN 978-0-8218-4222-5, MR2375028
• Lavrik, A. F. (2001), "Zeta-function" [14], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer,
ISBN 978-1556080104
• Lehmer, D. H. (1956), "Extended computation of the Riemann zeta-function", Mathematika. A Journal of Pure
and Applied Mathematics 3 (02): 102–108, doi:10.1112/S0025579300001753, MR0086083
• Levinson, N. (1974), "More than one-third of the zeros of Riemann's zeta function are on σ = 1/2", Adv. In Math.
13 (4): 383–436, doi:10.1016/0001-8708(74)90074-7, MR0564081
• Littlewood, J. E. (1962), "The Riemann hypothesis", The scientist speculates: an anthology of partly baked idea,
New York: Basic books
• van de Lune, J.; te Riele, H. J. J.; Winter, D. T. (1986), "On the zeros of the Riemann zeta function in the critical
strip. IV", Mathematics of Computation 46 (174): 667–681, doi:10.2307/2008005, JSTOR 2008005, MR829637
• Massias, J.-P.; Nicolas, Jean-Louis; Robin, G. (1988), "Évaluation asymptotique de l'ordre maximum d'un
élément du groupe symétrique" [15], Polska Akademia Nauk. Instytut Matematyczny. Acta Arithmetica 50 (3):
221–242, MR960551
• Montgomery, Hugh L. (1973), "The pair correlation of zeros of the zeta function", Analytic number theory, Proc.
Sympos. Pure Math., XXIV, Providence, R.I.: American Mathematical Society, pp. 181–193, MR0337821
Reprinted in (Borwein et al. 2008).
• Montgomery, Hugh L. (1983), "Zeros of approximations to the zeta function", in Erdős, Paul, Studies in pure
mathematics. To the memory of Paul Turán, Basel, Boston, Berlin: Birkhäuser, pp. 497–506,
ISBN 978-3-7643-1288-6, MR820245
• Nicely, Thomas R. (1999), "New maximal prime gaps and first occurrences" [16], Mathematics of Computation 68
(227): 1311–1315, doi:10.1090/S0025-5718-99-01065-0, MR1627813.
• Nyman, Bertil (1950), On the One-Dimensional Translation Group and Semi-Group in Certain Function Spaces,
PhD Thesis, University of Uppsala: University of Uppsala, MR0036444
• Odlyzko, A. M.; te Riele, H. J. J. (1985), "Disproof of the Mertens conjecture" [17], Journal für die reine und
angewandte Mathematik 357: 138–160, MR783538
• Odlyzko, A. M. (1987), "On the distribution of spacings between zeros of the zeta function", Mathematics of
Computation 48 (177): 273–308, doi:10.2307/2007890, JSTOR 2007890, MR866115
• Odlyzko, A. M. (1990), "Bounds for discriminants and related estimates for class numbers, regulators and zeros
of zeta functions: a survey of recent results" [18], Séminaire de Théorie des Nombres de Bordeaux. Série 2 2 (1):
119–141, MR1061762
• Odlyzko, A. M. (1992), The 1020-th zero of the Riemann zeta function and 175 million of its neighbors [19] This
unpublished book describes the implementation of the algorithm and discusses the results in detail.
• Odlyzko, A. M. (1998), The 1021st zero of the Riemann zeta function [20]
• Ono, Ken; Soundararajan, K. (1997), "Ramanujan's ternary quadratic form", Inventiones Mathematicae 130 (3):
415–454, doi:10.1007/s002220050191
• Patterson, S. J. (1988), An introduction to the theory of the Riemann zeta-function, Cambridge Studies in
Advanced Mathematics, 14, Cambridge University Press, ISBN 978-0-521-33535-5, MR933558
• Riemann, Bernhard (1859), "Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse" [21],
Monatsberichte der Berliner Akademie. In Gesammelte Werke, Teubner, Leipzig (1892), Reprinted by Dover,
New York (1953). Original manuscript [22] (with English translation). Reprinted in (Borwein et al. 2008) and
(Edwards 1874)
• Riesel, Hans; Göhl, Gunnar (1970), "Some calculations related to Riemann's prime number formula",
Mathematics of Computation 24 (112): 969–983, doi:10.2307/2004630, JSTOR 2004630, MR0277489
Riemann hypothesis 41

• Riesz, M. (1916), "Sur l'hypothèse de Riemann", Acta Mathematica 40: 185–190, doi:10.1007/BF02418544
• Robin, G. (1984), "Grandes valeurs de la fonction somme des diviseurs et hypothèse de Riemann", Journal de
Mathématiques Pures et Appliquées. Neuvième Série 63 (2): 187–213, MR774171
• Rockmore, Dan (2005), Stalking the Riemann hypothesis, Pantheon Books, ISBN 978-0-375-42136-5,
MR2269393
• Rosser, J. Barkley; Yohe, J. M.; Schoenfeld, Lowell (1969), "Rigorous computation and the zeros of the Riemann
zeta-function. (With discussion)", Information Processing 68 (Proc. IFIP Congress, Edinburgh, 1968), Vol. 1:
Mathematics, Software, Amsterdam: North-Holland, pp. 70–76, MR0258245
• Sabbagh, Karl (2003), The Riemann hypothesis, Farrar, Straus and Giroux, New York, ISBN 978-0-374-25007-2,
MR1979664
• Salem, Raphaël (1953), "Sur une proposition équivalente à l'hypothèse de Riemann", Les Comptes rendus de
l'Académie des sciences 236: 1127–1128, MR0053148
• Sarnak, Peter (2008), "Problems of the Millennium: The Riemann Hypothesis" [23], in Borwein, Peter; Choi,
Stephen; Rooney, Brendan et al. (PDF), The Riemann Hypothesis: A Resource for the Afficionado and Virtuoso
Alike, CMS Books in Mathematics, New York: Springer, pp. 107–115, ISBN 978-0387721255
• du Sautoy, Marcus (2003), The music of the primes, HarperCollins Publishers, ISBN 978-0-06-621070-4,
MR2060134
• Schoenfeld, Lowell (1976), "Sharper bounds for the Chebyshev functions θ(x) and ψ(x). II", Mathematics of
Computation 30 (134): 337–360, doi:10.2307/2005976, JSTOR 2005976, MR0457374
• Schumayer, Daniel; Hutchinson, David A. W. (2011), Physics of the Riemann Hypothesis, arXiv:1101.3116
• Selberg, Atle (1942), "On the zeros of Riemann's zeta-function", Skr. Norske Vid. Akad. Oslo I. 10: 59 pp,
MR0010712
• Selberg, Atle (1946), "Contributions to the theory of the Riemann zeta-function", Arch. Math. Naturvid. 48 (5):
89–155, MR0020594
• Selberg, Atle (1956), "Harmonic analysis and discontinuous groups in weakly symmetric Riemannian spaces with
applications to Dirichlet series", J. Indian Math. Soc. (N.S.) 20: 47–87, MR0088511
• Sheats, Jeffrey T. (1998), "The Riemann hypothesis for the Goss zeta function for Fq[T]", Journal of Number
Theory 71 (1): 121–157, doi:10.1006/jnth.1998.2232, MR1630979
• Siegel, C. L. (1932), "Über Riemanns Nachlaß zur analytischen Zahlentheorie", Quellen Studien zur Geschichte
der Math. Astron. und Phys. Abt. B: Studien 2: 45–80 Reprinted in Gesammelte Abhandlungen, Vol. 1. Berlin:
Springer-Verlag, 1966.
• Speiser, Andreas (1934), "Geometrisches zur Riemannschen Zetafunktion", Mathematische Annalen 110:
514–521, doi:10.1007/BF01448042, JFM 60.0272.04
• Stein, William; Mazur, Barry (2007) (PDF), What is Riemann’s Hypothesis? [24]
• Titchmarsh, Edward Charles (1935), "The Zeros of the Riemann Zeta-Function", Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences (The Royal Society) 151 (873): 234–255,
doi:10.1098/rspa.1935.0146, JSTOR 96545
• Titchmarsh, Edward Charles (1936), "The Zeros of the Riemann Zeta-Function", Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences (The Royal Society) 157 (891): 261–263,
doi:10.1098/rspa.1936.0192, JSTOR 96692
• Titchmarsh, Edward Charles (1986), The theory of the Riemann zeta-function (2nd ed.), The Clarendon Press
Oxford University Press, ISBN 978-0-19-853369-6, MR882550
• Turán, Paul (1948), "On some approximative Dirichlet-polynomials in the theory of the zeta-function of
Riemann", Danske Vid. Selsk. Mat.-Fys. Medd. 24 (17): 36, MR0027305 Reprinted in (Borwein et al. 2008).
• Turing, Alan M. (1953), "Some calculations of the Riemann zeta-function", Proceedings of the London
Mathematical Society. Third Series 3: 99–117, doi:10.1112/plms/s3-3.1.99, MR0055785
Riemann hypothesis 42

