Skip to main content

Worst-Case to Average-Case Reductions for Subclasses of P

  • Chapter
  • First Online:
Computational Complexity and Property Testing

Part of the book series: Lecture Notes in Computer Science ((LNTCS,volume 12050))

Abstract

For every polynomial q, we present (almost-linear-time) worst-case to average-case reductions for a class of problems in \(\mathcal P\) that are widely conjectured not to be solvable in time q. These classes contain, for example, the problems of counting the number of t-cliques in a graph, for any fixed \(t\ge 3\).

Specifically, we consider the class of problems that consist of counting the number of local neighborhoods in the input that satisfy some predetermined conditions, where the number of neighborhoods is polynomial, and the neighborhoods as well as the conditions can be specified by small uniform Boolean formulas. We show an almost-linear-time reduction from solving any such problem in the worst-case to solving some other problem (in the same class) on typical inputs. Furthermore, for some of these problems, we show that their average-case complexity almost equals their worst-case complexity.

En route we highlight a few issues and ideas such as sample-aided reductions, average-case analysis of randomized algorithms, and average-case emulation of arithmetic circuits by Boolean circuits. We also observe that adequately uniform versions of \(\mathcal{AC}^0[2]\) admit worst-case to average-case reductions that run in almost linear-time.

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 79.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 99.99
Price excludes VAT (USA)
  • Compact, lightweight 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.

    The basic idea underlying the worst-case to average-case reduction of the “permanent” is due to Lipton [24], but his proof implicitly presumes that the field is somehow fixed as a function of the dimension. This issue was addressed independently by [9] and in the proceeding version of [15]. In the current work, we shall be faced with the very same issue.

  2. 2.

    Actually, a worst-case to average-case reduction for a problem of this flavor was shown before by Goldreich and Wigderson [18]. They considered the problem of computing the function \(f_n(A_1,...,A_{\ell (n)}) = \sum _{S\subseteq [\ell (n)]}{} \mathtt{DET}(\sum _{i\in S} A_i)\), where the \(A_i\)’s are n-by-n matrices over a finite field and \(\ell (n)=O(\log n)\), conjectured that it cannot be computed in time \(2^{\ell (n)/3}\), and showed that it is random self-reducible (by O(n) queries). They also showed that it is downwards self-reducible when the field has the form \(GF(2^{m(n)})\) such that \(m(n) = 2^{{\lceil \log _2 n\rceil }}\) (or \(m(n) = 2\cdot 3^{{\lceil \log _3 n\rceil }}\)). We stress that the (subquadratic-time) reduction of [18] does not run in almost-linear time, and that the foregoing problem is not as well-studied as the problems considered in [5].

  3. 3.

    Furthermore, typically, these reductions make a small number of queries (since their queries are of at least linear length). In any case, composing them with a T-time algorithm for the target problem yields an “almost T-time” algorithm for the reduced problem, provided that \(T(n)\in [\varOmega (n),\mathrm{poly}(n)]\).

  4. 4.

    See Theorem 3.1 for a precise statement.

  5. 5.

    Here 0.76 stands for any constant greater than 3/4.

  6. 6.

    Here a “noticeable fraction” is the ratio of a linear function over an almost linear function. We stress that this is not the standard definition of this notion (at least not in cryptography).

  7. 7.

    See Theorem 4.2 for a precise statement. Note that Theorem 4.2 refers to sample-aided reductions, which imply non-uniform reductions. Theorem 1.3 refers to a liberal notion of almost-linearity that includes any function of the form \(f(n)=n^{1+o(1)}\). Under a strict notion that postulates that only functions of the form \(f(n)=n^{1+o(1)}\) are deemed almost-linear, the rare-case refers to an \(1/\mathrm{poly}(\log n)\) fraction of the n-bit long instances.

  8. 8.

    The sizes of these fields are lower-bounded not only by \(\varOmega (\log p(n))\) but also by the size of the Boolean formula used in the original counting problem. The latter lower bound guarantees that the field is larger than the degree of the polynomial that is computed by the Arithmetic problem used in the reduction.

  9. 9.

    Indeed, the original goal of the approximation method is to obtain low degree approximations of \(\mathcal{AC}^0[2]\) circuits.

  10. 10.

    This is the case since typically the randomized reduction of \(\varPi \) to \(\varPi '\) makes several queries to \(\varPi '\).

