DOKK Library

Some Limitations of the Sum of Small-Bias Distributions

Authors Chin Ho Lee, Emanuele Viola,

License CC-BY-3.0

Plaintext
                          T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23
                                       www.theoryofcomputing.org




                    Some Limitations of the
                 Sum of Small-Bias Distributions
                                 Chin Ho Lee∗                       Emanuele Viola†
              Received August 3, 2015; Revised September 22, 2016; Published December 14, 2017




       Abstract: We present two approaches to constructing ε-biased distributions D on n bits and
       functions f : {0, 1}n → {0, 1} such that the XOR of two independent copies (D + D) does
       not fool f . Using them, we give constructions for any of the following choices:
          1. ε = 2−Ω(n) and f is in P/poly;
          2. ε = 2−Ω(n/ log n) and f is in NC2 ;
          3. ε = n−c and f is a one-way space O(c log n) algorithm, for any c;
          4. ε = n−Ω(1) and f is a mod 3 linear function.
       All the results give one-sided distinguishers, and extend to the XOR of more copies for
       suitable ε. We also give conditional results for AC0 and DNF formulas.
           Meka and Zuckerman (RANDOM 2009) prove 4 with ε = O(1). Bogdanov,    √        Dvir,
       Verbin, and Yehudayoff (Theory of Computing 2013) prove 2 with ε = 2−O( n) . Chen and
       Zuckerman (personal communication) give an alternative proof of 3.

ACM Classification: F.1.3, G.3, F.2.3
AMS Classification: 68Q17
Key words and phrases: complexity theory, pseudorandomness, RL vs. L, error-correcting codes,
k-wise independence, small-bias distributions, sum of small bias

    A preliminary version of this paper appeared in ECCC [20].
   ∗ Supported by NSF grant CCF-1319206. Work done in part at Harvard University.
   † Supported by NSF grant CCF-1319206. Work done in part while a visiting scholar at Harvard University, with support

from Salil Vadhan’s Simons Investigator grant.


© 2017 Chin Ho Lee and Emanuele Viola
c b Licensed under a Creative Commons Attribution License (CC-BY)                         DOI: 10.4086/toc.2017.v013a016
                                   C HIN H O L EE AND E MANUELE V IOLA

1    Introduction and our results
We consider probability distributions over {0, 1}n , i. e., functions D : {0, 1}n → [0, 1] such that

                                                ∑ D(x) = 1 .
                                              x∈{0,1}n


We shall refer to such probability distributions simply as “distributions.”
    Small-bias distributions, introduced by Naor and Naor [24], cf. [1, 2, 5], are distributions that look
balanced to parity functions over {0, 1}n .

Definition 1.1. A distribution D over {0, 1}n is ε-biased if for every nonempty subset I ⊆ [n], we have

                                           Ex∼D (−1)∑i∈I xi ≤ ε .
                                                          


    An ε-biased distribution can be generated using a seed of O(log(n/ε)) bits. Since their introduction,
small-bias distributions have become a fundamental object in theoretical computer science and have
found their uses in many areas including derandomization and algorithm design.
    In the last decade or so researchers have considered the sum (i. e., bitwise XOR) of several independent
copies of small-bias distributions. The first paper to explicitly consider it is [8]. This distribution appears
to be significantly more powerful than a single small-bias copy, while retaining a modest seed length. In
particular, two main questions have been asked:

Question 1.2 (RL). Reingold and Vadhan (personal communication) asked whether there exists a constant
c such that the sum of two independent copies of any n−c -biased distribution fools one-way logarithmic
space, a. k. a. one-way polynomial-size branching programs, which would imply RL = L. It is known that
a small-bias distribution fools one-way width-2 branching programs (Saks and Zuckerman, see also [7]
where a generalization is obtained). No such result is known for width-3 programs.

Question 1.3 (Polynomials). The papers [8, 21, 39] show that the sum of d small-bias generators fools
F2 -polynomials of degree d. However, the proofs only apply when d ≤ (1 − Ω(1)) log n. It is an open
question whether the construction works for larger d. If the construction worked for any d = logO(1) n, it
would make progress on long-standing open problems in circuit complexity regarding AC0 with parity
gates [28]. This question is implicit in the works [8, 21, 39] and explicit in [38, Chapter 1] (Open question
4), cf. the survey [38, Chapter 1].