• de la Vallée-Poussin, Ch.J. (1896), "Recherches analytiques sur la théorie des nombers premiers", Ann. Soc. Sci.
Bruxelles 20: 183–256
• de la Vallée-Poussin, Ch.J. (1899–1900), "Sur la fonction ζ(s) de Riemann et la nombre des nombres premiers
inférieurs à une limite donnée", Mem. Couronnes Acad. Sci. Belg. 59 (1) Reprinted in (Borwein et al. 2008).
• Weil, André (1948), Sur les courbes algébriques et les variétés qui s'en déduisent, Actualités Sci. Ind., no. 1041 =
Publ. Inst. Math. Univ. Strasbourg 7 (1945), Hermann et Cie., Paris, MR0027151
• Weil, André (1949), "Numbers of solutions of equations in finite fields", Bulletin of the American Mathematical
Society 55 (5): 497–508, doi:10.1090/S0002-9904-1949-09219-4, MR0029393 Reprinted in Oeuvres
Scientifiques/Collected Papers by Andre Weil ISBN 0-387-90330-5
• Weinberger, Peter J. (1973), "On Euclidean rings of algebraic integers", Analytic number theory ( St. Louis Univ.,
1972), Proc. Sympos. Pure Math., 24, Providence, R.I.: Amer. Math. Soc., pp. 321–332, MR0337902
• Wiles, Andrew (2000), "Twenty years of number theory", Mathematics: frontiers and perspectives, Providence,
R.I.: American Mathematical Society, pp. 329–342, ISBN 978-0-8218-2697-3, MR1754786
• Zagier, Don (1977), "The first 50 million prime numbers" [25] (PDF), Math. Intelligencer (Springer) 0: 7–19,
doi:10.1007/BF03039306, MR643810
• Zagier, Don (1981), "Eisenstein series and the Riemann zeta function", Automorphic forms, representation theory
and arithmetic (Bombay, 1979), Tata Inst. Fund. Res. Studies in Math., 10, Tata Inst. Fundamental Res., Bombay,
pp. 275–301, MR633666

External links
• American institute of mathematics, Riemann hypothesis [26]
• Apostol, Tom, Where are the zeros of zeta of s? [27] Poem about the Riemann hypothesis, sung [28] by John
Derbyshire.
• Borwein, Peter (PDF), The Riemann Hypothesis [29] (Slides for a lecture)
• Conrad, K. (2010), Consequences of the Riemann hypothesis [30]
• Conrey, J. Brian; Farmer, David W, Equivalences to the Riemann hypothesis [31]
• Gourdon, Xavier; Sebah, Pascal (2004), Computation of zeros of the Zeta function [32] (Reviews the GUE
hypothesis, provides an extensive bibliography as well).
• Odlyzko, Andrew, Home page [33] including papers on the zeros of the zeta function [34] and tables of the zeros of
the zeta function [35]
• Odlyzko, Andrew (2002) (PDF), Zeros of the Riemann zeta function: Conjectures and computations [36] Slides of
a talk
• Pegg, Ed (2004), Ten Trillion Zeta Zeros [37], Math Games website. A discussion of Xavier Gourdon's calculation
of the first ten trillion non-trivial zeros
• Pugh, Glen, Java applet for plotting Z(t) [38]
• Rubinstein, Michael, algorithm for generating the zeros [39].
• du Sautoy, Marcus (2006), Prime Numbers Get Hitched [40], Seed Magazine [41]
• Stein, William A., What is Riemann's hypothesis [42]
• de Vries, Andreas (2004), The Graph of the Riemann Zeta function ζ(s) [43], a simple animated Java applet.
• Watkins, Matthew R. (2007-07-18), Proposed proofs of the Riemann Hypothesis [44]
• Zetagrid [45] (2002) A distributed computing project that attempted to disprove Riemann's hypothesis; closed in
November 2005
Riemann hypothesis 43