  11. 11.

    For each \(x\in \{0,1\}^n\), define a 0–1 random variable \(\zeta _x=\zeta _x(h)\) such that \(\zeta _x(h)=1\) if and only if \(C_h(x)\) is correct. Then, \(\mu _x{\mathop {=}\limits ^\mathrm{def}}\mathrm{E}[\zeta _x]\) equals the probability that A(x) is correct, and \({\mu }{\mathop {=}\limits ^\mathrm{def}}\sum _{x\in \{0,1\}^n}\mu _x\ge \rho (n)\cdot 2^n\). Using Chebyshev’s inequality, we get

    $$\begin{aligned} \mathrm{Pr}\left[ \left| \sum _{x\in \{0,1\}^n}\zeta _x-{\mu }\right| \ge 2^n\cdot \epsilon \cdot \rho (n)\right]\le & {} {{\mathrm{Var}\left[ \sum _{x\in \{0,1\}^n}\zeta _x\right] } \over {(2^n\cdot \epsilon \cdot \rho (n))^2}}\\= & {} {{\sum _{x\in \{0,1\}^n}\mathrm{Var}[\zeta _x]}\over {2^{2n}\epsilon ^2\rho (n)^2}}\\< & {} {{1}\over {2^n\epsilon ^2\rho (n)^2}} \end{aligned}$$

    where the equality is due to the pairwise-independence of the \(\zeta _x\)’s. Using \(\rho (n)>2^{-n/3}\), we obtain the desired bound (of \(2^{-n/3}/\epsilon ^2\)). Indeed, the same argument supports \(\rho (n)\ge 2^{-\beta n}\), for any constant \(\beta \in (0,0.5)\), and it yields a meaningful result for \(\epsilon =2^{-(1-2\beta )/3}\).

  12. 12.

    Recall that, given a formula F of size s, this transformation first locates a sub-formula \(F'\) of size \(s'\in [s/3,2s/3]\). Letting \(F_b\) be the formula that results from F when replacing \(F'\) by the constant \(b\in \{0,1\}\), the transformation is recursively applied to the formulae \(F_0,F_1\) and \(F'\), while noting that each of these formulae has size at most 2s/3. Lastly, we output the formula \((F'\wedge F_1)\vee ((\lnot F')\wedge F_0)\).

  13. 13.

    This polynomial is unique since we cannot have two different polynomials of degree \(D-1\) that each agree with 2D of the 3D points.

  14. 14.

    Recall that integer arithmetics is in \(\mathcal{NC}\); see, e.g., [23, Lect. 30].

  15. 15.

    The bound of 8/9 used in the proof of Proposition 3.4 is a consequence of lower-bounding by 2/3 the probability that two third of the points on a random line that passes through a given point are answered correctly. However, a lower-bound of \(0.5+o(1)\) would have been as good (when combined with error reduction).

  16. 16.

    Indeed, this move from the success rate (of P) to a fraction of correct answers (on a random curve) is analogous to the proof of Proposition 1.5; in the current context we get it “for free”.

  17. 17.

    This uses the hypothesis that \(\ell ''(n)=c\log n\), which implies that \(\ell ''(\exp (d'(n))\cdot n)=O(d'(n))+\ell ''(n)\).

  18. 18.

    Indeed, this follows a paradigm that can be traced to the work of Impagliazzo and Wigderson [20].

  19. 19.

    The sample can be used to test the candidate oracle machines (as outlined in the foregoing discussion). Note that this allows to distinguish machines that are correct on all inputs from machines that err on a noticeable fraction of the inputs, but not to rule out machines that err on a negligible fraction of the inputs. Hence, a worst-case to rare-case reduction of the list-decoding type only yields a sample-aided reduction from solving the original problem on a \(1-o(1)\) fraction of the instances.

  20. 20.

    Note that the formula \(\varPhi \) that underlies the counting problem \(\varPi ''\) can be implemented as a selector function that picks the corresponding formula \(\varPhi ^{(p)}\) that emulates the computation of \({\widehat{\varPhi }}\) in \(\mathrm{GF}(p)^n\).

  21. 21.

    In contrast, it is not clear how to approximate the agreement of \({\widehat{\varPhi }}\) and M over \(\mathrm{GF}(p)^n\), since we only have a solved sample of instances that are uniformly distributed in \(\{0,1\}^n\).

