Skip to main content

\(P\mathop{ =}\limits^{?}NP\)

  • Chapter
  • First Online:
Open Problems in Mathematics

Abstract

In 1950, John Nash sent a remarkable letter to the National Security Agency, in which—seeking to build theoretical foundations for cryptography—he all but formulated what today we call the \(\mathsf{P}\mathop{ =}\limits^{?}\mathsf{NP}\) problem, and consider one of the great open problems of science. Here I survey the status of this problem in 2016, for a broad audience of mathematicians, scientists, and engineers. I offer a personal perspective on what it’s about, why it’s important, why it’s reasonable to conjecture that PNP is both true and provable, why proving it is so hard, the landscape of related problems, and crucially, what progress has been made in the last half-century toward solving those problems. The discussion of progress includes diagonalization and circuit lower bounds; the relativization, algebrization, and natural proofs barriers; and the recent works of Ryan Williams and Ketan Mulmuley, which (in different ways) hint at a duality between impossibility proofs and algorithms.

For a more recent version of this survey, with numerous corrections and revisions as well as new material, the reader is invited to visit http://www.scottaaronson.com/papers/pvsnp.pdf

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 149.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 199.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 199.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Here we’re using the observation that, once we fix a formal system (say, first-order logic plus the axioms of ZF set theory), deciding whether a given statement has a proof at most n symbols long in that system is an NP problem, which can therefore be solved in time polynomial in n assuming P = NP. We’re also assuming that the other six Clay conjectures have ZF proofs that are not too enormous: say, 1012 symbols or fewer, depending on the exact running time of the assumed algorithm. In the case of the Poincaré Conjecture, this can almost be taken to be a fact, modulo the translation of Perelman’s proof [179] into the language of ZF.

  2. 2.

    The most celebrated examples of nonconstructive proofs that algorithms exist all come from the Robertson-Seymour graph minors theory, one of the great achievements of twentieth-century combinatorics (for an accessible introduction, see for example Fellows [75]). The Robertson-Seymour theory typically deals with parameterized problems: for example, “given a graph G, decide whether G can be embedded on a sphere with k handles.” In those cases, typically a fast algorithm A k can be abstractly shown to exist for every value of k. The central problem is that each A k requires hard-coded data—in the above example, a finite list of obstructions to the desired embedding—that no one knows how to find given k, and whose size might also grow astronomically as a function of k. On the other hand, once the finite obstruction set for a given k was known, one could then use it to solve the problem for any graph G in time \(O\left (\left \vert G\right \vert ^{3}\right )\), where the constant hidden by the big-O depended on k.

    Robertson-Seymour theory also provides a few examples of non-parameterized problems that are abstractly proved to be in P but with no bound on the exponent, or abstractly proved to be \(O\left (n^{3}\right )\) or \(O\left (n^{2}\right )\) but with no bound on the constant. Thus, one cannot rule out the possibility that the same would happen with an NP-complete problem, and Donald Knuth [131] has explicitly speculated that P = NP will be proven in that way. To me, however, it is unclear whether he speculates this because there is a positive reason for thinking it true, or just because it would be cool and interesting if it was true.

  3. 3.

    As an amusing side note, there is a trick called Levin’s universal search [141], in which one “dovetails” over all Turing machines M 1, M 2,  (that is, for all t, runs M 1, , M t for t steps each), halting when and if any M i has outputs a valid solution to one’s NP search problem. If we know P = NP, then we know this particular algorithm will find a valid solution, whenever one exists, in polynomial time—because clearly some M i does so, and all the machines other than M i increase the total running time by “only” a polynomial factor! With more work, one can even decrease this to a constant factor. Admittedly, however, the polynomial or constant factor will be so enormous as to negate this algorithm’s practical use.

  4. 4.

    The reason for this caveat is that, if a programming language were inherently limited to (say) 64K of memory, there would be only finitely many possible program behaviors, so in principle we could just cache everything in a giant lookup table. Many programming languages do impose a finite upper bound on the addressable memory, but they could easily be generalized to remove this restriction (or one could consider programs that store information on external I/O devices).

  5. 5.

    I should stress that, once we specify which computational models we have in mind—Turing machines, Intel machine code, etc.—the polynomial-time equivalence of those models is typically a theorem, though a rather tedious one. The “thesis” of the Extended Church-Turing Thesis, the part not susceptible to proof, is that all other “reasonable” models of digital computation will also be equivalent to those models.

  6. 6.

    In practice, often one only needs a special kind of Turing reduction called a many-one reduction or Karp reduction, which is a polynomial-time algorithm that maps every yes-instance of \(L^{{\prime}}\) to a yes-instance of L, and every no-instance of \(L^{{\prime}}\) to a no-instance of L. The additional power of Turing reductions—to make multiple queries to the L-oracle (with later queries depending on the outcomes of earlier ones), post-process the results of those queries, etc.—is needed only in a minority of cases. Nevertheless, for conceptual simplicity, throughout this survey I’ll talk in terms of Turing reductions.

  7. 7.

    Technically, Daskalakis et al. showed that the search problem of finding a Nash equilibrium is complete for a complexity class called PPAD. This could be loosely interpreted as saying that the problem is “as close to NP-hard as it could possibly be, subject to Nash’s theorem showing why the decision version is trivial.”

  8. 8.

    In defining the kth level of the hierarchy, we could also have given oracles for Π k−1 P rather than \(\Sigma _{k-1}^{\mathsf{P}}\): it doesn’t matter. Note also that “an oracle for complexity class \(\mathcal{C}\)” should be read as “an oracle for any \(\mathcal{C}\)-complete language L.”

  9. 9.

    This requires one nontrivial result, that every prime number has a succinct certificate—or in other words, that primality testing is in NP [180]. Since 2002, it is even known that primality testing is in P [14].

  10. 10.

    A further surprising result from 1987, called the Immerman-Szelepcsényi Theorem [110, 218], says that \(\mathsf{NSPACE}\left (f\left (n\right )\right ) = \mathsf{coNSPACE}\left (f\left (n\right )\right )\) for every “reasonable” memory bound \(f\left (n\right )\). (By contrast, Savitch’s Theorem produces a quadratic blowup when simulating nondeterministic space by deterministic space, and it remains open whether that blowup can be removed.) This further illustrates how space complexity behaves differently than we expect time complexity to behave.

  11. 11.

    Unlike P or PSPACE, classes like \(\mathsf{TIME}\left (n^{2}\right )\), \(\mathsf{SPACE}\left (n^{3}\right )\), etc. can be sensitive to whether we are talking about Turing machines, RAM machines, or some other model of computation. But in any case, one can simply fix one of those models any time the classes are mentioned in this survey, and nothing will go wrong.

  12. 12.

    I like to joke that, if computer scientists had been physicists, we’d simply have declared PNP to be an observed law of Nature, analogous to the laws of thermodynamics. A Nobel Prize would even be given for the discovery of that law. (And in the unlikely event that someone later proved P = NP, a second Nobel Prize would be awarded for the law’s overthrow.)

  13. 13.

    Strictly speaking, this is for the variant of 3Sat in which every clause must have exactly three literals, rather than at most three.

    Also note that, if we allow the use of randomness, then we can satisfy a 7∕8 fraction of the clauses in expectation by just setting each of the n variables uniformly at random! This is because a clause with three literals has 23 − 1 = 7 ways to be satisfied, and only one way to be unsatisfied. A deterministic polynomial-time algorithm that’s guaranteed to satisfy at least 7∕8 of the clauses requires only a little more work.

  14. 14.

    Indeed, a natural conjecture would be that the problem is NP-hard under randomized reductions for all k ≠ 7, but this remains open (Valiant, personal communication).

  15. 15.

    Note also that, by the Shoenfield absoluteness theorem [202], one’s beliefs about the Axiom of Choice, the Continuum Hypothesis, or other statements proven independent of ZF via forcing can have no effect on the provability of arithmetical statements such as PNP.

  16. 16.

    If a \(\Pi _{1}\)-sentence like the Goldbach Conjecture or the Riemann Hypothesis were known to be independent of ZF, then it would also be known to be true, since any counterexample would have a trivial finite proof! On the other hand, we could also imagine, say, the Goldbach Conjecture being proven equivalent to the consistency of ZF, in which case we could say only that either ZF is consistent and Goldbach is true but ZF doesn’t prove either, or else ZF proves anything. In any case, none of this directly applies to P ≠ NP, which is a \(\Pi _{2}\)-sentence.

  17. 17.

    I won’t have much to say about linear or semidefinite programming in this survey, so perhaps this is as good a place as any to mention that today, we know a great deal about the impossibility of solving NP-complete problems in polynomial time by formulating them as “natural” linear programs. This story starts in 1987, with a preprint by Swart [217] that claimed to prove P = NP by reducing the Traveling Salesperson Problem to a linear program with n 8 variables and constraints. Swart’s preprint inspired a landmark paper by Yannakakis [244] (making it possibly the most productive failed P = NP proof in history!), in which Yannakakis showed that there is no “symmetric” linear program with \(n^{o\left (n\right )}\) variables and constraints that has the “Traveling Salesperson Polytope” as its projection onto a subset of the variables. This ruled out Swart’s approach. Yannakakis also showed that the polytope corresponding to the maximum matching problem has no symmetric LP of subexponential size, but the polytope for the minimum spanning tree problem does have a polynomial-size LP. In general, expressibility by such an LP is sufficient for a problem to be in P, but not necessary.

    Later, in 2012, Fiorini et al. [76] substantially improved Yannakakis’s result, getting rid of the symmetry requirement. There have since been other major results in this direction: in 2014, Rothvoß[194] showed that the perfect matching polytope requires exponentially-large LPs (again with no symmetry requirement), while in 2015, Lee, Raghavendra, and Steurer [140] extended many of these lower bounds from linear to semidefinite programs.

    Collectively, these results rule out one “natural” approach to proving P = NP: namely, to start from famous NP-hard optimization problems like TSP, and then find a polynomial-size linear or semidefinite program that projects onto the polytope whose extreme points are the valid solutions. Of course, we can’t yet rule out the possibility that linear or semidefinite programs could help prove P = NP in some more indirect way (or via some NP-hard problem other than the specific ones that were studied); ruling that out seems essentially tantamount to proving PNP itself.

  18. 18.

    Despite the term “circuit,” which comes from electrical engineering, circuits in theoretical computer science are always free of cycles; they proceed from the inputs to the output via layers of logic gates.

  19. 19.

    This is a rare instance where the non-containment can actually be proved: for example, any unary language (i.e., language of the form \(\left \{0^{n}: n \in S\right \}\)) is clearly in P∕poly, but there is an uncountable infinity of such languages, so almost all of them cannot be in P.

  20. 20.

    Note that, if each logic gate depends on at most 2 inputs, then \(\log _{2}n\) is the smallest depth that allows the output to depend on all n input bits.

  21. 21.