References
[1] http:/ / arxiv. org/ find/ grp_math/ 1/ AND+ ti:+ AND+ Riemann+ hypothesis+ subj:+ AND+ General+ mathematics/ 0/ 1/ 0/ all/ 0/ 1
[2] http:/ / en. wikipedia. org/ wiki/ Oeis%3Aa114856
[3] http:/ / www. claymath. org/ millennium/ Riemann_Hypothesis/ riemann. pdf
[4] http:/ / www. digizeitschriften. de/ resolveppn/ GDZPPN002206781
[5] http:/ / www. ams. org/ notices/ 200303/ fea-conrey-web. pdf
[6] http:/ / www. numdam. org/ item?id=PMIHES_1974__43__273_0
[7] http:/ / www. numdam. org/ item?id=PMIHES_1980__52__137_0
[8] http:/ / www. mathematik. uni-bielefeld. de/ documenta/ xvol-icm/ 00/ Deninger. MAN. html
[9] http:/ / www. ams. org/ notices/ 200902/ rtx090200212p. pdf
[10] http:/ / numbers. computation. free. fr/ Constants/ Miscellaneous/ zetazeros1e13-1e24. pdf
[11] http:/ / www. numdam. org/ item?id=BSMF_1896__24__199_1
[12] http:/ / gallica. bnf. fr/ ark:/ 12148/ bpt6k3111d. image. f1014. langEN
[13] http:/ / www. jstor. org/ stable/ 2003098
[14] http:/ / eom. springer. de/ Z/ z099260. htm
[15] http:/ / matwbn. icm. edu. pl/ tresc. php?wyd=6& tom=50& jez=
[16] http:/ / www. trnicely. net/ gaps/ gaps. html
[17] http:/ / gdz. sub. uni-goettingen. de/ no_cache/ dms/ load/ img/ ?IDDOC=262633
[18] http:/ / www. numdam. org/ item?id=JTNB_1990__2_1_119_0
[19] http:/ / www. dtc. umn. edu/ ~odlyzko/ unpublished/ zeta. 10to20. 1992. pdf
[20] http:/ / www. dtc. umn. edu/ ~odlyzko/ unpublished/ zeta. 10to21. pdf
[21] http:/ / www. maths. tcd. ie/ pub/ HistMath/ People/ Riemann/ Zeta/
[22] http:/ / www. claymath. org/ millennium/ Riemann_Hypothesis/ 1859_manuscript/
[23] http:/ / www. claymath. org/ millennium/ Riemann_Hypothesis/ Sarnak_RH. pdf
[24] http:/ / modular. math. washington. edu/ edu/ 2007/ simuw07/ notes/ rh. pdf
[25] http:/ / modular. math. washington. edu/ edu/ 2007/ simuw07/ misc/ zagier-the_first_50_million_prime_numbers. pdf
[26] http:/ / www. aimath. org/ WWN/ rh/
[27] http:/ / www. math. wisc. edu/ ~robbin/ funnysongs. html#Zeta
[28] http:/ / www. olimu. com/ RIEMANN/ Song. htm
[29] http:/ / oldweb. cecm. sfu. ca/ ~pborwein/ COURSE/ MATH08/ LECTURE. pdf
[30] http:/ / mathoverflow. net/ questions/ 17232
[31] http:/ / aimath. org/ pl/ rhequivalences
[32] http:/ / numbers. computation. free. fr/ Constants/ Miscellaneous/ zetazeroscompute. html
[33] http:/ / www. dtc. umn. edu/ ~odlyzko/
[34] http:/ / www. dtc. umn. edu/ ~odlyzko/ doc/ zeta. html
[35] http:/ / www. dtc. umn. edu/ ~odlyzko/ zeta_tables/ index. html
[36] http:/ / www. dtc. umn. edu/ ~odlyzko/ talks/ riemann-conjectures. pdf
[37] http:/ / www. maa. org/ editorial/ mathgames/ mathgames_10_18_04. html
[38] http:/ / web. viu. ca/ pughg/ RiemannZeta/ RiemannZetaLong. html
[39] http:/ / pmmac03. math. uwaterloo. ca/ ~mrubinst/ l_function_public/ L. html
[40] http:/ / www. seedmagazine. com/ news/ 2006/ 03/ prime_numbers_get_hitched. php
[41] http:/ / www. seedmagazine. com
[42] http:/ / modular. math. washington. edu/ edu/ 2007/ simuw07/ index. html
[43] http:/ / math-it. org/ Mathematik/ Riemann/ RiemannApplet. html
[44] http:/ / secamlocal. ex. ac. uk/ ~mwatkins/ zeta/ RHproofs. htm
[45] http:/ / www. zetagrid. net/
Navier–Stokes existence and smoothness 44

Navier–Stokes existence and smoothness


Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

The Navier–Stokes equations are one of the pillars of fluid mechanics. These equations describe the motion of a
fluid (that is, a liquid or a gas) in space. Solutions to the Navier–Stokes equations are used in many practical
applications. However, theoretical understanding of the solutions to these equations is incomplete. In particular,
solutions of the Navier–Stokes equations often include turbulence, which remains one of the greatest unsolved
problems in physics despite its immense importance in science and engineering.
Even much more basic properties of the solutions to Navier–Stokes have never been proven. For the
three-dimensional system of equations, and given some initial conditions, mathematicians have not yet proved that
smooth solutions always exist, or that if they do exist they have bounded kinetic energy. This is called the
Navier–Stokes existence and smoothness problem.
Since understanding the Navier–Stokes equations is considered to be the first step for understanding the elusive
phenomenon of turbulence, the Clay Mathematics Institute offered in May 2000 a US$1,000,000 prize, not to
whomever constructs a theory of turbulence but (more modestly) to the first person providing a hint on the
phenomenon of turbulence. In that spirit of ideas, the Clay Institute set a concrete mathematical problem:[1]
Prove or give a counter-example of the following statement:
In three space dimensions and time, given an initial velocity field, there exists a vector velocity and a scalar
pressure field, which are both smooth and globally defined, that solve the Navier–Stokes equations.

The Navier–Stokes equations


In mathematics, the Navier–Stokes equations are a system of nonlinear partial differential equations for abstract
vector fields of any size. In physics and engineering, they are a system of equations that models the motion of liquids
or non-rarefied gases using continuum mechanics. The equations are a statement of Newton's second law, with the
forces modeled according to those in a viscous Newtonian fluid—as the sum of contributions by pressure, viscous
stress and an external body force. Since the setting of the problem proposed by the Clay Mathematics Institute is in
three dimensions, for an incompressible and homogeneous fluid, we will consider only that case.
Let be a 3-dimensional vector, the velocity of the fluid, and let be the pressure of the fluid.[2] The
Navier–Stokes equations are:

where is the kinematic viscosity, the external force, is the gradient operator and is the
Laplacian operator, which is also denoted by . Note that this is a vector equation, i.e. it has three scalar
equations. If we write down the coordinates of the velocity and the external force
Navier–Stokes existence and smoothness 45

then for each we have the corresponding scalar Navier–Stokes equation:

The unknowns are the velocity and the pressure . Since in three dimensions we have three
equations and four unknowns (three scalar velocities and the pressure), we need a supplementary equation. This
extra equation is the continuity equation describing the incompressibility of the fluid:

Due to this last property, the solutions for the Navier–Stokes equations are searched in the set of "divergence-free"
functions. For this flow of a homogeneous medium, density and viscosity are constants.
We can eliminate the pressure p by taking an operator rot (alternative notation curl) of both sides of the
Navier–Stokes equations. In this case the Navier–Stokes equations reduce to the Vorticity transport equations. In
two dimensions (2D), these equations are well known [6, p. 321]. In three dimensions (3D), it is known for a long
time that Vorticity transport equations have additional terms [6, p. 294]. However, why 1D, 2D and 3D
Navier–Stokes equations in the vector form are identical? In that case, probably, the vorticity transport equations in
the vector form must be identical too.

Two settings: unbounded and periodic space


There are two different settings for the one-million-dollar-prize Navier–Stokes existence and smoothness problem.
The original problem is in the whole space , which needs extra conditions on the growth behavior of the initial
condition and the solutions. In order to rule out the problems at infinity, the Navier–Stokes equations can be set in a
periodic framework, which implies that we are no longer working on the whole space but in the 3-dimensional
torus . We will treat each case separately.

Statement of the problem in the whole space

Hypotheses and growth conditions


The initial condition is assumed to be a smooth and divergence-free function (see smooth function) such
that, for every multi-index (see multi-index notation) and any , there exists a constant
(i.e. this "constant" depends on and K) such that

for all

The external force is assumed to be a smooth function as well, and satisfies a very analogous inequality
(now the multi-index includes time derivatives as well):

for all

For physically reasonable conditions, the type of solutions expected are smooth functions that do not grow large as
. More precisely, the following assumptions are made:

1.

2. There exists a constant such that for all

Condition 1 implies that the functions are smooth and globally defined and condition 2 means that the kinetic energy
of the solution is globally bounded.
Navier–Stokes existence and smoothness 46

The million-dollar-prize conjectures in the whole space


(A) Existence and smoothness of the Navier–Stokes solutions in
Let . For any initial condition satisfying the above hypotheses there exist smooth and globally
defined solutions to the Navier–Stokes equations, i.e. there is a velocity vector and a pressure
satisfying conditions 1 and 2 above.
(B) Breakdown of the Navier–Stokes solutions in
There exists an initial condition and an external force such that there exists no solutions
and satisfying conditions 1 and 2 above.