    In terms of negative results, Meka and Zuckerman [22] show that the sum of 2 distributions with
constant bias does√ not fool mod 3 linear functions. Bogdanov, Dvir, Verbin, and Yehudayoff [7] show
that for ε = 2−O( n/k) , the sum of k copies of ε-biased distributions does not fool circuits of size poly(n)
and depth O(log2 n) (NC2 ).
    This paper gives two different approaches to improving on both these results and obtain other
limitations of the sum of small-bias distributions. One is based on the complexity of decoding, and the
other one on bounding the mod 3 rank (see Definition 1.8). Either approach is a candidate to answer
negatively the “RL question” (Question 1.2).

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                2
                    S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

1.1   Our results
The following theorem states our main counterexamples. We denote by D + D the bitwise XOR of two
independent copies of a distribution D.
Theorem 1.4. For any c, there exists an explicit ε-biased distribution D over {0, 1}n and an explicit
function f , such that f (D + D) = 0 and Prx∼{0,1}n [ f (x) = 0] ≤ p, where ε, f , p are of any one of the
following choices:
    i. ε = 2−Ω(n) , f is a uniform poly(n)-size circuit, and p = 2−Ω(n) ;
   ii. ε = 2−Ω(n/ log n) , f is a uniform fan-in 2, poly(n)-size circuit of depth O(log2 n), and p = 2−n/4 ;
  iii. ε = 1/nc , f is a one-way O(c log n)-space algorithm, and p = O(1/nc );
   iv. ε = n−Ω(1) , f is a mod 3 linear function, and p = 1/2.
    Moreover, all our results extend to more copies of D as follows. The input D + D to f can be replaced
by the bitwise XOR of k independent copies of D if ε is replaced by ε 2/k , where k is at most the following
quantities corresponding to the above items: i. n/60; ii. n/6 log n; iii. 2c; iv. O(log n/ log log n).
    Theorem 1.4.i is tight up to the constant in the exponent because every ε2−n -biased distribution is
ε-close to uniform.
    Theorem 1.4.ii would also be true with ε = 2−Ω(n) , if a decoder for certain algebraic-geometric
                                                                                    √                codes
ran in NC2 , which we conjecture it does. [7] prove Theorem 1.4.ii with ε = 2−O( n/k) .
    Theorem 1.4.iii can also be obtained in the following way, pointed out to us by Chen and Zuckerman
(personal communication). Since one can distinguish a set of size s from uniform with a width s + 1
branching program, and there exist ε-bias distributions with support size O(n/ε 2 ), the sum of two such
distributions can be distinguished from uniform in space O(c log n) when ε = n−c . Actually, both their
proof and ours (presented later) apply to c > 0.01; but for smaller c Theorem 1.4.iv kicks in.
    Theorems 1.4.iii and 1.4.iv come close to addressing the “RL question,” without answering it:
Theorem 1.4.iv shows that polynomial bias is necessary even for width-3 regular branching programs,
while Theorem 1.4.iii shows that the bias is at least polynomial in the width. [22] prove Theorem 1.4.iv
with ε = Ω(1).
    We have not been able to say anything on the “Polynomials question” (Question 1.3).
     There exist other models of interest. For read-once DNFs no counterexample with large error is
possible because Chari, Rohatgi, and Srinivasan [12], building on [15], show that (just one) n−O(log(1/δ )) -
bias distribution fools any read-once DNF on n variables with error δ , cf. Section 5. The [12] result is
rediscovered by De, Etesami, Trevisan, and Tulsiani [14], who also show that it is essentially tight by
constructing a distribution which is n−Ω(log(1/δ )/ log log(1/δ )) -biased yet does not δ -fool a read-once DNF.
In particular, fooling with polynomial error requires super-polynomial bias.
     It would be interesting to know whether the XOR of two copies overcomes this limitation, i. e., if it
δ -fools any read-once DNF on n variables provided each copy has bias poly(δ /n). If true, this would
give a generator with seed length O(log(n/δ )), which is open.
     We are unable to resolve this for read-once DNFs. However, we show that the corresponding result
for general DNFs would resolve long-standing problems on circuit lower bounds [35]. This can be
interpreted as saying that such a result for DNFs is either false or extremely hard to prove. We also get
conditional counterexamples for depth-3 and AC0 circuits.

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                  3
                                  C HIN H O L EE AND E MANUELE V IOLA

Theorem 1.5. Suppose polynomial time (P) has fan-in 2 circuits of linear size and logarithmic depth.
Then Theorem 1.4 also applies to the following choices of parameters:
    i. ε = n−ω(1) , f is a depth-3 circuit of size no(1) and unbounded fan-in, and p = n−ω(1) .
   ii. ε = n−ω(1) , f is a DNF formula of size poly(n), and p = 1 − 1/no(1) .
Moreover, all our results extend to more copies of D as follows. The input D + D to f can be replaced by
the bitwise XOR of k ≤ log n independent copies of D if ε is replaced by ε 2/k .