    Bshouty, Cleve, and Eberly [53] showed that the size of the depth-reduced formula can even be taken to be \(O\left (S^{1+\varepsilon }\right )\) , for any constant \(\varepsilon > 0\).

  22. 22.

    But making matters more complicated still, survey propagation fails badly on random 4Sat.

  23. 23.

    We could also allow sampling from some distribution close to \(\mathcal{D}_{n}\), but we will ignore that complication here.

  24. 24.

    There are closely-related objects, such as “lossy” trapdoor OWFs (see [178]), that also suffice for building public-key cryptosystems.

  25. 25.

    At least, not for arbitrary polynomials computed by small formulas or circuits. A great deal of progress has been made derandomizing PIT for restricted classes of polynomials. In fact, the deterministic primality test of Agrawal et al. [14] was based on a derandomization of one extremely special case of PIT.

  26. 26.

    One can also consider the QMA-complete problems, which are a quantum generalization of the NP-complete problems themselves (see [48]), but we will not pursue that here.

  27. 27.

    One can artificially design an NP-complete problem with a superpolynomial quantum speedup over the best known classical algorithm by, for example, taking the language

    $$\displaystyle\begin{array}{rcl} L& =& \left \{0\varphi 0\cdots 0\ \vert \ \varphi \ \text{ is a satisfiable 3SAT instance of size }\ n^{0.01}\right \} \cup {}\\ & &\left \{1x\ \vert \ x\ \text{ is a binary encoding of a positive integer with an odd number of distinct prime factors}\right \}.{}\\ \end{array}$$

    Clearly L is NP-complete, and a quantum algorithm can decide L in \(O\left (c^{n^{0.01} }\right )\) time for some c, whereas the best known classical algorithm will take \(\exp \left (n\right )\) time.

    Conversely, there are also NP-complete problems with no significant quantum speedup known—say, because the best known classical algorithm is based on dynamic programming, and it’s unknown how to combine that with Grover’s algorithm. A candidate example is the Traveling Salesman Problem, which is solvable in \(O\left (2^{n}\mathop{\mathrm{poly}}\nolimits \left (n\right )\right )\) time using the Held-Karp dynamic programming algorithm [107], whereas Grover’s algorithm seems to yield only the worse bound \(O(\sqrt{n!})\).

  28. 28.

    Going even further, Raz [184] proved in 2010 that, if we manage to show that any explicit d-dimensional tensor \(A: \left [n\right ]^{d} \rightarrow \mathbb{F}\) has rank at least \(n^{d\left (1-o\left (1\right )\right )}\), then we’ve also shown that the n × n permanent function has no polynomial-size arithmetic formulas. It’s easy to construct explicit d-dimensional tensors with rank \(n^{\left \lfloor d/2\right \rfloor }\), but the current record is an explicit d-dimensional tensor with rank at least \(2n^{\left \lfloor d/2\right \rfloor } + n - O\left (d\log n\right )\) [18].

    Note that, if we could show that the permanent had no \(n^{O\left (\log n\right )}\)-size arithmetic formulas, that would imply Valiant’s famous Conjecture 66: that the permanent has no polynomial-size arithmetic circuits. However, Raz’s technique seems incapable of proving formula-size lower bounds better than \(n^{\Omega \left (\log \log n\right )}\).

  29. 29.

    With some effort, Shannon’s lower bound can be shown to be tight: that is, every n-variable Boolean function can be represented by a circuit of size \(O\left (2^{n}/n\right )\) . (The obvious upper bound is \(O\left (n2^{n}\right )\).)

  30. 30.

    Crucially, this will be a different language for each k; otherwise we would get \(\mathsf{PSPACE}\not\subset \mathsf{P/poly}\), which is far beyond our current ability to prove.

  31. 31.

    Note that, if we encode the input string using the so-called dual-rail representation—in which every 0 is represented by the 2-bit string 01, and every 1 by 10—then the monotone circuit complexities of Clique, Matching, and so on do become essentially equivalent to their non-monotone circuit complexities, since we can push all the NOT gates to the bottom layer of the circuit using de Morgan’s laws. Unfortunately, Razborov’s lower bound techniques also break down under dual-rail encoding.

  32. 32.

    Assume for simplicity that n is a power of 2. Then x 1 ⊕⋯ ⊕ x n can be written as yz, where y: = x 1 ⊕⋯ ⊕ x n∕2 and z: = x n∕2+1 ⊕⋯ ⊕ x n . This in turn can be written as \(\left (y \wedge \overline{z}\right ) \vee \left (\overline{y} \wedge z\right )\). Expanding recursively now yields a size-n 2 formula for Parity, made of AND, OR, and NOT gates.

  33. 33.

    In theoretical computer science, the term non-negligible means lower-bounded by \(1/n^{O\left (1\right )}\).

  34. 34.

    Technically, the problem of distinguishing random from pseudorandom functions is equivalent to the problem of inverting one-way functions, which is not quite as strong as solving NP-complete problems in polynomial time—only solving average-case NP-complete problems with planted solutions. For more see Sect. 5.3.

  35. 35.

    The problem, given as input the truth table of a Boolean function \(f: \left \{0, 1\right \}^{n} \rightarrow \left \{0, 1\right \}\), of computing or approximating the circuit complexity of f is called the Minimum Circuit Size Problem (MCSP). It is a longstanding open problem whether or not MCSP is NP-hard; at any rate, there are major obstructions to proving it NP-hard with existing techniques (see Kabanets and Cai [117] and Murray and Williams [169]). On the other hand, MCSP cannot be in P (or BPP) unless there are no cryptographically-secure pseudorandom generators. At any rate, what is relevant to natural proofs is just whether there is an efficient algorithm to certify a large fraction of Boolean functions as being hard, which is a weaker requirement than solving MCSP.

  36. 36.

    Allender and Koucký’s paper partly builds on 2003 work by Srinivasan [211], who showed that, to prove PNP, one would “merely” need to show that any algorithm to compute weak approximations for the MaxClique problem takes \(\Omega \left (n^{1+\varepsilon }\right )\) time, for some constant \(\varepsilon > 0\). The way Srinivasan proved this striking statement was, again, by using a sort of self-reduciblity: he showed that, if there’s a polynomial-time algorithm for MaxClique, then by running that algorithm on smaller graphs sampled from the original graph, one can solve approximate versions of MaxClique in \(n^{1+o\left (1\right )}\) time.

  37. 37.

    We can assume without loss of generality that Merlin’s strategy is deterministic, since Merlin is computationally unbounded, and any convex combination of strategies must contain a deterministic strategy that causes Arthur to accept with at least as great a probability as the convex combination does.

  38. 38.

    This is so because a polynomial-space Turing machine can treat the entire interaction between Merlin and Arthur as a game, in which Merlin is trying to get Arthur to accept with the largest possible probability. The machine can then evaluate the exponentially large game tree using depth-first recursion.

  39. 39.

    Actually, for this proof one does not really need either Toda’s Theorem, or the slightly-nontrivial result that \(\mathsf{\varSigma }_{2}^{\mathsf{P}}\) does not have circuits of size n k. Instead, one can just argue directly that at any rate, P #P does not have circuits of size n k, using a slightly more careful version of the argument of Theorem 37. For details see Aaronson [4].

  40. 40.

    In complexity theory, a promise problem is a pair of subsets \(\Pi _{\mathop{\mathrm{YES}}\nolimits },\Pi _{\mathop{\mathrm{NO}}\nolimits } \subseteq \left \{0, 1\right \}^{{\ast}}\) with \(\Pi _{\mathop{\mathrm{YES}}\nolimits } \cap \Pi _{\mathop{\mathrm{NO}}\nolimits } = \varnothing \). An algorithm solves the problem if it accepts all inputs in \(\Pi _{\mathop{\mathrm{YES}}\nolimits }\) and rejects all inputs in \(\Pi _{\mathop{\mathrm{NO}}\nolimits }\). Its behavior on inputs neither in \(\Pi _{\mathop{\mathrm{YES}}\nolimits }\) nor \(\Pi _{\mathop{\mathrm{NO}}\nolimits }\) (i.e., inputs that “violate the promise”) can be arbitrary. A typical example of a promise problem is: given a Boolean circuit C, decide whether C accepts at least 2∕3 of all inputs \(x \in \left \{0, 1\right \}^{n}\) or at most 1∕3 of them, promised that one of those is true. This problem is in BPP (or technically, PromiseBPP). The role of the promise here is to get rid of those inputs for which random sampling would accept with probability between 1∕3 and 2∕3, violating the definition of BPP.

  41. 41.

    Indeed, the hybrid circuit lower bounds of Sect. 6.3.2 could already be considered examples of ironic complexity theory. In this section, we discuss other examples.

  42. 42.

    On unrealistic models such as one-tape Turing machines, one can prove up to \(\Omega \left (n^{2}\right )\) lower bounds for 3Sat and many other problems (even recognizing palindromes), but only by exploiting the fact that the tape head needs to waste a lot of time moving back and forth across the input.

  43. 43.

    On the other hand, by proving size-depth tradeoffs for so-called branching programs, researchers have been able to obtain time-space tradeoffs for certain special problems in P. Unlike the 3Sat tradeoffs, the branching program tradeoffs involve only slightly superlinear time bounds; on the other hand, they really do represent a fundamentally different way to prove time-space tradeoffs, one that makes no appeal to NP-completeness, diagonalization, or hierarchy theorems. As one example, in 2000 Beame et al. [37], building on earlier work by Ajtai [17], used branching programs to prove the following: there exists a problem in P, based on binary quadratic forms, for which any RAM algorithm (even a nonuniform one) that uses \(n^{1-\Omega \left (1\right )}\) space must also use \(\Omega \left (n \cdot \sqrt{\log n/\log \log n}\right )\) time.

  44. 44.

    Interestingly, both the polynomial method and the proof of Lemma 64 are also closely related to the proof of Toda’s Theorem (Theorem 13), that \(\mathsf{PH} \subseteq \mathsf{P}^{\mathsf{\#P}}\).

  45. 45.

    Curiously, this step can only be applied to the ACC circuits themselves, which of course allow OR gates. It cannot be applied to the Boolean functions of low-degree polynomials that one derives from the ACC circuits.

  46. 46.

    Very recently, Kane and Williams [119] managed to give an explicit Boolean function that requires depth-2 threshold circuits with \(\Omega \left (n^{3/2}/\log ^{3}n\right )\) gates. However, their argument does not proceed via a better-than-brute-force algorithm for depth-2 TC 0 Sat; the latter problem appears to remain open.

  47. 47.

    I’ll restrict to fields here for simplicity, but one can also consider (e.g.) rings.

  48. 48.

    To clarify, 0 and 2x are equal as functions over the finite field \(\mathbb{F}_{2}\), but not equal as formal polynomials.

  49. 49.