Statement of the periodic problem

Hypotheses
The functions we seek now are periodic in the space variables of period 1. More precisely, let be the unitary
vector in the j- direction:

Then is periodic in the space variables if for any we have that

Notice that we are considering the coordinates mod 1. This allows us to work not on the whole space but on the
quotient space , which turns out to be the 3-dimensional torus

We can now state the hypotheses properly. The initial condition is assumed to be a smooth and
divergence-free function and the external force is assumed to be a smooth function as well. The type of
solutions that are physically relevant are those who satisfy these conditions:
3.

4. There exists a constant such that for all

Just as in the previous case, condition 3 implies that the functions are smooth and globally defined and condition 4
means that the kinetic energy of the solution is globally bounded.

The periodic million-dollar-prize theorems


(C) Existence and smoothness of the Navier–Stokes solutions in
Let . For any initial condition satisfying the above hypotheses there exist smooth and globally
defined solutions to the Navier–Stokes equations, i.e. there is a velocity vector and a pressure
satisfying conditions 3 and 4 above.
(D) Breakdown of the Navier–Stokes solutions in
There exists an initial condition and an external force such that there exists no solutions
and satisfying conditions 3 and 4 above.
Navier–Stokes existence and smoothness 47

Partial results
1. The Navier–Stokes problem in two dimensions has already been solved positively since the 1960s: there exist
smooth and globally defined solutions.[3]
2. If the initial velocity is sufficiently small then the statement is true: there are smooth and globally
defined solutions to the Navier–Stokes equations.[1]
3. Given an initial velocity there exists a finite time T, depending on such that the Navier–Stokes
equations on have smooth solutions and . It is not known if the solutions exist
beyond that "blowup time" T.[1]
4. The mathematician Jean Leray in 1934 proved the existence of so called weak solutions to the Navier–Stokes
equations, satisfying the equations in mean value, not pointwise.[4]

Notes
[1] Official statement of the problem (http:/ / www. claymath. org/ millennium/ Navier-Stokes_Equations/ navierstokes. pdf), Clay Mathematics
Institute.
[2] More precisely, is the pressure divided by the fluid density, and the density is constant for this incompressible and homogeneous
fluid.
[3] Ladyzhenskaya, O. (1969), The Mathematical Theory of Viscous Incompressible Flows (2nd ed.), New York: Gordon and Breach.
[4] Leray, J. (1934), "Sur le mouvement d'un liquide visqueux emplissant l'espace", Acta Mathematica 63: 193–248, doi:10.1007/BF02547354

References

External links
• The Clay Mathematics Institute's Navier–Stokes equation prize (http://www.claymath.org/millennium/
Navier-Stokes_Equations/)
• Why global regularity for Navier–Stokes is hard (http://terrytao.wordpress.com/2007/03/18/
why-global-regularity-for-navier-stokes-is-hard) — Possible routes to resolution are scrutinized by Terence Tao.
• Fuzzy Fluid Mechanics (http://sgrajeev.com/fuzzy-fluids/)
• Navier–Stokes existence and smoothness (Millennium Prize Problem) (http://vimeo.com/18185364/) A lecture
on the problem by Luis Caffarelli.
Birch and Swinnerton-Dyer conjecture 48

Birch and Swinnerton-Dyer conjecture


Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture

In mathematics, the Birch and Swinnerton-Dyer conjecture is an open problem in the field of number theory. Its
status as one of the most challenging mathematical questions has become widely recognized; the conjecture was
chosen as one of the seven Millennium Prize Problems listed by the Clay Mathematics Institute, which has offered a
$1,000,000 prize for the first correct proof. As of 2010, only special cases of the conjecture have been proved
correct.
The conjecture relates arithmetic data associated to an elliptic curve E over a number field K to the behaviour of the
Hasse-Weil L-function L(E, s) of E at s = 1. More specifically, it is conjectured that the rank of the abelian group
E(K) of points of E is the order of the zero of L(E, s) at s = 1, and the first non-zero coefficient in the Taylor
expansion of L(E, s) at s = 1 is given by more refined arithmetic data attached to E over K.[1]

Background
In 1922 Louis Mordell proved Mordell's theorem: the group of rational points on an elliptic curve has a finite basis.
This means that for any elliptic curve there is a finite sub-set of the rational points on the curve, from which all
further rational points may be generated.
If the number of rational points on a curve is infinite then some point in a finite basis must have infinite order. The
number of independent basis points with infinite order is called the rank of the curve, and is an important invariant
property of an elliptic curve.
If the rank of an elliptic curve is 0, then the curve has only a finite number of rational points. On the other hand, if
the rank of the curve is greater than 0, then the curve has an infinite number of rational points.
Although Mordell's theorem shows that the rank of an elliptic curve is always finite, it does not give an effective
method for calculating the rank of every curve. The rank of certain elliptic curves can be calculated using numerical
methods but (in the current state of knowledge) these cannot be generalised to handle all curves.
An L-function L(E, s) can be defined for an elliptic curve E by constructing an Euler product from the number of
points on the curve modulo each prime p. This L-function is analogous to the Riemann zeta function and the
Dirichlet L-series that is defined for a binary quadratic form. It is a special case of a Hasse-Weil L-function.
The natural definition of L(E, s) only converges for values of s in the complex plane with Re(s) > 3/2. Helmut Hasse
conjectured that L(E, s) could be extended by analytic continuation to the whole complex plane. This conjecture was
first proved by Max Deuring for elliptic curves with complex multiplication. It was subsequently shown to be true
for all elliptic curves over Q, as a consequence of the modularity theorem.
Finding rational points on a general elliptic curve is a difficult problem. Finding the points on an elliptic curve
modulo a given prime p is conceptually straightforward, as there are only a finite number of possibilities to check.
However, for large primes it is computationally intensive.
Birch and Swinnerton-Dyer conjecture 49

History
In the early 1960s Peter Swinnerton-Dyer used the EDSAC computer at the University of Cambridge Computer
Laboratory to calculate the number of points modulo p (denoted by Np) for a large number of primes p on elliptic
curves whose rank was known. From these numerical results Bryan Birch and Swinnerton-Dyer conjectured that Np
for a curve E with rank r obeys an asymptotic law

A plot of for the curve y2 = x3 − 5x as X varies over the first 100000 primes.