  22. 22.

    We shall reduce solving the original (worst-case) instance of length n to solving (in the rare-case sense) instances of various lengths, which correspond to different prime fields. Hence, for each input length for the original problem, we rely on being able to rarely solve the reduced problem on several input lengths (rather than on one input length as in the reductions presented so far). We note that this disadvantage is inherent to the downwards self-reduction paradigm of [20], so we may just take advantage of it for this additional purpose.

  23. 23.

    Specifically, we use a multi-linear extension \(\mathtt{SEL}:\mathrm{GF}(p)^{\log _2n}\times \mathrm{GF}(p)^n\rightarrow \mathrm{GF}(p)\) of the selection function \(\mathtt{sel}:\{0,1\}^{\log _2n}\times \mathrm{GF}(p)^n\rightarrow \mathrm{GF}(p)\), which satisfies \(\mathtt{sel}(\alpha ,x)=x_{\mathtt{int}(\alpha )+1}\), where \(\mathtt{int}(\alpha )\) is the integer represented by the binary string \(\alpha \); that is,

    $$\mathtt{SEL}(\zeta ,x) = \sum _{\beta \in \{0,1\}^{\log _2n}} \prod _{~~k\in [\log _2n]}(\beta _k\zeta _k+(1-\beta _k)(1-\zeta _k)) \cdot x_{\mathtt{int}(\beta )+1}$$

    (where a crucial point is that \(\mathtt{SEL}(\zeta ,x)\) is the sum of n terms such that the \(i^\mathrm{th}\) term is a multilinear function of the \(\zeta _k\)’s and \(x_i\)). Then,

    $${\widehat{\psi }}_{n}(vw',x) = {\widehat{\phi }}_{n}(vw',F_{1}(vw',x),...,F_{m(n)}(vw',x))$$

    where \(F_k(w,x)=\mathtt{SEL}({\widehat{\sigma }}_{n,k}(w),x)\) and \({\widehat{\sigma }}_{n,k}:\mathrm{GF}(p_i)^{\ell (n)}\rightarrow \mathrm{GF}(p_i)^{\log n}\) is a low degree polynomial that agrees with \({\sigma }_{n,k}:\{0,1\}^{\ell (n)}\rightarrow [n]\) (analogously to the way \({\widehat{\phi }}_n\) is derived from \(\phi _n\) (cf. the proof of Proposition 3.3)). Note that, for \(w\in \{0,1\}^{\ell (n)}\), it holds that \(\mathtt{SEL}({\widehat{\sigma }}_{n,k}(w),x)=\mathtt{SEL}({\sigma }_{n,k}(w),x)=x_{{\sigma }_{n,k}(w)}\).

  24. 24.

    We stress that we use the foregoing length convention, by which the length refers to the number of field elements in a sequence (rather than to the length of the binary representation of the sequence).

  25. 25.

    As noted in Sect. 4.1, although the statement of [31, Thm. 29] only claims running time that is polynomial in all relevant parameters, it is clear that the running time is linear in the number of variables.

  26. 26.

    We could not do this in Sect. 4.1, since the solved instances we obtained there were all in \(\{0,1\}^{{\widetilde{n}}_i}\), whereas different low degree polynomials over \(\mathrm{GF}(p_i)^{{\widetilde{n}}_i}\) may agree on \(\{0,1\}^{{\widetilde{n}}_i}\).

  27. 27.

    Actually, this is a slightly revised version, which is essentially equivalent to the original: In [16, Def. 2], instead of w, the formula \(\phi _n\) got as part of its input the sequence \(({\sigma }_{n,1}(w),...,{\sigma }_{n,p(\log n)}(w))\). This is essentially equivalent to the form used here, since, on the one hand, \(\phi _n\) can compute the \({\sigma }_{n,i}\)’s (given w), and on the other hand w can be reconstructed from \(\frac{\ell (n)}{\log _2n}=c\) auxiliary \({\sigma }_{n,i}\)’s.

  28. 28.

    Indeed, it is required that in case of inputs in S, the predicate \(\phi _n\) evaluates to 0 (rather than to 1). This choice was made in [16] in order to simplify the expansion. We stress that since n is presented in binary, the algorithm runs in \(\mathrm{poly}(\log n)\)-time.