    The central difference between the \(\mathsf{P}_{\mathbb{R}}\mathop{ =}\limits^{?}\mathsf{NP}_{\mathbb{R}}\) and \(\mathsf{P}_{\mathbb{C}}\mathop{ =}\limits^{?}\mathsf{NP}_{\mathbb{C}}\) questions is simply that, because \(\mathbb{R}\) is an ordered field, one defines Turing machines over \(\mathbb{R}\) to allow comparisons ( < , ≤ ) and branching on their results.

  50. 50.

    In later work, Bürgisser [59] showed that the same conjecture about \(\tau ^{{\ast}}\left (n!\right )\) (called the τ-conjecture) would also imply Valiant’s Conjecture  66, that the permanent has no polynomial-size arithmetic circuits.

  51. 51.

    Fields of characteristic 2, such as \(\mathbb{F}_{2}\) , are a special case: there, the permanent and determinant are equivalent, so in particular \(\mathop{\mathrm{Per}}\nolimits \left (X\right )\) has polynomial-size arithmetic circuits.

  52. 52.

    In the literature, Conjecture 66 is often called the VPVNP conjecture, with VP and VNP being arithmetic analogues of P and NP respectively. I won’t use that terminology in this survey, for several reasons: (1) VP is arguably more analogous to NC than to P, (2) VNP is arguably more analogous to #P than to NP, and (3) Conjecture 66 is almost always studied as a nonuniform conjecture, more analogous to \(\mathsf{NP}\not\subset \mathsf{P/poly}\) than to PNP.

  53. 53.

    Indeed, #P would even have polynomial-size circuits of depth \(\log ^{O\left (1\right )}n\).

  54. 54.

    An immediate corollary is that any multilinear circuit for \(\mathop{\mathrm{Per}}\nolimits \left (X\right )\) or \(\mathop{\mathrm{Det}}\nolimits \left (X\right )\) requires depth \(\Omega \left (\log ^{2}n\right )\).

  55. 55.

    For simplicity, here I’ll assume that we mean an “ordinary” (Boolean) polynomial-time algorithm, though one could also require polynomial-time algorithms in the arithmetic model.

  56. 56.

    Although so far, I haven’t seen a conjectured role for topology or logic in GCT.

  57. 57.

    Furthermore, just as string theory didn’t predict what new data there has been in fundamental physics in recent decades (e.g., the dark energy), so GCT played no role in, e.g., the proof of \(\mathsf{NEXP}\not\subset \mathsf{ACC}\) (Sect. 6.4.2), or the breakthrough lower bounds for small-depth circuits computing the permanent (Sect. 6.5.2). In both cases, to point this out is to hold the theory to a high and probably unfair standard, but also, ipso facto, to pay the theory a compliment.

  58. 58.

    MaxFlow can easily be reduced to linear programming, but Ford and Fulkerson [78] also gave a much faster and direct way to solve it, which can be found in any undergraduate algorithms textbook. There have been further improvements since.

  59. 59.

    We can assume, if we like, that G, s, and t are hardwired into the circuit. We can also allow constants such as 0 and 1 as inputs, but this is not necessary, as we can also generate constants ourselves using comparison gates and arithmetic.

  60. 60.

    More broadly, I’ve found, there are many confusing points about GCT whose resolutions require reminding ourselves that we’re talking about orbit closures, and not only about orbits. For example, the plan of GCT is ultimately to show, roughly speaking, that the n × n permanent has “too little symmetry” to be embedded into the m × m determinant, unless m is much larger than n. But in that case, what about (say) a random arithmetic formula f of size n, which has no nontrivial symmetries, but which clearly can be embedded into the \(\left (n + 1\right ) \times \left (n + 1\right )\) determinant, by Theorem 68? Even though this f clearly isn’t characterized by its symmetries, mustn’t the embedding obstructions for f be a strict superset of the embedding obstructions for the permanent—since f’s symmetries are a strict subset of the permanent’s symmetries—and doesn’t that give rise to a contradiction? According to Mulmuley (personal communication), the solution to the apparent paradox is that this argument would be valid if we were talking only about orbits, but it’s not valid for orbit closures. With orbit closures, the set of obstructions doesn’t depend in a simple way on the symmetries of the original function, so that it’s possible that an obstruction for the permanent would fail to be an obstruction for f.

  61. 61.

    Note that we don’t even need to assume the symmetry \(p\left (X\right ) = p\left (X^{T}\right )\) ; that comes as a free byproduct. Also, it might seem like “cheating” that we use the permanent to state the symmetries that characterize the permanent, and likewise for the determinant. But we’re just using the permanent and determinant as convenient ways to specify which matrices A,B we want, and could give slightly more awkward symmetry conditions that avoided them. (This is especially clear for the permanent, since if A is diagonal, then \(\mathop{\mathrm{Per}}\nolimits \left (A\right )\) is just the product of the diagonal entries.)

  62. 62.

    See Grochow [95, Proposition 3.4.9] for a simple proof of this, via a dimension argument.

  63. 63.

    In principle, \(\rho _{\mathop{\mathrm{Det}}\nolimits }\) and \(\rho _{\mathop{\mathrm{Per}}\nolimits }\) are infinite-dimensional representations, so an algorithm could search them forever for obstructions without halting. On the other hand, if we impose some upper bound on the degrees of the polynomials in the coordinate ring, we get an algorithm that takes “merely” doubly- or triply-exponential time.

  64. 64.

    A very loose analogy: well before Andrew Wiles proved Fermat’s Last Theorem [233], that x n + y n = z n has no nontrivial integer solutions for any n ≥ 3, number theorists knew a reasonably efficient algorithm that took an exponent n as input, and that (in practice, in all the cases that were tried) proved FLT for that n. Using that algorithm, in 1993—just before Wiles announced his proof—Buhler et al. [55] proved FLT for all n up to 4 million.

  65. 65.

    As pointed out in Sect. 2.2.6, there are other cases in complexity theory where deciding positivity is much easier than exact counting: for example, deciding whether a graph has at least one perfect matching (counting the number of perfect matchings is #P-complete).

  66. 66.

    A language L is called P-complete if (1) L ∈ P, and (2) every \(L^{{\prime}}\in \mathsf{P}\) can be reduced to L by some form of reduction weaker than arbitrary polynomial-time ones (LOGSPACE reductions are often used for this purpose).

  67. 67.

    This result of Lipton’s provided the germ of the proof that IP = PSPACE; see Sect. 6.3.1.

    Mulmuley’s test improves over Lipton’s by, for example, requiring only nonadaptive queries to C rather than adaptive ones.

  68. 68.

    A standard example is the 2 × 2 × 2 tensor whose \(\left (2, 1, 1\right )\), \(\left (1, 2, 1\right )\), and \(\left (1, 1, 2\right )\) entries are all 1, and whose 5 remaining entries are all 0. One can check that this tensor has a rank of 3 but border rank of 2.

  69. 69.

    It would be interesting to find a function in P, more natural than the P-complete H function, that’s completely characterized by its symmetries, and then try to understand explicitly why there’s no representation-theoretic obstruction to that function’s orbit closure being contained in χ P —something we already know must be true.

  70. 70.

    In fact, the list can be found in the class ZPP NP, where ZPP stands for Zero-Error Probabilistic Polynomial-Time. This means that, whenever the randomized algorithm succeeds in constructing the list, it’s certain that it’s done so.

  71. 71.

    The polynomial identity testing problem was defined in Sect. 5.4. Also, by a “black-box derandomization,” we mean a deterministic polynomial-time algorithm that outputs a hitting set: that is, a list of points x1 ,…,x such that, for all small arithmetic circuits C that don’t compute the identically-zero polynomial, there exists an i such that \(C\left (x_{i}\right )\neq 0\) . What makes the derandomization “black-box” is that the choice of x 1 ,…,x doesn’t depend on C.

  72. 72.

    In 1999, Allender [19] showed that the permanent, and various other natural #P-complete problems, can’t be solved by LOGTIME -uniform TC 0 circuits: in other words, constant-depth threshold circuits for which there’s an \(O\left (\log n\right )\)-time algorithm to output the ith bit of their description, for any i. Indeed, these problems can’t be solved by LOGTIME-uniform TC 0 circuits of size \(f\left (n\right )\), where f is any function that can yield an exponential when iterated a constant number of times. The proof uses a hierarchy theorem, it would be interesting to know whether it relativizes.

  73. 73.

    Examples are deleting two successive NOT gates, or applying de Morgan’s laws. By the completeness of Boolean algebra, one can give local transformation rules that suffice to convert any n-input Boolean circuit into any equivalent circuit using at most \(\exp \left (n\right )\) moves.

    From talking to experts, this problem seems closely related to the problem of proving superpolynomial lower bounds for so-called Frege proofs, but is possibly easier.

  74. 74.

    There are even what Richard Lipton has termed “galactic algorithms” [147], which beat their asymptotically-worse competitors, but only for values of n that likely exceed the information storage capacity of the galaxy or the observable universe. The currently-fastest matrix multiplication algorithms might fall into this category, although the constants don’t seem to be known in enough detail to say for sure.

References

  1. S. Aaronson. Is P versus NP formally independent? Bulletin of the EATCS, (81), October 2003.

    Google Scholar 

  2. S. Aaronson. Multilinear formulas and skepticism of quantum computing. In Proc. ACM STOC, pages 118–127, 2004. quant-ph/0311039, www.scottaaronson.com/papers/mlinsiam.pdf.

  3. S. Aaronson. NP-complete problems and physical reality. SIGACT News, March 2005. quant-ph/0502072.

    Google Scholar 

  4. S. Aaronson. Oracles are subtle but not malicious. In Proc. Conference on Computational Complexity, pages 340–354, 2006. ECCC TR05-040.

    Google Scholar 

  5. S. Aaronson. Arithmetic natural proofs theory is sought, 2008. www.scottaaronson.com/blog/?p=336.

    Google Scholar 

  6. S. Aaronson. Quantum Computing Since Democritus. Cambridge University Press, 2013.

    Book  MATH  Google Scholar 

  7. S. Aaronson. The scientific case for PNP, 2014. www.scottaaronson.com/blog/?p=1720.

  8. S. Aaronson et al. The Complexity Zoo. www.complexityzoo.com.

  9. S. Aaronson, R. Impagliazzo, and D. Moshkovitz. AM with multiple Merlins. In Proc. Conference on Computational Complexity, pages 44–55, 2014. arXiv:1401.6848.

    Google Scholar 

  10. S. Aaronson and A. Wigderson. Algebrization: a new barrier in complexity theory. ACM Trans. on Computation Theory, 1(1), 2009. Earlier version in Proc. ACM STOC’2008.