The X-axis is log(log(X)) and Y-axis is in a logarithmic scale so the conjecture predicts
that the data should form a line of slope equal to the rank of the curve, which is 1 in this
case. For comparison, a line of slope 1 is drawn in red on the graph.

where C is a constant.
Initially this was based on somewhat tenuous trends in graphical plots; which induced a measure of skepticism in J.
W. S. Cassels (Birch's Ph.D. advisor). Over time the numerical evidence stacked up.
This in turn led them to make a general conjecture about the behaviour of a curve's L-function L(E, s) at s = 1,
namely that it would have a zero of order r at this point. This was a far-sighted conjecture for the time, given that the
analytic continuation of L(E, s) there was only established for curves with complex multiplication, which were also
the main source of numerical examples. (NB that the reciprocal of the L-function is from some points of view a more
natural object of study; on occasion this means that one should consider poles rather than zeroes.)
The conjecture was subsequently extended to include the prediction of the precise leading Taylor coefficient of the
L-function at s = 1. It is conjecturally given by

where the quantities on the right hand side are invariants of the curve, studied by Cassels, Tate, Shafarevich and
others: these include the order of the torsion group, the order of the Tate-Shafarevich group, and the canonical
heights of a basis of rational points.[1]
Birch and Swinnerton-Dyer conjecture 50

Current status
The Birch and Swinnerton-Dyer conjecture has been proved only in special cases :
1. In 1976, John Coates and Andrew Wiles proved that if E is a curve over a number field F with complex
multiplication by an imaginary quadratic field K of class number 1, F = K or Q, and L(E, 1) is not 0 then E(F) is a
finite group.[2] This was extended to the case where F is any finite abelian extension of K by Nicole
Arthaud-Kuhman.[3]
2. In 1983, Benedict Gross and Don Zagier showed that if a modular elliptic curve has a first-order zero at s = 1 then
it has a rational point of infinite order;[4] see Gross–Zagier theorem.
3. In 1990, Victor Kolyvagin showed that a modular elliptic curve E for which L(E,1) is not zero has rank 0, and a
modular elliptic curve E for which L(E,1) has a first-order zero at s = 1 has rank 1.
4. In 1991, Karl Rubin showed that for elliptic curves defined over an imaginary quadratic field K with complex
multiplication by K, if the L-series of the elliptic curve was not zero at s=1, then the p-part of the
Tate–Shafarevich group had the order predicted by the Birch and Swinnerton-Dyer conjecture, for all primes p >
7.[5]
5. In 2001, Christophe Breuil, Brian Conrad, Fred Diamond and Richard Taylor, extending work of Wiles, proved
that all elliptic curves defined over the rational numbers are modular (the Taniyama-Shimura theorem), which
extends results 2 and 3 to all elliptic curves over the rationals, and shows that the L-functions of all elliptic curves
over Q are defined at s = 1.[6]
6. In 2010, Manjul Bhargava and Arul Shankar announced a proof that the average rank of the Mordell–Weil group
of an elliptic curve over Q is bounded above by 7/6.[7] Combining this with the announced proof of the main
conjecture of Iwasawa theory for GL(2) by Chris Skinner and Éric Urban,[8] they conclude that a positive
proportion of elliptic curves over Q have analytic rank zero, and hence, by Kolyvagin's result, satisfy the Birch
and Swinnerton-Dyer conjecture.
Nothing has been proved for curves with rank greater than 1, although there is extensive numerical evidence for the
truth of the conjecture.

Clay Mathematics Institute Prize


The Birch and Swinnerton-Dyer conjecture is one of the seven Millennium Problems selected by the Clay
Mathematics Institute, which is offering a prize of $1 million for the first proof or disproof of the whole
conjecture.[9]

Notes
[1] Wiles 2006
[2] Coates, J.; Wiles, A. (1977). "On the conjecture of Birch and Swinnerton-Dyer". Inventiones Mathematicae 39 (3): 223–251.
doi:10.1007/BF01402975.
[3] Arthaud, Nicole (1978). "On Birch and Swinnerton-Dyer's conjecture for elliptic curves with complex multiplication". Compositio
Mathematica 37 (2): 209–232. MR504632.
[4] Gross, Benedict H.; Zagier, Don B. (1986). "Heegner points and derivatives of L-series". Inventiones Mathematicae 84 (2): 225–320.
doi:10.1007/BF01388809. MR0833192.
[5] Rubin, Karl (1991). "The 'main conjectures' of Iwasawa theory for imaginary quadratic fields". Inventiones Mathematicae 103 (1): 25–68.
doi:10.1007/BF01239508.
[6] Breuil, Christophe; Conrad, Brian; Diamond, Fred; Taylor, Richard (2001). "On the Modularity of Elliptic Curves over Q: Wild 3-Adic
Exercises". Journal of the American Mathematical Society 14 (4): 843–939. doi:10.1090/S0894-0347-01-00370-8.
[7] Bhargava, Manjul; Shankar, Arul (2010). "Ternary cubic forms having bounded invariants, and the existence of a positive proportion of
elliptic curves having rank 0". Preprint. arXiv:1007.0052.
[8] Skinner, Chris; Urban, Éric (2010). "The Iwasawa main conjectures for GL2" (http:/ / www. math. columbia. edu/ ~urban/ eurp/ MC. pdf). In
preparation. .
[9] Birch and Swinnerton-Dyer Conjecture (http:/ / www. claymath. org/ millennium/ Birch_and_Swinnerton-Dyer_Conjecture/ ) at Clay
Mathematics Institute
Birch and Swinnerton-Dyer conjecture 51

References
• Wiles, Andrew (2006), "The Birch and Swinnerton-Dyer conjecture" (http://www.claymath.org/millennium/
Birch_and_Swinnerton-Dyer_Conjecture/birchswin.pdf), in Carlson, James; Jaffe, Arthur; Wiles, Andrew, The
Millennium prize problems, American Mathematical Society, pp. 31–44, ISBN 978-0-821-83679-8

External links
• Weisstein, Eric W., " Swinnerton-Dyer Conjecture (http://mathworld.wolfram.com/
Swinnerton-DyerConjecture.html)" from MathWorld.
• Birch and Swinnerton-Dyer Conjecture (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=4561)
on PlanetMath
• The Birch and Swinnerton-Dyer Conjecture (http://sums.mcgill.ca/delta-epsilon/mag/0610/mmm061024.
pdf): An Interview with Professor Henri Darmon by Agnes F. Beaudry

Yang–Mills existence and mass gap


The Clay Mathematics Institute has offered a prize of US$1,000,000 to a person solving any one of the Millennium
Prize Problems of modern mathematics. One of these seven problems is phrased as follows:
Yang–Mills Existence and Mass Gap. Prove that for any compact simple gauge group G, a non-trivial
quantum Yang–Mills theory exists on R4 and has a mass gap Δ > 0. Existence includes establishing axiomatic
properties at least as strong as those cited in [45, 35].
In this statement, Yang–Mills theory is the (non-Abelian) quantum field theory underlying the Standard Model of
particle physics; R4 is space-time; the mass gap Δ is the mass of the least massive particle predicted by the theory.
Therefore, the winner must first prove that Yang–Mills theory exists and that it satisfies the standard of rigor that
characterizes contemporary mathematical physics, in particular constructive quantum field theory, which is
referenced in the papers 45 and 35 cited in the official problem description by Jaffe and Witten. The winner must
then prove that the mass of the least massive particle of the force field predicted by the theory is strictly positive. For
example, in the case of G=SU(3) - the strong nuclear interaction - the winner must prove that glueballs have a lower
mass bound, and thus cannot be arbitrarily light.

Millennium Prize Problems

P versus NP problem

Hodge conjecture

Poincaré conjecture (solution)

Riemann hypothesis

Yang–Mills existence and mass gap

Navier–Stokes existence and smoothness

Birch and Swinnerton-Dyer conjecture


Yang–Mills existence and mass gap 52

Background


[...] one does not yet have a mathematically complete example of a quantum gauge theory in four-dimensional space-time, nor even a precise
definition of quantum gauge theory in four dimensions. Will this change in the 21st century? We hope so!

—From the Clay Institute's official problem description by Arthur Jaffe and Edward Witten.