  29. 29.

    Recall that this construction replaces each or-gate by a conjunction of \(O(\log n^c)\) random linear combinations of the values that feed the original or-gate.

  30. 30.

    Using the third construction in [2], we need to perform exponentiation in a field of size \(2^{k/2}\), where \(k=O(\log n^c)\) is the length of the seed. By [19, Thm. 4], this operation can be performed by highly uniform constant-depth circuit (with parity gates) of size \(\exp ({\widetilde{O}}({\sqrt{k/2}}))=n^{o(1)}\).

  31. 31.

    Recall that taking t independent linear combinations of the output of an \(\epsilon \)-bias generator yields a distribution that equal \(0^t\) with probability at most \(2^{-t}+\epsilon \). Also recall that the aforementioned generator of [2] produces a \(n^c\)-bit long \(\epsilon \)-biased sequence using a seed of length \(O(\log (n^c/\epsilon ))\), and so we can set \(t=O(c\log n)\) and \(\epsilon =2^{-t}\).

References

  1. Ajtai, M.: \({\Sigma }_{1}^{1}\)-formulae on finite structures. Ann. Pure Appl. Logic 24(1), 1–48 (1983)

    Article  MathSciNet  Google Scholar 

  2. Alon, N., Goldreich, O., Hastad, J., Peralta, R.: Simple construction of almost k-wise independent random variables. Random Struct. Algorithms 3(3), 289–304 (1992)

    Article  MathSciNet  Google Scholar 

  3. Babai, L.: Random oracles separate PSPACE from the polynomial-time hierarchy. IPL 26, 51–53 (1987)

    Article  MathSciNet  Google Scholar 

  4. Babai, L., Fortnow, L., Nisan, N., Wigderson, A.: BPP has subexponential time simulations unless EXPTIME has publishable proofs. Complex. Theory 3, 307–318 (1993)

    MathSciNet  MATH  Google Scholar 

  5. Ball, M., Rosen, A., Sabin, M., Vasudevan, P.N.: Average-case fine-grained hardness. In: The Proceedings of STOC, pp. 483–496 (2017)

    Google Scholar 

  6. Barak, B.: A probabilistic-time hierarchy theorem for “slightly non-uniform” algorithms. In: Rolim, J.D.P., Vadhan, S. (eds.) RANDOM 2002. LNCS, vol. 2483, pp. 194–208. Springer, Heidelberg (2002). https://doi.org/10.1007/3-540-45726-7_16

    Chapter  Google Scholar 

  7. Barkol, O., Ishai, Y.: Secure computation of constant-depth circuits with applications to database search problems. In: Shoup, V. (ed.) CRYPTO 2005. LNCS, vol. 3621, pp. 395–411. Springer, Heidelberg (2005). https://doi.org/10.1007/11535218_24

    Chapter  Google Scholar 

  8. Bogdanov, A., Trevisan, L.: On worst-case to average-case reductions for NP problems. SIAM J. Comput. 36(4), 1119–1159 (2006)

    Article  MathSciNet  Google Scholar 

  9. Cai, J.-Y., Pavan, A., Sivakumar, D.: On the hardness of permanent. In: Meinel, C., Tison, S. (eds.) STACS 1999. LNCS, vol. 1563, pp. 90–99. Springer, Heidelberg (1999). https://doi.org/10.1007/3-540-49116-3_8

    Chapter  Google Scholar 

  10. Feigenbaum, J., Fortnow, L.: Random-self-reducibility of complete sets. SIAM J. Comput. 22(5), 994–1005 (1993)

    Article  MathSciNet  Google Scholar 

  11. Gemmell, P., Lipton, R.J., Rubinfeld, R., Sudan, M., Wigderson, A.: Self-testing/correcting for polynomials and for approximate functions. In: The Proceedings of ACM Symposium on the Theory of Computing, pp. 32–42 (1991)

    Google Scholar 

  12. Goldmann, M., Grape, P., Hastad, J.: On average time hierarchies. Inf. Process. Lett. 49(1), 15–20 (1994)

    Article  MathSciNet  Google Scholar 

  13. Goldreich, O.: Computational Complexity: A Conceptual Perspective. Cambridge University Press, Cambridge (2008)

    Book  Google Scholar 

  14. Goldreich, O., Nisan, N., Wigderson, A.: On Yao’s XOR-Lemma. In: ECCC, TR95-050 (1995)

    Google Scholar 

  15. Goldreich, O., Ron, D., Sudan, M.: Chinese remaindering with errors. IEEE Trans. Inf. Theory 46(4), 1330–1338 (2000). Preliminary version in 31st STOC (1999)