    Google Scholar 

  11. L. Adleman. Two theorems on random polynomial time. In Proc. IEEE FOCS, pages 75–83, 1978.

    Google Scholar 

  12. L. Adleman, J. DeMarrais, and M.-D. Huang. Quantum computability. SIAM J. Comput., 26(5):1524–1540, 1997.

    Article  MathSciNet  MATH  Google Scholar 

  13. M. Agrawal. Determinant versus permanent. In Proceedings of the International Congress of Mathematicians, 2006.

    Google Scholar 

  14. M. Agrawal, N. Kayal, and N. Saxena. PRIMES is in P. Annals of Mathematics, 160(2):781–793, 2004. Preprint released in 2002.

    Google Scholar 

  15. M. Agrawal and V. Vinay. Arithmetic circuits: a chasm at depth four. In Proc. IEEE FOCS, pages 67–75, 2008.

    Google Scholar 

  16. M. Ajtai. \(\Sigma _{1}^{1}\)-formulae on finite structures. Annals of Pure and Applied Logic, 24(1):1–48, 1983.

    Google Scholar 

  17. M. Ajtai. A non-linear time lower bound for Boolean branching programs. Theory of Computing, 1(1):149–176, 2005. Earlier version in Proc. IEEE FOCS’1999, pp. 60–70.

    Google Scholar 

  18. B. Alexeev, M. A. Forbes, and J. Tsimerman. Tensor rank: some lower and upper bounds. In Proc. Conference on Computational Complexity, pages 283–291, 2011.

    Google Scholar 

  19. E. Allender. The permanent requires large uniform threshold circuits. Chicago Journal of Theoretical Computer Science, 7:19, 1999.

    MathSciNet  MATH  Google Scholar 

  20. E. Allender. Cracks in the defenses: scouting out approaches on circuit lower bounds. In Computer Science in Russia, pages 3–10, 2008.

    Google Scholar 

  21. E. Allender. A status report on the P versus NP question. Advances in Computers, 77:117–147, 2009.

    Article  Google Scholar 

  22. E. Allender and V. Gore. On strong separations from AC 0. In Fundamentals of Computation Theory, pages 1–15. Springer Berlin Heidelberg, 1991.

    Google Scholar 

  23. E. Allender and V. Gore. A uniform circuit lower bound for the permanent. SIAM J. Comput., 23(5):1026–1049, 1994.

    Article  MathSciNet  MATH  Google Scholar 

  24. E. Allender and M. Koucký. Amplifying lower bounds by means of self-reducibility. J. of the ACM, 57(3):1–36, 2010. Earlier version in Proc. IEEE Complexity’2008, pp. 31–40.

    Google Scholar 

  25. N. Alon and R. B. Boppana. The monotone circuit complexity of Boolean functions. Combinatorica, 7(1):1–22, 1987.

    Article  MathSciNet  MATH  Google Scholar 

  26. A. E. Andreev. On a method for obtaining more than quadratic effective lower bounds for the complexity of π-schemes. Moscow Univ. Math. Bull., 42:63–66, 1987. In Russian.

    MATH  Google Scholar 

  27. S. Arora and B. Barak. Complexity Theory: A Modern Approach. Cambridge University Press, 2009. Online draft at www.cs.princeton.edu/theory/complexity/.

  28. S. Arora, R. Impagliazzo, and U. Vazirani. Relativizing versus nonrelativizing techniques: the role of local checkability. Manuscript, 1992.

    Google Scholar 

  29. S. Arora, C. Lund, R. Motwani, M. Sudan, and M. Szegedy. Proof verification and the hardness of approximation problems. J. of the ACM, 45(3):501–555, 1998. Earlier version in Proc. IEEE FOCS’1992, pp. 14–23.

    Google Scholar 

  30. S. Arora and S. Safra. Probabilistic checking of proofs: a new characterization of NP. J. of the ACM, 45(1):70–122, 1998. Earlier version in Proc. IEEE FOCS’1992, pp. 2–13.

    Google Scholar 

  31. A. Atserias. Distinguishing SAT from polynomial-size circuits, through black-box queries. In Proc. Conference on Computational Complexity, pages 88–95, 2006.

    Google Scholar 

  32. L. Babai. Graph isomorphism in quasipolynomial time. arXiv:1512.03547, 2015.

    Google Scholar 

  33. A. Backurs and P. Indyk. Edit distance cannot be computed in strongly subquadratic time (unless SETH is false). In Proc. ACM STOC, pages 51–58, 2015.

    Google Scholar 

  34. T. Baker, J. Gill, and R. Solovay. Relativizations of the P=?NP question. SIAM J. Comput., 4:431–442, 1975.

    Article  MathSciNet  MATH  Google Scholar 

  35. A. Banerjee, C. Peikert, and A. Rosen. Pseudorandom functions and lattices. In Proc. of EUROCRYPT, pages 719–737, 2012.

    Google Scholar 

  36. P. Beame and T. Pitassi. Propositional proof complexity: past, present, and future. Current Trends in Theoretical Computer Science, pages 42–70, 2001.

    Google Scholar 

  37. P. Beame, M. E. Saks, X. Sun, and E. Vee. Time-space trade-off lower bounds for randomized computation of decision problems. J. of the ACM, 50(2):154–195, 2003. Earlier version in Proc. IEEE FOCS’2000, pp. 169–179.

    Google Scholar 

  38. R. Beigel and J. Tarui. On ACC. Computational Complexity, 4:350–366, 1994. Earlier version in Proc. IEEE FOCS’1991, pp. 783–792.

    Google Scholar 

  39. S. Ben-David and S. Halevi. On the independence of P versus NP. Technical Report TR714, Technion, 1992.

    Google Scholar 

  40. C. Bennett, E. Bernstein, G. Brassard, and U. Vazirani. Strengths and weaknesses of quantum computing. SIAM J. Comput., 26(5):1510–1523, 1997. quant-ph/9701001.

    Google Scholar 

  41. C. H. Bennett and J. Gill. Relative to a random oracle A, P ANP AcoNP A with probability 1. SIAM J. Comput., 10(1):96–113, 1981.

    Article  MathSciNet  MATH  Google Scholar 

  42. E. Bernstein and U. Vazirani. Quantum complexity theory. SIAM J. Comput., 26(5):1411–1473, 1997. Earlier version in Proc. ACM STOC’1993.

    Google Scholar 

  43. D. Bini. Relations between exact and approximate bilinear algorithms. applications. Calcolo, 17(1):87–97, 1980.

    Google Scholar 

  44. D. Bini, G. Lotti, and F. Romani. Approximate solutions for the bilinear form computational problem. SIAM J. Comput., 9(4):692–697, 1980.

    Article  MathSciNet  MATH  Google Scholar 

  45. J. Blasiak, K. Mulmuley, and M. Sohoni. Geometric complexity theory IV: nonstandard quantum group for the Kronecker problem, volume 235 of Memoirs of the American Mathematical Society. 2015. arXiv:cs.CC/0703110.

    Google Scholar 

  46. L. Blum, F. Cucker, M. Shub, and S. Smale. Complexity and Real Computation. Springer-Verlag, 1997.

    MATH  Google Scholar 

  47. A. Bogdanov and L. Trevisan. Average-case complexity. Foundations and Trends in Theoretical Computer Science, 2(1), 2006. ECCC TR06-073.

    Google Scholar 

  48. A. D. Bookatz. QMA-complete problems. Quantum Information and Computation, 14(5-6):361–383, 2014. arXiv:1212.6312.

    Google Scholar 

  49. R. B. Boppana, J. Håstad, and S. Zachos. Does co-NP have short interactive proofs? Inform. Proc. Lett., 25:127–132, 1987.

    Article  MathSciNet  MATH  Google Scholar 

  50. A. Braunstein, M. Mézard, and R. Zecchina. Survey propagation: An algorithm for satisfiability. Random Structures and Algorithms, 27(2):201–226, 2005.

    Article  MathSciNet  MATH  Google Scholar 

  51. M. Braverman. Poly-logarithmic independence fools AC 0 circuits. In Proc. Conference on Computational Complexity, pages 3–8, 2009. ECCC TR09-011.

    Google Scholar 

  52. R. P. Brent. The parallel evaluation of general arithmetic expressions. J. of the ACM, 21:201–206, 1974.

    Article  MathSciNet  MATH  Google Scholar 

  53. N. H. Bshouty, R. Cleve, and W. Eberly. Size-depth tradeoffs for algebraic formulae. SIAM J. Comput., 24(4):682–705, 1995. Earlier version in Proc. IEEE FOCS’1991, pp. 334–341.

    Google Scholar 

  54. N. H. Bshouty, R. Cleve, R. Gavaldà, S. Kannan, and C. Tamon. Oracles and queries that are sufficient for exact learning. J. Comput. Sys. Sci., 52(3):421–433, 1996.

    Article  MathSciNet  MATH  Google Scholar 

  55. J. Buhler, R. Crandall, R. Ernvall, and T. Metsänkylä. Irregular primes and cyclotomic invariants to four million. Mathematics of Computation, 61(203):151–153, 1993.

    Article  MathSciNet  MATH  Google Scholar 

  56. H. Buhrman, L. Fortnow, and T. Thierauf. Nonrelativizing separations. In Proc. Conference on Computational Complexity, pages 8–12, 1998.

    Google Scholar 

  57. P. Bürgisser. Completeness and reduction in algebraic complexity theory. 2000. Available at math-www.uni-paderborn.de/agpb/work/habil.ps.

    Google Scholar 

  58. P. Bürgisser. Cook’s versus Valiant’s hypothesis. Theoretical Comput. Sci., 235(1):71–88, 2000.

    Article  MATH  Google Scholar 

  59. P. Bürgisser. On defining integers and proving arithmetic circuit lower bounds. Computational Complexity, 18(1):81–103, 2009. Earlier version in Proc. STACS’2007, pp. 133–144.

    Google Scholar 

  60. P. Bürgisser and C. Ikenmeyer. Deciding positivity of Littlewood-Richardson coefficients. SIAM J. Discrete Math., 27(4):1639–1681, 2013.

    Article  MathSciNet  MATH  Google Scholar 

  61. P. Bürgisser and C. Ikenmeyer. Explicit lower bounds via geometric complexity theory. In Proc. ACM STOC, pages 141–150, 2013. arXiv:1210.8368.

    Google Scholar 

  62. P. Bürgisser, C. Ikenmeyer, and J. Hüttenhain. Permanent versus determinant: not via saturations. arXiv:1501.05528, 2015.

    Google Scholar 

  63. S. R. Buss and R. Williams. Limits on alternation trading proofs for time-space lower bounds. Computational Complexity, 24(3):533–600, 2015. Earlier version in Proc. IEEE Complexity’2012, pp. 181–191.

    Google Scholar 

  64. J.-Y. Cai, X. Chen, and D. Li. Quadratic lower bound for permanent vs. determinant in any characteristic. 19(1):37–56, 2010. Earlier version in Proc. ACM STOC’2008, pp. 491–498.

    Google Scholar 

  65. A. Cobham. The intrinsic computational difficulty of functions. In Proceedings of Logic, Methodology, and Philosophy of Science II. North Holland, 1965.