Most known and nontrivial (i.e. interacting) quantum field theories in 4 dimensions are effective field theories with a
cutoff scale. Since the beta-function is positive for most models, it appears that most such models have a Landau
pole as it is not at all clear whether or not they have nontrivial UV fixed points. This means that if such a QFT is
well-defined at all scales, as it has to be to satisfy the axioms of axiomatic quantum field theory, it would have to be
trivial (i.e. a free field theory).
Quantum Yang-Mills theory with a non-abelian gauge group and no quarks is an exception, because asymptotic
freedom characterizes this theory, meaning that it has a trivial UV fixed point. Hence it is the simplest nontrivial
constructive QFT in 4 dimensions. (QCD is a more complicated theory because it involves quarks.)
It has already been well proven—at least at the level of rigor of theoretical physics but not that of mathematical
physics—that the quantum Yang–Mills theory for a non-abelian Lie group exhibits a property known as
confinement. This property is covered in more detail in the relevant QCD articles (QCD, color confinement, lattice
gauge theory, etc.), although not at the level of rigor of mathematical physics. A consequence of this property is that
beyond a certain scale, known as the QCD scale (more properly, the confinement scale, as this theory is devoid of
quarks), the color charges are connected by chromodynamic flux tubes leading to a linear potential between the
charges. Hence free color charge and free gluons cannot exist. In the absence of confinement, we would expect to see
massless gluons, but since they are confined, all we see are color-neutral bound states of gluons, called glueballs. If
glueballs exist, they are massive, which is why we expect a mass gap.
Results from lattice gauge theory have convinced many that quantum Yang–Mills theory for a non-abelian Lie group
model exhibits confinement—as indicated, for example, by an area law for the falloff of the vacuum expectation
value (VEV) of a Wilson loop. However, these methods and results are not mathematically rigorous.

References
• Arthur Jaffe and Edward Witten "Quantum Yang-Mills theory. [1]" Official problem description.

External links
• The Millennium Prize Problems: Yang–Mills and Mass Gap [2]

References
[1] http:/ / www. claymath. org/ millennium/ Yang-Mills_Theory/ yangmills. pdf
[2] http:/ / www. claymath. org/ millennium/ Yang-Mills_Theory/
Article Sources and Contributors 53

Article Sources and Contributors


Poincaré conjecture  Source: http://en.wikipedia.org/w/index.php?oldid=428574621  Contributors: 67-21-48-122, Acer, Aeusoes1, Ahoerstemeier, Akriasas, Alba, AlexBG72, Alpharigel,
AnRtist, Anonymous Dissident, Antonio Prates, Arcfrk, Arne Hermann, ArnoldReinhold, Asyndeton, Avihu, AxelBoldt, BACbKA, Bender235, BlackPanther, Bobblewik, Brian0918, Bryan
Derksen, Btball, Bugwit, C S, CBM, CRGreathouse, Caeruleancentaur, Chadloder, Charles Matthews, Chasingamy, Chatool, Chinju, Choess, ChongDae, ChrisHardie, Closedmouth,
Cmdrjameson, Colonies Chris, Congyu, Connelly, Conversion script, Crazyarabian, Csigabi, Css, Cullinane, Cybercobra, Daqu, Davemck, David Haslam, Db099221, Deathwing23, Delirium,
Dominus, Dorftrottel, Downwards, Dr Zak, Dr mindbender, Edinborgarstefan, Eigenlambda, ElTchanggo, Eleassar777, Emperor657, Eraserhead1, Ewen, Excirial, Filippowiki, Fintelia, Francis
Schonken, Freakofnurture, Fredrik, Fropuff, FvdP, Gareth Owen, Garybrimley, Gauss, Geometry guy, Gg1966, Giftlite, Glenn, Godzilla00, GraemeMcRae, Gregregregre, Gruen, Headbomb,
Henry Delforn, Hermel, Hofingerandi, Inoculatedcities, Isomorphic, Issildur, Jayen466, Jcarroll, Jedishive, Jesse Viviano, Jitse Niesen, JohnyDog, JorgePeixoto, Jowa fan, Jtwdog, KT322,
Kaimiddleton, Karada, Kasreyn, Keenan Pepper, Kidiawipe, Klaus, Kwamikagami, Kylemahan, Kylu, L Kensington, Lailoken, Lethe, Lfh, Lightdarkness, Lightmouse, Ling.Nut, Lousyd,
LoveEncounterFlow, Lumos3, Macintosh User, Magioladitis, Mal4mac, MarkSweep, MathMartin, Mathslover, Maxal, Mehmety, Michael Hardy, Mignon, Mike Peel, Mikeblas, Misza13, Mm
202, Mnemo, Moreton bay bug, Msh210, Mwarren us, Myaca, Myasuda, Nbarth, Nick C, Numbo3, Nv8200p, Oleg Alexandrov, Ox don, Ozob, Partialdifferential, Paul August, Populus,
Qmwne235, Qutezuce, Qwfp, R.e.b., Rade Kutil, Rake, Ranicki, Rcinda1, Redixx, ReyBrujo, Rfc1394, Rich Farmbrough, Rjwilmsi, Rl, Russavia, Russgeom, S2000magician, ST47, Salix alba,
Sam nead, Schizobullet, Sclerolith, Senordingdong, Sevenstones, Sfahey, Shalom Yechiel, Shmuelw, Shrike, Singularity, Sk2006 8, Skyfiler, Squarecircle, Stephen B Streater, Superior IQ
Genius, Suruena, Susvolans, Sverdrup, Sławomir Biały, Tamfang, Tarquin, Tayste, Tesseran, The Anome, The Thing That Should Not Be, Thehotelambush, Timwi, Tkuvho, Toby, Toby Bartels,
Tonysu, Tosha, UkPaolo, Uncia, Uvgm, Vaughan Pratt, Vespristiano, W foxx, Wapcaplet, Wiki alf, Wikiklrsc, Wissenschaft, Wolfkeeper, Wwannsda, Xyb, Yill577, Ysangkok, Zundark, 272
anonymous edits