    Article  MathSciNet  Google Scholar 

  16. Goldreich, O., Rothblum, G.N.: Simple doubly-efficient interactive proof systems for locally-characterizable sets. In: The Proceedings of ITCS, pp. 18:1–18:19 (2018)

    Google Scholar 

  17. Goldreich, O., Rothblum, G.N.: Counting t-cliques: worst-case to average-case reductions and direct interactive proof systems. In: The Proceedings of FOCS, pp. 77–88 (2018)

    Google Scholar 

  18. Goldreich, O., Wigderson, A.: Derandomization that is rarely wrong from short advice that is typically good. In: Rolim, J.D.P., Vadhan, S. (eds.) RANDOM 2002. LNCS, vol. 2483, pp. 209–223. Springer, Heidelberg (2002). https://doi.org/10.1007/3-540-45726-7_17

    Chapter  Google Scholar 

  19. Healy, A., Viola, E.: Constant-depth circuits for arithmetic in finite fields of characteristic two. In: Durand, B., Thomas, W. (eds.) STACS 2006. LNCS, vol. 3884, pp. 672–683. Springer, Heidelberg (2006). https://doi.org/10.1007/11672142_55

    Chapter  Google Scholar 

  20. Impagliazzo, R., Wigderson, A.: Randomness vs time: derandomization under a uniform assumption. J. Comput. Syst. Sci. 63(4), 672–688 (2001)

    Article  MathSciNet  Google Scholar 

  21. Impagliazzo, R., Jaiswal, R., Kabanets, V., Wigderson, A.: Uniform direct product theorems: simplified, optimized, and derandomized. SIAM J. Comput. 39(4), 1637–1665 (2010)

    Article  MathSciNet  Google Scholar 

  22. Jukna, S.: Boolean Function Complexity: Advances and Frontiers. Algorithms and Combinatorics, vol. 27. Springer, Heidelberg (2012)

    Book  Google Scholar 

  23. Kozen, D.: The Design and Analysis of Algorithms. Springer, New York (1991). https://doi.org/10.1007/978-1-4612-4400-4

    Book  MATH  Google Scholar 

  24. Lipton, R.J.: New directions in testing. In: Feigenbaum, J., Merritt, M. (eds.) Distributed Computing and Cryptography. DIMACS Series in Discrete Mathematics and Theoretical Computer Science, vol. 2, pp. 191–202. American Mathematics Society, Providence (1991)

    Chapter  Google Scholar 

  25. Naor, J., Naor, M.: Small-bias probability spaces: efficient constructions and applications. SIAM J. Comput. 22(4), 838–856 (1993). Preliminary version in 22nd STOC (1990)

    Article  MathSciNet  Google Scholar 

  26. Razborov, A.A.: Lower bounds on the size of bounded-depth networks over a complete basis with logical addition. Matematicheskie Zametki 41(4), 598–607 (1987). (in Russian). English translation in Math. Notes Acad. Sci. USSR 41(4), 333–338 (1987)

    MathSciNet  MATH  Google Scholar 

  27. Rubinfeld, R., Sudan, M.: Self-testing polynomial functions efficiently and over rational domains. In: The Proceedings of 3rd SODA, pp. 23–32 (1992)

    Google Scholar 

  28. Rubinfeld, R., Sudan, M.: Robust characterization of polynomials with applications to program testing. SIAM J. Comput. 25(2), 252–271 (1996). Unifies and extends part of the results contained in [11] and [27]

    Article  MathSciNet  Google Scholar 

  29. Smolensky, R.: Algebraic methods in the theory of lower bounds for Boolean circuit complexity. In: 19th ACM Symposium on the Theory of Computing, pp. 77–82 (1987)

    Google Scholar 

  30. Spira, P.M.: On time-hardware complexity tradeoffs for Boolean functions. In: Proceedings of 4th Hawaii Symposium on System Sciences, pp. 525–527 (1971)

    Google Scholar 

  31. Sudan, M., Trevisan, L., Vadhan, S.P.: Pseudorandom generators without the XOR Lemma. J. Comput. Syst. Sci. 62(2), 236–266 (2001)

    Article  MathSciNet  Google Scholar 

  32. Vassilevska Williams, V.: Hardness of easy problems: basing hardness on popular conjectures such as the strong exponential time hypothesis. In: 10th International Symposium on Parameterized and Exact Computation, pp. 17–29 (2015)

    Google Scholar 

Download references

Acknowledgements

We are grateful to Madhu Sudan for many useful discussions regarding list decoding of multivariate polynomials and related issues.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Oded Goldreich or Guy N. Rothblum .