    Google Scholar 

  66. S. Cook. The P versus NP problem, 2000. Clay Math Institute official problem description. At www.claymath.org/sites/default/files/pvsnp.pdf.

  67. S. A. Cook. The complexity of theorem-proving procedures. In Proc. ACM STOC, pages 151–158, 1971.

    Google Scholar 

  68. S. A. Cook. A hierarchy for nondeterministic time complexity. In Proc. ACM STOC, pages 187–192, 1972.

    Google Scholar 

  69. D. Coppersmith. Rapid multiplication of rectangular matrices. SIAM J. Comput., 11(3):467–471, 1982.

    Article  MathSciNet  MATH  Google Scholar 

  70. D. Coppersmith and S. Winograd. Matrix multiplication via arithmetic progressions. Journal of Symbolic Computation, 9(3):251–280, 1990. Earlier version in Proc. ACM STOC’1987.

    Google Scholar 

  71. C. Daskalakis, P. W. Goldberg, and C. H. Papadimitriou. The complexity of computing a Nash equilibrium. Commun. of the ACM, 52(2):89–97, 2009. Earlier version in Proc. ACM STOC’2006.

    Google Scholar 

  72. M. Davis, G. Logemann, and D. Loveland. A machine program for theorem proving. Commun. of the ACM, 5(7):394–397, 1962.

    Article  MathSciNet  MATH  Google Scholar 

  73. V. Deolalikar. PNP. Archived version available at www.win.tue.nl/~gwoegi/P-versus-NP/Deolalikar.pdf, 2010.

  74. J. Edmonds. Paths, trees, and flowers. Canadian Journal of Mathematics, 17(3):449–467, 1965.

    Article  MathSciNet  MATH  Google Scholar 

  75. M. R. Fellows. The Robertson-Seymour theorems: a survey of applications. Contemporary Mathematics, 89:1–18, 1989.

    Article  MathSciNet  MATH  Google Scholar 

  76. S. Fiorini, S. Massar, S. Pokutta, H. R. Tiwary, and R. de Wolf. Exponential lower bounds for polytopes in combinatorial optimization. J. of the ACM, 62(2):17, 2015. Earlier version in Proc. ACM STOC’2012, pp. 95–106.

    Google Scholar 

  77. M. A. Forbes and A. Shpilka. Explicit Noether normalization for simultaneous conjugation via polynomial identity testing. pages 527–542, 2013. arXiv:1303.0084.

    Google Scholar 

  78. L. R. Ford and D. R. Fulkerson. Flows in Networks. Princeton, 1962.

    MATH  Google Scholar 

  79. L. Fortnow. The status of the P versus NP problem. Commun. of the ACM, 52(9):78–86, 2009.

    Article  Google Scholar 

  80. L. Fortnow. The Golden Ticket: P, NP, and the Search for the Impossible. Princeton University Press, 2009.

    MATH  Google Scholar 

  81. L. Fortnow and D. van Melkebeek. Time-space tradeoffs for nondeterministic computation. In Proc. Conference on Computational Complexity, pages 2–13, 2000.

    Google Scholar 

  82. L. Fortnow, A. Pavan, and S. Sengupta. Proving SAT does not have small circuits with an application to the two queries problem. J. Comput. Sys. Sci., 74(3):358–363, 2008. Earlier version in Proc. IEEE Complexity’2003, pp. 347–350.

    Google Scholar 

  83. L. Fortnow and M. Sipser. Are there interactive protocols for co-NP languages? Inform. Proc. Lett., 28:249–251, 1988.

    Article  MathSciNet  MATH  Google Scholar 

  84. H. M. Friedman. Mathematically natural concrete incompleteness. At u.osu.edu/friedman.8/ files/2014/01/Putnam062115pdf-15ku867.pdf, 2015.

    Google Scholar 

  85. M. Furst, J. B. Saxe, and M. Sipser. Parity, circuits, and the polynomial time hierarchy. Math. Systems Theory, 17:13–27, 1984. Earlier version in Proc. IEEE FOCS’1981, pp. 260–270.

    Google Scholar 

  86. M. R. Garey and D. S. Johnson. Computers and Intractability: A Guide to the Theory of NP-Completeness. W. H. Freeman, 1979.

    MATH  Google Scholar 

  87. W. Gasarch. The P=?NP poll. SIGACT News, 33(2):34–47, June 2002.

    Google Scholar 

  88. O. Goldreich, S. Goldwasser, and S. Micali. How to construct random functions. J. of the ACM, 33(4):792–807, 1984. Earlier version in Proc. IEEE FOCS’1984, pp. 464–479.

    Google Scholar 

  89. O. Goldreich, S. Micali, and A. Wigderson. Proofs that yield nothing but their validity or all languages in NP have zero-knowledge proof systems. J. of the ACM, 38(1):691–729, 1991.

    MathSciNet  MATH  Google Scholar 

  90. S. Goldwasser and M. Sipser. Private coins versus public coins in interactive proof systems. In Randomness and Computation, volume 5 of Advances in Computing Research. JAI Press, 1989.

    Google Scholar 

  91. R. Goodstein. On the restricted ordinal theorem. J. Symbolic Logic, 9:33–41, 1944.

    Article  MathSciNet  MATH  Google Scholar 

  92. B. Grenet. An upper bound for the permanent versus determinant problem. www.lirmm.fr/~grenet/publis/Gre11.pdf, 2011.

  93. D. Grigoriev and M. Karpinski. An exponential lower bound for depth 3 arithmetic circuits. In Proc. ACM STOC, pages 577–582, 1998.

    Google Scholar 

  94. D. Grigoriev and A. A. Razborov. Exponential lower bounds for depth 3 arithmetic circuits in algebras of functions over finite fields. Appl. Algebra Eng. Commun. Comput., 10(6):465–487, 2000. Earlier version in Proc. IEEE FOCS’1998, pp. 269–278.

    Google Scholar 

  95. J. A. Grochow. Symmetry and equivalence relations in classical and geometric complexity theory. PhD thesis, 2012.

    Google Scholar 

  96. J. A. Grochow. Unifying known lower bounds via Geometric Complexity Theory. Computational Complexity, 24(2):393–475, 2015. Earlier version in Proc. IEEE Complexity’2014, pp. 274–285.

    Google Scholar 

  97. L. K. Grover. A fast quantum mechanical algorithm for database search. In Proc. ACM STOC, pages 212–219, 1996. quant-ph/9605043.

    Google Scholar 

  98. A. Gupta, P. Kamath, N. Kayal, and R. Saptharishi. Arithmetic circuits: a chasm at depth three. In Proc. IEEE FOCS, pages 578–587, 2013.

    Google Scholar 

  99. A. Gupta, P. Kamath, N. Kayal, and R. Saptharishi. Approaching the chasm at depth four. J. of the ACM, 61(6):1–16, 2014. Earlier version in Proc. IEEE Complexity’2013, pp. 65–73.

    Google Scholar 

  100. L. Gurvits. On the complexity of mixed discriminants and related problems. In Mathematical Foundations of Computer Science, pages 447–458, 2005.

    Google Scholar 

  101. A. Haken. The intractability of resolution. Theoretical Comput. Sci., 39:297–308, 1985.

    Article  MathSciNet  MATH  Google Scholar 

  102. J. Hartmanis and R. E. Stearns. On the computational complexity of algorithms. Transactions of the American Mathematical Society, 117:285–306, 1965.

    Article  MathSciNet  MATH  Google Scholar 

  103. J. Håstad. Computational Limitations for Small Depth Circuits. MIT Press, 1987.

    Google Scholar 

  104. J. Håstad. The shrinkage exponent of De Morgan formulas is 2. SIAM J. Comput., 27(1):48–64, 1998. Earlier version in Proc. IEEE FOCS’1993, pp. 114–123.

    Google Scholar 

  105. J. Håstad. Some optimal inapproximability results. J. of the ACM, 48:798–859, 2001. Earlier version in Proc. ACM STOC’1997, pp. 1–10.

    Google Scholar 

  106. J. Håstad, R. Impagliazzo, L. A. Levin, and M. Luby. A pseudorandom generator from any one-way function. SIAM J. Comput., 28(4):1364–1396, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  107. M. Held and R. M. Karp. A dynamic programming approach to sequencing problems. Journal of the Society for Industrial and Applied Mathematics, 10(1):196–210, 1962.

    Article  MathSciNet  MATH  Google Scholar 

  108. C. Ikenmeyer, K. Mulmuley, and M. Walter. On vanishing of Kronecker coefficients. arXiv:1507.02955, 2015.

    Google Scholar 

  109. C. Ikenmeyer and G. Panova. Rectangular Kronecker coefficients and plethysms in geometric complexity theory. arXiv:1512.03798, 2015.

    Google Scholar 

  110. N. Immerman. Nondeterministic space is closed under complementation. SIAM J. Comput., 17(5):935–938, 1988. Earlier version in Proc. IEEE Structure in Complexity Theory, 1988.

    Google Scholar 

  111. N. Immerman. Descriptive Complexity. Springer, 1998.

    MATH  Google Scholar 

  112. R. Impagliazzo, V. Kabanets, and A. Kolokolova. An axiomatic approach to algebrization. In Proc. ACM STOC, pages 695–704, 2009.

    Google Scholar 

  113. R. Impagliazzo, V. Kabanets, and A. Wigderson. In search of an easy witness: exponential time vs. probabilistic polynomial time. J. Comput. Sys. Sci., 65(4):672–694, 2002. Earlier version in Proc. IEEE Complexity’2001, pp. 2–12.

    Google Scholar 

  114. R. Impagliazzo and N. Nisan. The effect of random restrictions on formula size. Random Structures and Algorithms, 4(2):121–134, 1993.

    Article  MathSciNet  MATH  Google Scholar 

  115. R. Impagliazzo and A. Wigderson. P=BPP unless E has subexponential circuits: derandomizing the XOR Lemma. In Proc. ACM STOC, pages 220–229, 1997.

    Google Scholar 

  116. M. Jerrum and M. Snir. Some exact complexity results for straight-line computations over semirings. J. of the ACM, 29(3):874–897, 1982.

    Article  MathSciNet  MATH  Google Scholar 

  117. V. Kabanets and J.-Y. Cai. Circuit minimization problem. In Proc. ACM STOC, pages 73–79, 2000. TR99-045.

    Google Scholar 

  118. V. Kabanets and R. Impagliazzo. Derandomizing polynomial identity testing means proving circuit lower bounds. Computational Complexity, 13(1-2):1–46, 2004. Earlier version in Proc. ACM STOC’2003. ECCC TR02-055.

    Google Scholar 

  119. D. M. Kane and R. Williams. Super-linear gate and super-quadratic wire lower bounds for depth-two and depth-three threshold circuits. arXiv:1511.07860, 2015.