P versus NP problem  Source: http://en.wikipedia.org/w/index.php?oldid=429733008  Contributors: 128.138.87.xxx, 194.117.133.xxx, 62.202.117.xxx, A bit iffy, Action Jackson IV, Adam
McMaster, Adityad, Alexwatson, Algebraist, Alksentrs, Altenmann, Andejons, Andreas Kaufmann, Andris, Anog, Anonymous Dissident, Anonymousacademic, Anonymousone2, Archibald
Fitzchesterfield, Arichnad, Arthur Frayn, Arthur Rubin, Arvindn, Aseld, Asmeurer, AstroNomer, AxelBoldt, Azimuth1, Banus, Bdesham, Beb0s, Bender235, Bevo74, Bigtimepeace, Bihco,
Bkell, Blazotron, Blokhead, Blotwell, Bofa, Booyabazooka, Brianjd, Brighterorange, C S, CALR, CBKAtTopsails, CBM, CRGreathouse, Calculuslover, Calvin 1998, Canon, Cf. Hay,
ChandraKarChandra, Charles Matthews, Chinju, Chocolateboy, Chrisjwmartin, Civil Engineer III, Ckplato, CloudNine, Cngoulimis, Cointyro, Cole Kitchen, Conversion script, Creidieki,
Cybercobra, Cyde, DGG, Damian Yerrick, DanielMayer, Dantheox, Dark Charles, David Eppstein, David Gerard, David.Monniaux, Davidhorman, Dayewalker, Dcoetzee, Deadcode, Debresser,
Delmet, Dereckson, Dillon256, Dissident, Dlakavi, Docu, Donarreiskoffer, Doradus, Doulos Christos, Download, Dragon guy, Dspart, Duncan, Duncancumming, ELLinng, ESkog, Eescardo,
Eggishorn, Elizabeyth, EmersonLowry, EmilJ, Emurphy42, Epbr123, Eric119, Ewa5050, FarzanehSarafraz, Favonian, Fcady2007, Feinoha, Feynman81, Fiachra10003, Frazzydee, Fredrik,
Frungi, Gakrivas, Gandalf61, Gary King, Gavia immer, Gdr, Giftlite, Gonzolito, GraemeMcRae, Graham87, GromXXVII, Gtcostello, Guoguo12, Gyllstromk, Hairy Dude, Hcsradek, Hideyuki,
Hobophobe, Hofingerandi, Hritcu, Humanengr, Husky, Ianhowlett, Icairns, Ihope127, Illuminatedwax, Iman.saleh, Intgr, IvanAndreevich, Ixfd64, JRSpriggs, Jacob grace, Jammycakes, Jan
Hidders, Jaybuffington, Jayme, Jeff G., Jeff3000, Jeremy W Powell, Jiayq, Jitse Niesen, Jmah, John Vandenberg, Jon Awbrey, Jonnty, Jossi, Jtwdog, Julesd, Justin Stafford, Justin W Smith,
Kayau, Kmote, Kralizec!, Kufat, Kvikram, Kzzl, L Kensington, LC, LJosil, LOL, Lambiam, Landroo, Laurusnobilis, LeoNomis, Libertyrights, Liron00, Logan, Lotu, LouScheffer, Lowellian,
Luckas Blade, Lukeh1, MATThematical, MER-C, MSGJ, Marknau, Markvs88, MartinMusatov, Mathiastck, MatthewH, Mattiabona, Mcsee, Mellum, Mentifisto, Mgiganteus1, Mgraham831,
Michael Hardy, Mikaey, Minesweeper, Miym, Mpeisenbr, Mrmagoo2006, Ms2ger, Mshebanow, Msikma, Nageh, Najoj, Narsil, Navigatr85, Nbhatla, Neutrality, Nicolaennio, Night Gyr, Nith
Sahor, Noamz, NormDor, NuclearWarfare, Obradovic Goran, Oddity-, Olivier, Osias, Otisjimmy1, Paddu, ParotWise, Patelhiren.101, Pcap, Perey, Phil Boswell, Phil Sandifer, Pichpich, Pmadrid,
PoolGuy, Poor Yorick, Predawn, Prodego, Pxtreme75, QuantitativeFinanceKinderChocolate, Qwertyus, R.e.b., R6144, RTC, Raiden09, Rajpaj, Raven4x4x, Rbarreira, Rdnk, Remy B, Rhythm,
Rich Farmbrough, Rick Block, Rjlipton, Rjwilmsi, Robbar, Robert K S, Robert Merkel, RobinK, Robinh, Rockiesfan19, Rorro, RoySmith, Rs561, Rspeer, Rspence1234, Rtc, Rupert Clayton,
Ruud Koot, Ryan Postlethwaite, Sabb0ur, Salgueiro, ScWizard, Schneelocke, Shawn comes, Shreevatsa, Sith Lord 13, Smj2118, Speculatrix, Srinivasasha, Staecker, Stan Marian C-tin,
Stardust8212, Stephen B Streater, Stifle, Stuart P. Bentley, Subtilior, Sundar, Suruena, Svnsvn, Tbhotch, Tcotco, Telekid, Template namespace initialisation script, Teorth, The Obfuscator, The
Thing That Should Not Be, Thehotelambush, Thorbjørn Ravn Andersen, Thore, Tide rolls, Timc, Timwi, Tizio, Tobias Bergemann, Tommy2010, Trovatore, Ttonyb1, Twin Bird, Twri, UkPaolo,
Ultimus, Ustadny, Vadim Makarov, VernoWhitney, Versus22, Vivekk, Waldir, WavePart, Wikiklrsc, Wikinick, Wikipediatrist, Wmahan, WookieInHeat, Wtt, Wtuvell, X7q, Xiaodai,
Yellowdesk, Ylloh, ^demon, 672 anonymous edits

Hodge conjecture  Source: http://en.wikipedia.org/w/index.php?oldid=419614816  Contributors: Baccyak4H, BenTheMen, Bender235, Benzh, BerndGehrmann, Bihco, Changbao, Charles
Matthews, ChicXulub, ComplexZeta, Cícero, DanielDeibler, David Haslam, Dfrg.msc, Drbreznjev, Eoladis, Gareth Jones, Giftlite, Gtrmp, Hofingerandi, Ilmari Karonen, Jakob.scholbach,
Jtwdog, KSmrq, LJosil, Master Bratac, Michael Hardy, Michael Slone, Nzseries1, Ozob, Pruneau, Silvonen, Snowdog, Spireguy, Steven1234321, Suruena, Sławomir Biały, The Anome, Tong,
UkPaolo, Willking1979, ZX81, 46 anonymous edits