Editor information

Editors and Affiliations

Appendices

Appendices

Appendix A.1 relates Definition 2.3 to [16, Def. 2], and Appendix A.2 presents a worst-case to average-case reduction for uniform \(\mathcal{AC}^0[2]\).

1.1 A.1 A Related Class (for Context Only)

In this appendix, we review the definition of locally-characterizable sets [16], and discuss its relation to Definition 2.3.

Definition A.1

(locally-characterizable sets [16, Def. 2]):Footnote 27 A set S is locally-characterizable if there exist a constant c, a polynomial p and a polynomial-time algorithm that on input n outputs \(\mathrm{poly}(\log n)\)-sized formulae \(\phi _n:\{0,1\}^{c\cdot \log n}\times \{0,1\}^{p(\log n)}\rightarrow \{0,1\}\) and \({\sigma }_{n,1},...,{\sigma }_{n,p(\log n)}:\{0,1\}^{c\cdot \log n}\rightarrow [n]\) such that, for every \(x\in \{0,1\}^n\), it holds that \(x\in S\) if and only if for all \(w\in \{0,1\}^{c\log n}\)

$$\begin{aligned} \varPhi _x(w){\mathop {=}\limits ^\mathrm{def}}\phi _n(w,x_{{\sigma }_{n,1}(w)},...,x_{{\sigma }_{n,p(\log n)}(w)}) \end{aligned}$$
(10)

equals 0.Footnote 28

That is, each value of \(w\in \{0,1\}^{c\log n}\) corresponds to a local condition that refers to polylogarithmically many locations in the input (i.e., the locations \({\sigma }_{n,1}(w),...,{\sigma }_{n,p(\log n)}(w)\in [n]\)). This local condition is captured by \(\phi _n\), and in its general form it depends both on the selected locations (equiv., on w) and on the value on the input at these locations. A simplified form, which suffices in many case, uses a local condition that only depends on the values of the input at these locations (i.e., \(\phi _n:[n]^{ca\cdot \log n}\times \{0,1\}^{p(\log n)}\rightarrow \{0,1\}\) only depends on the \(p(\log n)\)-bit long suffix).

A locally-characterizable set corresponds to the set of inputs for the counting problem (of Definition 2.3) that have value 0 (i.e., the number of satisfied local conditions is 0). This correspondence is not an equality, because the sizes of the formulae in the two definitions are different. Whereas in Definition 2.3 the number of formulae and their sizes are exponential in a function d that is in the admissible class, in Definition A.1 the number and size is poly-logarithmic in n. In both definitions, the formulae are constructed in time that is polynomial in their number and size.

1.2 A.2 Worst-Case to Average-Case Reduction for Uniform \(\mathcal{AC}^0[2]\)

For constants \(c,d\in \mathbb N\), let \(\mathtt{C}^{(d,c)}\) denote the class of decision problems on n-bit inputs that can be solved by uniform families of (unbounded fan-in) Boolean circuits of depth d and size \(n^c\), with and, or, parity, and not gates. Specifically, a problem is parameterized by an efficient algorithm that on input n and \(i,j\in [n^{c}]\) returns the type of the \(i^\mathrm{th}\) gate and whether or not the \(j^\mathrm{th}\) gate feeds into it (in the circuit that corresponds to n-bit inputs). The term “efficient” is left unspecified on purpose; possible choices include \(\mathrm{poly}(n)\)-time and \(O(\log n)\)-space. Note that any decision problem having sufficiently uniform \(\mathcal{AC}^0[2]\) circuits is in \(\mathtt{C}^{(d,c)}\) for some constants \(c,d\in \mathbb N\).