    Google Scholar 

  120. R. Kannan. Circuit-size lower bounds and non-reducibility to sparse sets. Information and Control, 55:40–56, 1982. Earlier version in Proc. IEEE FOCS’1981, pp. 304–309.

    Google Scholar 

  121. M. Karchmer and A. Wigderson. Monotone circuits for connectivity require super-logarithmic depth. SIAM J. Comput., 3:255–265, 1990. Earlier version in Proc. ACM STOC’1988, pp. 539–550.

    Google Scholar 

  122. R. M. Karp. Reducibility among combinatorial problems. In R. E. Miller and J. W. Thatcher, editors, Complexity of Computer Computations, pages 85–103. Plenum Press, 1972.

    Google Scholar 

  123. R. M. Karp and R. J. Lipton. Turing machines that take advice. Enseign. Math., 28:191–201, 1982. Earlier version in Proc. ACM STOC’1980, pp. 302–309.

    Google Scholar 

  124. N. Kayal. Affine projections of polynomials: extended abstract. In Proc. ACM STOC, pages 643–662, 2012.

    Google Scholar 

  125. N. Kayal, N. Limaye, C. Saha, and S. Srinivasan. An exponential lower bound for homogeneous depth four arithmetic formulas. In Proc. IEEE FOCS, pages 61–70, 2014.

    Google Scholar 

  126. N. Kayal, C. Saha, and R. Saptharishi. A super-polynomial lower bound for regular arithmetic formulas. In Proc. ACM STOC, pages 146–153, 2014.

    Google Scholar 

  127. S. Khot. On the Unique Games Conjecture. In Proc. Conference on Computational Complexity, pages 99–121, 2010.

    Google Scholar 

  128. V. M. Khrapchenko. A method of determining lower bounds for the complexity of π schemes. Matematischi Zametki, 10:83–92, 1971. In Russian.

    MATH  Google Scholar 

  129. L. Kirby and J. Paris. Accessible independence results for Peano arithmetic. Bulletin of the London Mathematical Society, 14:285–293, 1982.

    Article  MathSciNet  MATH  Google Scholar 

  130. A. Klivans and D. van Melkebeek. Graph nonisomorphism has subexponential size proofs unless the polynomial-time hierarchy collapses. SIAM J. Comput., 31:1501–1526, 2002. Earlier version in Proc. ACM STOC’1999.

    Google Scholar 

  131. D. E. Knuth and E. G. Daylight. Algorithmic Barriers Falling: P=NP? Lonely Scholar, 2014.

    MATH  Google Scholar 

  132. A. Knutson and T. Tao. The honeycomb model of \(GL_{n}(\mathbb{C})\) tensor products I: proof of the saturation conjecture. J. Amer. Math. Soc., 12(4):1055–1090, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  133. P. Koiran. Arithmetic circuits: the chasm at depth four gets wider. Theor. Comput. Sci., 448:56–65, 2012.

    Article  MathSciNet  MATH  Google Scholar 

  134. E. Kushilevitz and N. Nisan. Communication Complexity. Cambridge, 1997.

    Book  MATH  Google Scholar 

  135. R. E. Ladner. On the structure of polynomial time reducibility. J. of the ACM, 22:155–171, 1975.

    Article  MathSciNet  MATH  Google Scholar 

  136. J. M. Landsberg. The border rank of the multiplication of two by two matrices is seven. J. Amer. Math. Soc., 19(2):447–459, 2006. arXiv:math/0407224.

    Google Scholar 

  137. J. M. Landsberg. Geometric complexity theory: an introduction for geometers. Annali dell’Universita di Ferrara, 61(1):65–117, 2015. arXiv:1305.7387.

    Google Scholar 

  138. J. M. Landsberg and G. Ottaviani. New lower bounds for the border rank of matrix multiplication. Theory of Computing, 11:285–298, 2015. arXiv:1112.6007.

    Google Scholar 

  139. C. Lautemann. BPP and the polynomial hierarchy. Inform. Proc. Lett., 17:215–217, 1983.

    Article  MathSciNet  MATH  Google Scholar 

  140. J. R. Lee, P. Raghavendra, and D. Steurer. Lower bounds on the size of semidefinite programming relaxations. In Proc. ACM STOC, pages 567–576, 2015.

    Google Scholar 

  141. L. A. Levin. Universal sequential search problems. Problems of Information Transmission, 9(3):115–116, 1973.

    MATH  Google Scholar 

  142. L. A. Levin. Laws of information conservation (non-growth) and aspects of the foundations of probability theory. Problems of Information Transmission, 10(3):206–210, 1974.

    Google Scholar 

  143. M. Li and P. M. B. Vitányi. Average case complexity under the universal distribution equals worst-case complexity. Inform. Proc. Lett., 42(3):145–149, 1992.

    Article  MathSciNet  MATH  Google Scholar 

  144. N. Linial, Y. Mansour, and N. Nisan. Constant depth circuits, Fourier transform, and learnability. J. of the ACM, 40(3):607–620, 1993. Earlier version in Proc. IEEE FOCS’1989, pp. 574–579.

    Google Scholar 

  145. N. Linial and N. Nisan. Approximate inclusion-exclusion. Combinatorica, 10(4):349–365, 1990. Earlier version in Proc. ACM STOC’1990.

    Google Scholar 

  146. R. J. Lipton. New directions in testing. In Distributed Computing and Cryptography, pages 191–202. AMS, 1991.

    Google Scholar 

  147. R. J. Lipton. Galactic algorithms, 2010. rjlipton. wordpress.com/2010/10/23/galactic-algorithms/.

    Google Scholar 

  148. R. J. Lipton and K. W. Regan. Practically P=NP?, 2014. rjlipton.wordpress.com/2014/02/28/practically-pnp/.

    Google Scholar 

  149. R. J. Lipton and A. Viglas. On the complexity of SAT. In Proc. IEEE FOCS, pages 459–464, 1999.

    Google Scholar 

  150. D. Luna. Slices étales. Mémoires de la Société Mathématique de France, 33:81–105, 1973.

    MATH  Google Scholar 

  151. C. Lund, L. Fortnow, H. Karloff, and N. Nisan. Algebraic methods for interactive proof systems. J. of the ACM, 39:859–868, 1992. Earlier version in Proc. IEEE FOCS’1990, pp. 2–10.

    Google Scholar 

  152. D. van Melkebeek. A survey of lower bounds for satisfiability and related problems. Foundations and Trends in Theoretical Computer Science, 2:197–303, 2007. ECCC TR07-099.

    Google Scholar 

  153. T. Mignon and N. Ressayre. A quadratic bound for the determinant and permanent problem. International Mathematics Research Notices, (79):4241–4253, 2004.

    Article  MathSciNet  MATH  Google Scholar 

  154. G. L. Miller. Riemann’s hypothesis and tests for primality. J. Comput. Sys. Sci., 13:300–317, 1976. Earlier version in Proc. ACM STOC’1975.

    Google Scholar 

  155. C. Moore and S. Mertens. The Nature of Computation. Oxford University Press, 2011.

    Book  MATH  Google Scholar 

  156. S. Moran. Some results on relativized deterministic and nondeterministic time hierarchies. J. Comput. Sys. Sci., 22(1):1–8, 1981.

    Article  MathSciNet  MATH  Google Scholar 

  157. K. Mulmuley. GCT publications web page. ramakrishnadas.cs.uchicago.edu.

    Google Scholar 

  158. K. Mulmuley. Lower bounds in a parallel model without bit operations. SIAM J. Comput., 28(4):1460–1509, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  159. K. Mulmuley. Geometric complexity theory VII: nonstandard quantum group for the plethysm problem. Technical Report TR-2007-14, University of Chicago, 2007. arXiv:0709.0749.

    Google Scholar 

  160. K. Mulmuley. Geometric complexity theory VIII: on canonical bases for the nonstandard quantum groups. Technical Report TR-2007-15, University of Chicago, 2007. arXiv:0709.0751.

    Google Scholar 

  161. K. Mulmuley. Explicit proofs and the flip. arXiv:1009.0246, 2010.

    Google Scholar 

  162. K. Mulmuley. Geometric complexity theory VI: the flip via positivity. Technical report, University of Chicago, 2011. arXiv:0704.0229.

    Google Scholar 

  163. K. Mulmuley. On P vs. NP and geometric complexity theory: dedicated to Sri Ramakrishna. J. of the ACM, 58(2):5, 2011.

    Google Scholar 

  164. K. Mulmuley. The GCT program toward the P vs. NP problem. Commun. of the ACM, 55(6):98–107, June 2012.

    Google Scholar 

  165. K. Mulmuley. Geometric complexity theory V: equivalence between blackbox derandomization of polynomial identity testing and derandomization of Noether’s Normalization Lemma. In Proc. IEEE FOCS, pages 629–638, 2012. Full version available at arXiv:1209.5993.

    Google Scholar 

  166. K. Mulmuley, H. Narayanan, and M. Sohoni. Geometric complexity theory III: on deciding nonvanishing of a Littlewood-Richardson coefficient. Journal of Algebraic Combinatorics, 36(1):103–110, 2012.

    Article  MathSciNet  MATH  Google Scholar 

  167. K. Mulmuley and M. Sohoni. Geometric complexity theory I: An approach to the P vs. NP and related problems. SIAM J. Comput., 31(2):496–526, 2001.

    Google Scholar 

  168. K. Mulmuley and M. Sohoni. Geometric complexity theory II: Towards explicit obstructions for embeddings among class varieties. SIAM J. Comput., 38(3):1175–1206, 2008.

    Article  MathSciNet  MATH  Google Scholar 

  169. C. D. Murray and R. R. Williams. On the (non) NP-hardness of computing circuit complexity. In Proc. Conference on Computational Complexity, pages 365–380, 2015.

    Google Scholar 

  170. J. Naor and M. Naor. Number-theoretic constructions of efficient pseudo-random functions. J. of the ACM, 51(2):231–262, 2004. Earlier version in Proc. IEEE FOCS’1997.

    Google Scholar 

  171. J. Nash. Letter to the United States National Security Agency, 1950. Available at www.nsa.gov/public_info/_files/nash_letters/nash_letters1.pdf.

    Google Scholar 

  172. M. Nielsen and I. Chuang. Quantum Computation and Quantum Information. Cambridge University Press, 2000.

    MATH  Google Scholar 

  173. N. Nisan and A. Wigderson. Lower bounds on arithmetic circuits via partial derivatives. Computational Complexity, 6(3):217–234, 1997. Earlier version in Proc. IEEE FOCS’1995, pp. 16–25.

    Google Scholar 

  174. B. Aydınlıoğlu and E. Bach. Affine relativization: unifying the algebrization and relativization barriers. 2016.

    Google Scholar 

  175. C. H. Papadimitriou. Computational Complexity. Addison-Wesley, 1994.

    MATH  Google Scholar 

  176. M. Paterson and U. Zwick. Shrinkage of De Morgan formulae under restriction. Random Structures and Algorithms, 4(2):135–150, 1993.