Riemann hypothesis  Source: http://en.wikipedia.org/w/index.php?oldid=430834453  Contributors: 13shajiia, 345Kai, A-Doo, Aiden Fisher, Albert Einstein2011, AliceNovak, Allmightyduck,
Andrei Stroe, Anonymous Dissident, Antandrus, Anupam, Army1987, Arthur Rubin, Atlantia, AxelBoldt, Aydinkutlu, Baiji, Barticus88, Bender235, BigFatBuddha, Bkkbrad, Boing! said
Zebedee, Bubba73, C S, CRGreathouse, Calton, Can't sleep, clown will eat me, Carnildo, Cedars, Cenarium, Charles Matthews, Chartguy, Ched Davis, Chinju, Chocolateboy, Chrisguidry, Cole
Kitchen, Conscious, Conversion script, Cool Blue, Copedance, Corkgkagj, Crisófilax, DBrane, DYLAN LENNON, Daqu, David Eppstein, David Gerard, David Haslam, Deadbarnacle,
DerHexer, Dicconb, DiceDiceBaby, Discospinster, Dmharvey, Dod1, Dominic, Doomed Rasher, Dr. Leibniz, Dr. Megadeth, Droog Andrey, Dto, Dysprosia, ERcheck, Egg, Einsteinino,
Ekaratsuba, Emersoni, Emholvoet, EmilJ, Epbr123, Eric119, Ericamick, Erud, Estel, Evercat, F3et, Fetchcomms, Fournax, Fredrik, Frenchwhale, Freticat, Fullmetal2887, Fumblebruschi,
GTBacchus, Gandalf61, Gary King, Gavia immer, Gene Ward Smith, Gershwinrb, Giftlite, Gika, Gingermint, GirasoleDE, Gsmgm, Guardian of Light, Gurch, HaeB, Halo, Harleyjamesmunro,
He Who Is, Headbomb, Helohe, Henry Delforn, Herbee, Hofingerandi, Hu, Hut 8.5, Iamunknown, IanOsgood, Ianmacm, Icairns, Ilyanep, InvertRect, IronSwallow, J.delanoy, JW1805,
JackSchmidt, Jacobolus, Jakob.scholbach, Jeppesn, Jheald, Jijinmachina, Jitse Niesen, Joel.Gilmore, Jtwdog, JuPitEer, Jujutacular, Julesd, Kapalama, Karl-H, Karl-Henner, Katsushi, Kdoto, Kh7,
Kingdon, Klaus, KnowledgeOfSelf, Krea, LDH, LJosil, LiDaobing, Linas, Loadmaster, Looxix, LouisWins, Lzur, Madmath789, Magicxcian, Maha ts, MarkSutton, Marudubshinki, Mcarling,
Mcsee, Meekohi, Michael Hardy, Million Little Gods, Mindmatrix, Mistamagic28, Mon4, Motomuku, Mpatel, Mpeisenbr, Mrhawley, Myasuda, NatusRoma, Neilc, Nicolasqueen, Numerao,
Obradovic Goran, Ocolon, Ofap, Oleg Alexandrov, Olivier, Openshac, Opustylnikov, Oshanker, Overlord 77520, PL290, Pasky, Peterungar, Phil Boswell, Piet Delport, Pip2andahalf,
Plastikspork, Pleasantville, PrimeHunter, Profstein, Pstudier, Pt, Qui1che, R.e.b., REGULAR-NORMAL, Raul654, Raulshc, Ravi12346, Rchamberlin, Reinyday, Revolver, Rgclegg, Rich
Farmbrough, Riemann'sZeta, Ripe, Rivertorch, Rjwilmsi, RobHar, RobertG, Rodhullandemu, Ruakh, Rylann, STRANGELUV3, Saccerzd, Saga City, Salix alba, Sallypally, Sbjf, Schneelocke,
Scythe33, Seglea, Seraphita, Sherbrooke, Shimgray, Shreevatsa, Sidis24, Slonzor, Snthdiueoa, Speight, Suruena, Svetovid, Symplectic Map, TakuyaMurata, Tarquin, Tassedethe, Taxman,
Teeesssyyy, The Anome, TheCustomOfLife, TheSeven, Thehotelambush, Timhoooey, Timothy Clemans, Tobias Bergemann, Tpbradbury, Travelbird, TravisAF, Twilsonb, UkPaolo,
Venona2007, Vinoo Cameron, Vipul, Voidxor, WAREL, Wafulz, Waprap, Wile E. Heresiarch, William Avery, Woseph, Wshun, Wwwwolf, XJamRastafire, Yurakm, ZX81, Zaian, Zondor, 408
anonymous edits

Navier–Stokes existence and smoothness  Source: http://en.wikipedia.org/w/index.php?oldid=425571686  Contributors: -Kerplunk-, Amble, Bender235, BryanD, Bunder, Bvdano, C S,
CRGreathouse, Calton, Cfp, Charlesreid1, Crowsnest, Cybercobra, Dicklyon, Dmitri Gorskin, Egbertus, Flup, Freelance Intellectual, Gareth E Kegg, Gbuffett, Giftlite, Goluskin, Gramolin,
Harryboyles, Ixfd64, Jhausauer, Jitse Niesen, Jon513, Leondumontfollower, LouScheffer, Mevalabadie, Michael Hardy, Mpatel, Nneonneo, Noetica, Oleg Alexandrov, Oub, Patrick Denny,
Perturbationist, Rbcoulter, RekishiEJ, ScWizard, Sincurp, Suruena, TV4Fun, Tjschuck, User A1, Wikiboyzz, Wzhx cc, 99 anonymous edits

Birch and Swinnerton-Dyer conjecture  Source: http://en.wikipedia.org/w/index.php?oldid=428916980  Contributors: Ampassag, Bender235, Borat fan, CBM, CRGreathouse, Charles
Matthews, CryptoDerk, David Eppstein, David Haslam, DiceDiceBaby, Epsilon17, Gandalf61, Gauge, Giftlite, Hesam7, Hofingerandi, J.delanoy, Jaraalbe, Jpbowen, Jtwdog, Lenthe, Linas,
Lowellian, MathMartin, Maximus Rex, Michael Hardy, Oleg Yunakov, Phe, Qutezuce, R.e.b., Rajpaj, Raulshc, Rhythm, RobHar, Robertwb, Saga City, Samwb123, Squeezy, Suruena, Tbsmith,
Thewhyman, Timothy Clemans, Timwi, UkPaolo, Waltpohl, Zoicon5, 40 anonymous edits

Yang–Mills existence and mass gap  Source: http://en.wikipedia.org/w/index.php?oldid=420976450  Contributors: Abandoned, Aldynin, Bender235, CBM, CDN99, CRGreathouse, Charles
Matthews, ClaudeDes, Edpell3, Ernie shoemaker, Evercat, Guy Harris, Hillgentleman, Hofingerandi, Jon Awbrey, Jtwdog, Klaus, Loodog, Nousernamesleft, Paul Foxworthy, Phys, Pt, Reyk,
Rich Farmbrough, SQGibbon, Suruena, Tabarr, Thesocool, Traitor de, UkPaolo, Was a bee, Werewolf Bar Mitzvah, Ysangkok, 青子守歌, 31 anonymous edits
Image Sources, Licenses and Contributors 54

Image Sources, Licenses and Contributors


Image:P1S2all.jpg  Source: http://en.wikipedia.org/w/index.php?title=File:P1S2all.jpg  License: GNU Free Documentation License  Contributors: User:Salix alba
Image:Ricci flow.png  Source: http://en.wikipedia.org/w/index.php?title=File:Ricci_flow.png  License: Public Domain  Contributors: CBM
File:Complexity classes.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Complexity_classes.svg  License: Public Domain  Contributors: Booyabazooka, Mdd, Mike1024, 1
anonymous edits
File:P np np-complete np-hard.svg  Source: http://en.wikipedia.org/w/index.php?title=File:P_np_np-complete_np-hard.svg  License: Creative Commons Attribution-Sharealike 3.0,2.5,2.0,1.0
 Contributors: Original uploader was Behnam at en.wikipedia
File:KnapsackEmpComplexity.GIF  Source: http://en.wikipedia.org/w/index.php?title=File:KnapsackEmpComplexity.GIF  License: Public Domain  Contributors: Cngoulimis
File:RiemannCriticalLine.svg  Source: http://en.wikipedia.org/w/index.php?title=File:RiemannCriticalLine.svg  License: Public Domain  Contributors: Slonzor
File:Riemann zeta function absolute value.png  Source: http://en.wikipedia.org/w/index.php?title=File:Riemann_zeta_function_absolute_value.png  License: Creative Commons
Attribution-ShareAlike 3.0 Unported  Contributors: Conscious, Kilom691
File:Zeta polar.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Zeta_polar.svg  License: GNU Free Documentation License  Contributors: Original artwork created by Linas Vepstas
<linas@linas.org> User:Linas New smooth and precise plotcurve version by User:Geek3
Image:BSD data plot for elliptic curve 800h1.svg  Source: http://en.wikipedia.org/w/index.php?title=File:BSD_data_plot_for_elliptic_curve_800h1.svg  License: Creative Commons
Attribution-Sharealike 3.0  Contributors: RobHar
License 55

License
Creative Commons Attribution-Share Alike 3.0 Unported
http:/ / creativecommons. org/ licenses/ by-sa/ 3. 0/

Вам также может понравиться