Theorem A.2

(worst-case to average-case reduction for \(\mathcal{AC}_0\)): There exists a universal constant \(\gamma \) such that for any \(c,d\in \mathbb N\), solving any problem in \(\mathtt{C}^{(d,c)}\) on the worst-case reduces in almost linear time to solving some problem in \(\mathtt{C}^{(\gamma \cdot d,c+o(1))}\) on at least 90% of the instances.

Proof Sketch:

We proceed in three steps: First we reduce the Boolean problem (in \(\mathtt{C}^{(d,c)}\)) to an Arithmetic problem, next we show that the latter problem supports a worst-case to average-case reduction, and lastly we reduce the Arithmetic problem to a Boolean problem (in \(\mathtt{C}^{(\gamma \cdot d,c+o(1))}\)). This is very similar to what was done in the main part of this work, except that the first step is fundamentally different.

A straightforward emulation of the Boolean circuit by the Arithmetic circuit would yield multiplication gates of polynomial fan-in, which in turn would mean that the polynomial computed by this circuit is of polynomial degree. In contrast, using the approximation method of Razborov [26] and Smolensky [29], we can get multiplication gates of logarithmic fan-in, and arithmetic circuits that compute polynomials of polylogarithmic degree. The approximation error is of no real concern, since we are actually interested in the value of the circuit at a single point. We need, however, to perform the foregoing randomized reduction using an almost linear amount of randomness, since we need to run in almost linear time, but this is possible using small-bias generators (cf. [25]). Details follow.

The first step is a randomized reduction of solving the Boolean problem in the worst-case to solving a corresponding Arithmetic problem on the worst-case. This reduction uses the ideas underlying the approximation method of Razborov [26] and Smolensky [29], while working with the field \(\mathrm{GF}(2)\) (as [26], rather than with \(\mathrm{GF}(p)\) for some prime \(p>2\) (as [29])).Footnote 29 When doing so, we replace the random choices made at each gate by pseudorandom choices that are generated by a small bias generator [25]; specifically, we use a “highly uniform” generator \(G:\{0,1\}^{O(\log n^c)}\rightarrow \{0,1\}^{{\widetilde{O}}(n^c)}\) such that the individual bits of G(s) are computed by uniform \(n^{o(1)}\)-size circuits of constant depth (and parity gates) [2, 19].Footnote 30 We stress that the same pseudorandom sequence can be used for all gates in the circuit, and in each gate we can use \(O(\log n^c)\) disjoint portions of the pseudorandom sequence for the \(O(\log n^c)\) different linear combinations.Footnote 31

Hence, for a fixed Boolean circuit \(C_n\), on input \(x\in \{0,1\}^n\), we uniformly select a seed \(s\in \{0,1\}^{O(\log n)}\) for the aforementioned small-bias generator G, and construct the corresponding Arithmetic circuit \(A_n^{(s)}:\mathrm{GF}(2)^n\rightarrow \mathrm{GF}(2)\), in which or-gates of \(C_n\) are replaced by \(O(c\log n)\)-way multiplications of linear combinations of the original gates that are determined by G(s). For any fixed x, we may have \(\mathrm{Pr}_s[A_n^{(s)}(x)\!\ne \!C_n(x)]=1/\mathrm{poly}(n)\). Note that the depth of \(A_n^{(s)}\) is only O(1) times larger than the depth of \(C_n\), where the constant is determined by various (local) manipulations (which include replacing and-gates by or-gates, computing inner products of gates’ values and generator outputs, and adding \(O(\log n)\)-wise multiplication gates). Furthermore, \(A_n^{(s)}\) uses multiplication gates of \(O(c\log n)\) arity, its size is at most \(O(\log n)^d\cdot n^{c}\), and it computes a polynomial of degree \(O(c\log n)^d\).

The next step is to embed \(\mathrm{GF}(2)\) in an extension field of size greater than the foregoing degree so that the standard process of self-correction of polynomials can be performed. Hence, \(A_n^{(s)}\) is now viewed as an arithmetic circuit over \(\mathcal{F}=\mathrm{GF}(2^\ell )\) (i.e., \(A_n^{(s)}:\mathcal{F}^{n}\rightarrow \mathcal{F}\)), where \(\ell =d\log \log n + O(d)\), since we need \(2^\ell \ge O(c\cdot \log n)^d\). Now, a worst-case to average-case reduction is applied to \(A_n^{(s)}\) (i.e., evaluating \(A_n^{(s)}\) on the worst case reduces to evaluating \(A_n^{(s)}\) correctly on at least a \(51\%\) fraction of the instances).