    Article  MathSciNet  MATH  Google Scholar 

  177. W. J. Paul, N. Pippenger, E. Szemerédi, and W. T. Trotter. On determinism versus non-determinism and related problems. In Proc. IEEE FOCS, pages 429–438, 1983.

    Google Scholar 

  178. C. Peikert and B. Waters. Lossy trapdoor functions and their applications. SIAM J. Comput., 40(6):1803–1844, 2011. Earlier version in Proc. ACM STOC’2008.

    Google Scholar 

  179. G. Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv:math/0211159, 2002.

    Google Scholar 

  180. V. R. Pratt. Every prime has a succinct certificate. SIAM J. Comput., 4(3):214–220, 1975.

    Article  MathSciNet  MATH  Google Scholar 

  181. M. O. Rabin. Probabilistic algorithm for testing primality. J. Number Theory, 12(1):128–138, 1980.

    Article  MathSciNet  MATH  Google Scholar 

  182. R. Raz. Multi-linear formulas for permanent and determinant are of super-polynomial size. J. of the ACM, 56(2):8, 2009. Earlier version in Proc. ACM STOC’2004, pp. 633–641. ECCC TR03-067.

    Google Scholar 

  183. R. Raz. Elusive functions and lower bounds for arithmetic circuits. Theory of Computing, 6:135–177, 2010. Earlier version in Proc. ACM STOC’2008.

    Google Scholar 

  184. R. Raz. Tensor-rank and lower bounds for arithmetic formulas. J. of the ACM, 60(6):40, 2013. Earlier version in Proc. ACM STOC’2010, pp. 659–666.

    Google Scholar 

  185. R. Raz and A. Yehudayoff. Lower bounds and separations for constant depth multilinear circuits. Computational Complexity, 18(2):171–207, 2009. Earlier version in Proc. IEEE Complexity’2008, pp. 128–139.

    Google Scholar 

  186. R. Raz and A. Yehudayoff. Multilinear formulas, maximal-partition discrepancy and mixed-sources extractors. J. Comput. Sys. Sci., 77(1):167–190, 2011. Earlier version in Proc. IEEE FOCS’2008, pp. 273–282.

    Google Scholar 

  187. A. A. Razborov. Lower bounds for the monotone complexity of some Boolean functions. Doklady Akademii Nauk SSSR, 281(4):798–801, 1985. English translation in Soviet Math. Doklady 31:354–357, 1985.

    Google Scholar 

  188. A. A. Razborov. Lower bounds on monotone complexity of the logical permanent. Mathematical Notes, 37(6):485–493, 1985. Original Russian version in Matematischi Zametki.

    Google Scholar 

  189. A. A. Razborov. Lower bounds for the size of circuits of bounded depth with basis \(\left \{\&,\oplus \right \}\). Mathematicheskie Zametki, 41(4):598–607, 1987. English translation in Math. Notes. Acad. Sci. USSR 41(4):333–338, 1987.

    Google Scholar 

  190. A. A. Razborov. Unprovability of lower bounds on circuit size in certain fragments of bounded arithmetic. Izvestiya Math., 59(1):205–227, 1995.

    Article  MathSciNet  MATH  Google Scholar 

  191. A. A. Razborov and S. Rudich. Natural proofs. J. Comput. Sys. Sci., 55(1):24–35, 1997. Earlier version in Proc. ACM STOC’1994, pp. 204–213.

    Google Scholar 

  192. K. W. Regan. Understanding the Mulmuley-Sohoni approach to P vs. NP. Bulletin of the EATCS, 78:86–99, 2002.

    Google Scholar 

  193. B. Rossman, R. A. Servedio, and L.-Y. Tan. An average-case depth hierarchy theorem for Boolean circuits. In Proc. IEEE FOCS, pages 1030–1048, 2015. ECCC TR15-065.

    Google Scholar 

  194. T. Rothvoß. The matching polytope has exponential extension complexity. In Proc. ACM STOC, pages 263–272, 2014.

    Google Scholar 

  195. R. Santhanam. Circuit lower bounds for Merlin-Arthur classes. In Proc. ACM STOC, pages 275–283, 2007.

    Google Scholar 

  196. S. Saraf. Recent progress on lower bounds for arithmetic circuits. In Proc. Conference on Computational Complexity, pages 155–160, 2014.

    Google Scholar 

  197. W. J. Savitch. Relationships between nondeterministic and deterministic tape complexities. J. Comput. Sys. Sci., 4(2):177–192, 1970.

    Article  MathSciNet  MATH  Google Scholar 

  198. U. Schöning. A probabilistic algorithm for k-SAT and constraint satisfaction problems. In Proc. IEEE FOCS, pages 410–414, 1999.

    Google Scholar 

  199. U. Schöning and R. J. Pruim. Gems of Theoretical Computer Science. Springer, 1998.

    Book  MATH  Google Scholar 

  200. A. Shamir. IP=PSPACE. J. of the ACM, 39(4):869–877, 1992. Earlier version in Proc. IEEE FOCS’1990, pp. 11–15.

    Google Scholar 

  201. C. Shannon. The synthesis of two-terminal switching circuits. Bell System Technical Journal, 28(1):59–98, 1949.

    Article  MathSciNet  Google Scholar 

  202. J. Shoenfield. The problem of predicativity. In Y. Bar-Hillel et al., editor, Essays on the Foundations of Mathematics, pages 132–142. Hebrew University Magnes Press, 1961.

    Google Scholar 

  203. P. W. Shor. Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer. SIAM J. Comput., 26(5):1484–1509, 1997. Earlier version in Proc. IEEE FOCS’1994. quant-ph/9508027.

    Google Scholar 

  204. A. Shpilka and A. Wigderson. Depth-3 arithmetic circuits over fields of characteristic zero. Computational Complexity, 10(1):1–27, 2001. Earlier version in Proc. IEEE Complexity’1999.

    Google Scholar 

  205. A. Shpilka and A. Yehudayoff. Arithmetic circuits: a survey of recent results and open questions. Foundations and Trends in Theoretical Computer Science, 5(3-4):207–388, 2010.

    Article  MathSciNet  MATH  Google Scholar 

  206. M. Sipser. A complexity theoretic approach to randomness. In Proc. ACM STOC, pages 330–335, 1983.

    Google Scholar 

  207. M. Sipser. The history and status of the P versus NP question. In Proc. ACM STOC, pages 603–618, 1992.

    Google Scholar 

  208. M. Sipser. Introduction to the Theory of Computation (Second Edition). Course Technology, 2005.

    Google Scholar 

  209. R. Smolensky. Algebraic methods in the theory of lower bounds for Boolean circuit complexity. In Proc. ACM STOC, pages 77–82, 1987.

    Google Scholar 

  210. R. Solovay and V. Strassen. A fast Monte-Carlo test for primality. SIAM J. Comput., 6(1):84–85, 1977.

    Article  MathSciNet  MATH  Google Scholar 

  211. A. Srinivasan. On the approximability of clique and related maximization problems. J. Comput. Sys. Sci., 67(3):633–651, 2003. Earlier version in Proc. ACM STOC’2000, pp. 144–152.

    Google Scholar 

  212. L. J. Stockmeyer. The complexity of approximate counting. In Proc. ACM STOC, pages 118–126, 1983.

    Google Scholar 

  213. A. J. Stothers. On the complexity of matrix multiplication. PhD thesis, 2010.

    Google Scholar 

  214. V. Strassen. Gaussian elimination is not optimal. Numerische Mathematik, 14(13):354–356, 1969.

    Article  MathSciNet  MATH  Google Scholar 

  215. V. Strassen. Vermeidung von divisionen. Journal für die Reine und Angewandte Mathematik, 264:182–202, 1973.

    MathSciNet  MATH  Google Scholar 

  216. B. A. Subbotovskaya. Realizations of linear functions by formulas using +, ×, −. Doklady Akademii Nauk SSSR, 136(3):553–555, 1961. In Russian.

    Google Scholar 

  217. E. R. Swart. P = NP. Technical report, University of Guelph, 1986. Revision in 1987.

    Google Scholar 

  218. R. Szelepcsényi. The method of forced enumeration for nondeterministic automata. Acta Informatica, 26(3):279–284, 1988.

    Article  MathSciNet  MATH  Google Scholar 

  219. A. Tal. Shrinkage of De Morgan formulae by spectral techniques. In Proc. IEEE FOCS, pages 551–560, 2014. ECCC TR14-048.

    Google Scholar 

  220. É. Tardos. The gap between monotone and non-monotone circuit complexity is exponential. Combinatorica, 8(1):141–142, 1988.

    Article  MathSciNet  MATH  Google Scholar 

  221. S. Tavenas. Improved bounds for reduction to depth 4 and depth 3. Inf. Comput., 240:2–11, 2015. Earlier version in Proc. MFCS’2013, pp. 813–824.

    Google Scholar 

  222. S. Toda. PP is as hard as the polynomial-time hierarchy. SIAM J. Comput., 20(5):865–877, 1991. Earlier version in Proc. IEEE FOCS’1989, pp. 514–519.

    Google Scholar 

  223. B. A. Trakhtenbrot. A survey of Russian approaches to perebor (brute-force search) algorithms. Annals of the History of Computing, 6(4):384–400, 1984.

    Article  MathSciNet  MATH  Google Scholar 

  224. L. G. Valiant. Graph-theoretic arguments in low-level complexity. In Mathematical Foundations of Computer Science, pages 162–176, 1977.

    Google Scholar 

  225. L. G. Valiant. Completeness classes in algebra. In Proc. ACM STOC, pages 249–261, 1979.

    Google Scholar 

  226. L. G. Valiant. The complexity of computing the permanent. Theoretical Comput. Sci., 8(2):189–201, 1979.

    Article  MathSciNet  MATH  Google Scholar 

  227. L. G. Valiant. Accidental algorithms. In Proc. IEEE FOCS, pages 509–517, 2006.

    Google Scholar 

  228. L. G. Valiant, S. Skyum, S. Berkowitz, and C. Rackoff. Fast parallel computation of polynomials using few processors. SIAM J. Comput., 12(4):641–644, 1983.

    Article  MathSciNet  MATH  Google Scholar 

  229. Various authors. Deolalikar P vs NP paper (wiki page). Last modified 30 September 2011. michaelnielsen.org/polymath1/index.php?title=Deolalikar_P_vs_NP_paper.

    Google Scholar 

  230. V. Vassilevska Williams. Multiplying matrices faster than Coppersmith-Winograd. In Proc. ACM STOC, pages 887–898, 2012.

    Google Scholar 

  231. N. V. Vinodchandran. A note on the circuit complexity of PP. Theor. Comput. Sci., 347:415–418, 2005. ECCC TR04-056.

    Google Scholar 

  232. A. Wigderson. P, NP and mathematics - a computational complexity perspective. In Proceedings of the International Congress of Mathematicians 2006 (Madrid), pages 665–712. EMS Publishing House, 2007. www.math.ias.edu/~avi/PUBLICATIONS/MYPAPERS/W06/w06.pdf.

  233. A. Wiles. Modular elliptic curves and Fermat’s Last Theorem. Annals of Mathematics, 141(3):443–551, 1995.

    Article  MathSciNet  MATH  Google Scholar 

  234. R. Williams. Better time-space lower bounds for SAT and related problems. In Proc. Conference on Computational Complexity, pages 40–49, 2005.

    Google Scholar 

  235. R. Williams. Time-space tradeoffs for counting NP solutions modulo integers. Computational Complexity, 17(2):179–219, 2008. Earlier version in Proc. IEEE Complexity’2007, pp. 70–82.

    Google Scholar 

  236. R. Williams. Guest column: a casual tour around a circuit complexity bound. ACM SIGACT News, 42(3):54–76, 2011.

    Article  Google Scholar 

  237. R. Williams. Alternation-trading proofs, linear programming, and lower bounds. ACM Trans. on Computation Theory, 5(2):6, 2013. Earlier version in Proc. STACS’2010, pp. 669–680.

    Google Scholar 

  238. R. Williams. Improving exhaustive search implies superpolynomial lower bounds. SIAM J. Comput., 42(3):1218–1244, 2013. Earlier version in Proc. ACM STOC’2010.

    Google Scholar 

  239. R. Williams. Natural proofs versus derandomization. In Proc. ACM STOC, pages 21–30, 2013.

    Google Scholar 

  240. R. Williams. Algorithms for circuits and circuits for algorithms: connecting the tractable and intractable. In Proceedings of the International Congress of Mathematicians, 2014.

    Google Scholar 

  241. R. Williams. New algorithms and lower bounds for circuits with linear threshold gates. In Proc. ACM STOC, pages 194–202, 2014.

    Google Scholar 

  242. R. Williams. Nonuniform ACC circuit lower bounds. J. of the ACM, 61(1):1–32, 2014. Earlier version in Proc. IEEE Complexity’2011.

    Google Scholar 

  243. C. B. Wilson. Relativized circuit complexity. J. Comput. Sys. Sci., 31(2):169–181, 1985.

    Article  MathSciNet  MATH  Google Scholar 

  244. M. Yannakakis. Expressing combinatorial optimization problems by linear programs. J. Comput. Sys. Sci., 43(3):441–466, 1991. Earlier version in Proc. ACM STOC’1988, pp. 223–228.

    Google Scholar 

  245. A. C-C. Yao. Separating the polynomial-time hierarchy by oracles (preliminary version). In Proc. IEEE FOCS, pages 1–10, 1985.

    Google Scholar 

  246. A. C-C. Yao. On ACC and threshold circuits. In Proc. IEEE FOCS, pages 619–627, 1990.

    Google Scholar 

Download references

Acknowledgements

I thank Andy Drucker, Adam Klivans, Ashley Montanaro, Leslie Valiant, and Ryan Williams for helpful observations and for answering questions. I especially thank Michail Rassias for his exponential patience with my delays completing this article. Supported by an NSF Waterman Award, under grant no. 1249349.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Scott Aaronson .

Editor information

Editors and Affiliations

Appendix: Glossary of Complexity Classes

Appendix: Glossary of Complexity Classes

To help you remember all the supporting characters in the ongoing soap opera of which P and NP are the stars, this appendix contains short definitions of the complexity classes that appear in this survey, with references to the sections where the classes are discussed in more detail. For a fuller list, containing over 500 classes, see for example my Complexity Zoo [8]. All the classes below are classes of decision problems—that is, languages \(L \subseteq \left \{0,1\right \}^{{\ast}}\). The known inclusion relations among many of the classes are also depicted in Fig. 3.

Fig. 3
figure 3

Known inclusion relations among 21 of the complexity classes that appear in this survey

Π 2 P::

coNP NP, the second level of the polynomial hierarchy (with universal quantifier in front). See Sect. 2.2.3.

\(\mathsf{\varSigma }_{2}^{\mathsf{P}}\)::

NP NP, the second level of the polynomial hierarchy (with existential quantifier in front). See Sect. 2.2.3.

AC 0::

The class decidable by a nonuniform family of polynomial-size, constant-depth, unbounded-fanin circuits of AND, OR, and NOT gates. See Sect. 6.2.3.

\(\mathsf{AC}^{0}\left [m\right ]\)::

AC 0 enhanced by MOD m gates, for some specific value of m (the case of m a prime power versus a non-prime-power are dramatically different). See Sect. 6.2.4.

ACC::

AC 0 enhanced by MOD m gates, for every m simultaneously. See Sect. 6.4.2.

BPP::

Bounded-Error Probabilistic Polynomial-Time. The class decidable by a polynomial-time randomized algorithm that errs with probability at most 1∕3 on each input. See Sect. 5.4.1.

BQP::

Bounded-Error Quantum Polynomial-Time. The same as BPP except that we now allow quantum algorithms. See Sect. 5.5.

coNP::

The class consisting of the complements of all languages in NP. Complete problems include unsatisfiability, graph non-3-colorability, etc. See Sect. 2.2.3.

\(\mathsf{DTISP}\left (f\left (n\right ),g\left (n\right )\right )\)::

See Sect. 6.4.1.

EXP::

Exponential-Time, or \(\bigcup _{k}\mathsf{TIME}\left (2^{n^{k} }\right )\). Note the permissive definition of “exponential,” which allows any polynomial in the exponent. See Sect. 2.2.7.

EXPSPACE::

Exponential-Space, or \(\bigcup _{k}\mathsf{SPACE}\left (2^{n^{k} }\right )\). See Sect. 2.2.7.

IP::

Interactive Proofs, the class for which a “yes” answer can be proven (to statistical certainty) via an interactive protocol in which a polynomial-time verifier Arthur exchanges a polynomial number of bits with a computationally-unbounded prover Merlin. Turns out to equal PSPACE [200]. See Sect. 6.3.1.

LOGSPACE::

Logarithmic-Space, or \(\mathsf{SPACE}\left (\log n\right )\). Note that only read/write memory is restricted to \(O\left (\log n\right )\) bits; the n-bit input itself is stored in a read-only memory. See Sect. 6.4.1.

MA::

Merlin-Arthur, the class for which a “yes” answer can be proven to statistical certainty via a polynomial-size message from a prover (“Merlin”), which the verifier (“Arthur”) then verifies in probabilistic polynomial time. Same as NP except that the verification can be probabilistic. See Sect. 6.3.2.

MA EXP ::

The exponential-time analogue of MA, where now Merlin’s proof can be \(2^{n^{O\left (1\right )} }\) bits long, and Arthur’s probabilistic verification can also take \(2^{n^{O\left (1\right )} }\) time. See Sect. 6.3.2.

NC 1::

The class decidable by a nonuniform family of polynomial-size Boolean formulas—or equivalently, polynomial-size Boolean circuits of fanin 2 and depth \(O\left (\log n\right )\). The subclass of P∕poly that is “highly parallelizable.” See Sect. 5.2.

NEXP::

Nondeterministic Exponential-Time, or \(\bigcup _{k}\mathsf{NTIME}\left (2^{n^{k} }\right )\). The exponential-time analogue of NP. See Sect. 2.2.7.

NP::

Nondeterministic Polynomial-Time, or \(\bigcup _{k}\mathsf{NTIME}\left (n^{k}\right )\). The class for which a “yes” answer can be proven via a polynomial-size witness, which is verified by a deterministic polynomial-time algorithm. See Sect. 2.

\(\mathsf{NTIME}\left (f\left (n\right )\right )\)::

Nondeterministic \(f\left (n\right )\)-Time. The class for which a “yes” answer can be proven via an \(O\left (f\left (n\right )\right )\)-bit witness, which is verified by a deterministic \(O\left (f\left (n\right )\right )\)-time algorithm. Equivalently, the class solvable by a nondeterministic \(O\left (f\left (n\right )\right )\)-time algorithm. See Sect. 2.2.7.

P::

Polynomial-Time, or \(\bigcup _{k}\mathsf{TIME}\left (n^{k}\right )\). The class solvable by a deterministic polynomial-time algorithm. See Sect. 2.

P #P::

P with an oracle for #P problems (i.e., for counting the exact number of accepting witnesses for any problem in NP). See Sect. 2.2.6.

P∕poly::

P enhanced by polynomial-size “advice strings” \(\left \{a_{n}\right \}_{n}\), which depend only on the input size n but can otherwise be chosen to help the algorithm as much as possible. Equivalently, the class solvable by a nonuniform family of polynomial-size Boolean circuits (i.e., a different circuit is allowed for each input size n). See Sect. 5.2.

PH::

The Polynomial Hierarchy. The class expressible via a polynomial-time predicate with a constant number of alternating universal and existential quantifiers over polynomial-size strings. Equivalently, the union of \(\mathsf{\varSigma }_{1}^{\mathsf{P}} = \mathsf{NP}\), Π 1 P = coNP, \(\mathsf{\varSigma }_{2}^{\mathsf{P}} = \mathsf{NP}^{\mathsf{NP}}\), Π 2 P = coNP NP, and so on. See Sect. 2.2.3.

PP::

Probabilistic Polynomial-Time. The class decidable by a polynomial-time randomized algorithm that, for each input x, guesses the correct answer with probability greater than 1∕2. Like BPP but without the bounded-error (1∕3 versus 2∕3) requirement, and accordingly believed to be much more powerful. See Sect. 2.2.6.

PSPACE::

Polynomial-Space, or \(\bigcup _{k}\mathsf{SPACE}\left (n^{k}\right )\). See Sect. 2.2.5.

\(\mathsf{SPACE}\left (f\left (n\right )\right )\)::

The class decidable by a serial, deterministic algorithm that uses \(O\left (f\left (n\right )\right )\) bits of memory (and possibly up to \(2^{O\left (f\left (n\right )\right )}\) time). See Sect. 2.2.7.

TC 0::

AC 0 enhanced by MAJORITY gates. Also corresponds to “neural networks” (polynomial-size, constant-depth circuits of threshold gates). See Sect. 6.2.5.

\(\mathsf{TIME}\left (f\left (n\right )\right )\)::

The class decidable by a serial, deterministic algorithm that uses \(O\left (f\left (n\right )\right )\) time steps (and therefore, \(O\left (f\left (n\right )\right )\) bits of memory). See Sect. 2.2.7.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Aaronson, S. (2016). \(P\mathop{ =}\limits^{?}NP\) . In: Nash, Jr., J., Rassias, M. (eds) Open Problems in Mathematics. Springer, Cham. https://doi.org/10.1007/978-3-319-32162-2_1

Download citation

Publish with us

Policies and ethics