Lastly, we wish to get back to a class of Boolean problems. We can do so as follows. First, we replace each (unbounded) \(\mathrm{GF}(2^\ell )\)-addition gate by \(\ell \) parity gates (which add-up the \(\ell \) corresponding bits in the sequence of field elements). Next, we replace each \(\mathcal{F}\)-multiplication gate of arity at most \(m=O(\log n)\) by a \(\mathcal{F}\)-multiplication gate of arity \(\sqrt{m}\) that is fed by \(\sqrt{m}\) multiplication gates that cover the original m wires. Finally, we implement each of the latter gates by a small Boolean circuit of depth two (via a look-up table of size \(|\mathcal{F}|^{\sqrt{m}}=\exp (\ell \cdot {\sqrt{m}})=\exp ({\widetilde{O}}({\sqrt{\log n}}))=n^{o(1)}\)). Hence, given the Arithmetic circuit \(A_n^{(s)}:\mathrm{GF}(2^\ell )^{n}\rightarrow \mathrm{GF}(2^\ell )\), we obtain a Boolean circuit \(B_n^{(s)}:\{0,1\}^{n\ell }\rightarrow \{0,1\}^\ell \) that emulates \(A_n^{(s)}\). Furthermore, using the small circuits that produce the output bits of the small-bias generator G on seed \(s\in \{0,1\}^k\), where \(k=O(\log n)\), we obtain a Boolean circuit \(B_n:\{0,1\}^{n\ell +k}\rightarrow \{0,1\}^\ell \) such that \(B_n(y,s)=B_n^{(s)}(y)\). Recalling that the aforementioned circuits that compute the bits of G have constant depth and size \(n^{o(1)}\), it follows that \(B_n\) has depth \(O(d)+O(1)\) and size \(n^{c+o(1)}\).

We are almost done, except that we need to reduce the evaluation of \(B_n:\{0,1\}^{n\ell +k}\rightarrow \{0,1\}^\ell \) to the evaluation of a Boolean circuit that has a single output bit, and we need this reduction to work in the average-case setting. We can do so by using the Boolean circuit \(B'_n:\{0,1\}^{n\ell +k+\ell }\rightarrow \{0,1\}\) that, on input \((z,r)\in \{0,1\}^{n\ell +k}\times \{0,1\}^\ell \), returns the inner product mod 2 of \(B_n(z)\) and r. Note that the depth of \(B'_n\) exceeds the depth of \(B_n\) only by a constant term, and that we can correctly retrieve \(B_n(z)\) (with high probability) if we can obtain the correct value of \(B'_n(z,r)\) correctly on \(0.76\%\) of the r’s. Recalling that evaluating \(C_n\) on x was reduced to evaluating \(B_n\) correctly on \(0.51\%\) of the instances, it follows that it suffices to compute \(B'_n\) correctly on a \(\rho \) fraction of the instances, provided that \(\rho \ge 1-0.24\cdot 0.49\). (Of course, 0.24 and 0.49 stand for any constants smaller than 0.25 and 0.5 respectively.)

To summarize, evaluating \(C_n:\{0,1\}^n\rightarrow \{0,1\}\) on the worst case reduces to evaluating \(B'_n:\{0,1\}^{\log \log n+O(\log n)}\rightarrow \{0,1\}\) on at least \(1-0.24\cdot 0.49\approx 0.88\) fraction of the instances. Using an adequate indexing of the gates in \(B'_n\), the uniformity of \(B'_n\) follows from the uniformity of \(C_n\), since all modifications we have performed are local.    \(\blacksquare \)

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Goldreich, O., Rothblum, G.N. (2020). Worst-Case to Average-Case Reductions for Subclasses of P. In: Goldreich, O. (eds) Computational Complexity and Property Testing. Lecture Notes in Computer Science(), vol 12050. Springer, Cham. https://doi.org/10.1007/978-3-030-43662-9_15

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-43662-9_15

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-43661-2

  • Online ISBN: 978-3-030-43662-9

  • eBook Packages: Computer ScienceComputer Science (R0)

Publish with us

Policies and ethics