    Recall that it is still open whether NP has linear-size circuits of logarithmic depth.

Theorem 1.6. Suppose for every δ > 0 there exists a constant d such that NC2 has AC0 circuits of size
                                                                                                   c
2n and depth d. Then Theorem 1.4 also applies to the following choice of parameters: ε = n− log n , f is
  δ

                                                            Ω(1)
an AC0 circuit of size nO(c) and depth O(c) , and p = n− log n/4 .
    Moreover, our result extends to more copies of D as follows. The input D + D to f can be replaced by
the bitwise XOR of k ≤ logc+1 n/6(c + 1) log log n independent copies of D if ε is replaced by ε 2/k .

    Recall that the assumption in Theorem 1.6 holds for NC1 instead of NC2 , in fact it holds even for log-
                                                                                                        O(d)
space. Moreover, the parameters in the conclusion of Theorem 1.6 are tight in the sense that n−(log n)
bias fools AC0 circuits of size nd and depth d, as shown in the sequence of works [4, 29, 10, 34].
    All the above results except Theorem 1.4.iv are based on a new, simple connection between small-bias
generators and error-correcting codes, discussed in Section 1.2.

Definition 1.7. A distribution over X n is k-wise independent if its marginal distribution on any k positions
is uniformly distributed.

Definition 1.8. Let S ⊆ F p be a set of vectors. Define the mod p rank of S, denoted by rank p (S) to be
the rank of S over F p . We define the mod p rank of a distribution D to be the mod p rank of its support.

Definition 1.9. The correlation of two functions f , g : {0, 1}n → {0, 1} is

                                  Cor( f , g) := Ex∼{0,1}n [(−1) f (x)+g(x) ] .

    Theorem 1.4.iv instead follows [22] and bounds the mod 3 rank of small-bias distributions. It turns
out an upper bound on the mod 3 rank of some k-wise independent distributions over bits would allow
us to reduce the bias in Theorem 1.4.iv, assuming long-standing conjectures on correlation bounds for
low-degree polynomials (which may be taken as standard).

Claim 1.10. Suppose
   1. the parity of k copies of mod 3 on disjoint inputs of length m has correlation 2−Ω(k) with any
                                  √
      F2 -polynomial of degree ε m for some constant ε > 0, and
   2. for every c, there exists an n−c -almost c log n-wise independent distribution whose support on
      {0, 1}n ⊆ Fn3 = {0, 1, 2}n has mod 3 rank at most n0.49 .
Then the “RL question” has a negative answer, i. e., for every c, there exists an n−c -biased distribution D
such that D + D does not fool a one-way O(log n)-space algorithm. More specifically, D + D does not
fool a mod 3 linear function.

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                               4
                    S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

    Contrapositively, an affirmative answer to the “RL question,” even for permutation, width-3 branching
programs, implies lower bounds on the mod 3 rank of k-wise independent distributions, or that the
aforementioned correlation bounds are false.
    What we know about the second assumption in Claim 1.10 is in Section 3, where we initiate a
systematic study of the mod 3 rank of (almost) k-wise independent distributions, and obtain the following
lower and upper bounds. First, we give an Ω(k log n) lower bound on the mod 3 rank for almost k-wise
independent distributions, specifically, distributions such that any k coordinates are 1/10 close to being
uniform over {0, 1}k (Claim 3.1). This also gives an exponential separation between mod 3 rank and seed
length for such distributions.
    We then prove the following upper bounds, see Claim 3.9.

Theorem 1.11. For infinitely many n, there exist k-wise independent distributions over {0, 1}n with
mod 3 rank d for k = 2 and d ≤ n0.73 .

     We note that an upper bound of n − 1 on the mod 3 rank of a k-wise independent distribution implies
that the distribution is constant on a mod 3 linear test. We ask what is the largest k∗ = k∗ (n) such
that there exists a k-wise independent distribution with mod 3 rank ≤ n − 1. We conjectured the bound
k∗ (n) = Ω(n). Partial progress towards this conjecture appeared in a preliminary version of this paper [20].
This conjecture was later verified [9].

1.2   Our techniques
All our counterexamples in Theorem 1.4 and 1.5, except Theorem 1.4.iv, come from a new connection
between small-bias distributions and linear codes, which we now explain. Let C ⊆ Fn be a linear error
correcting code over a finite field of characteristic 2. (Using characteristic 2 allows us to work with
small-bias over bits, as opposed to large alphabets, which makes things slightly simpler.) We also use C
to denote the uniform distribution over the code C. It is well-known that if C⊥ has minimum distance d ⊥ ,
then C is (d ⊥ − 1)-wise independent.
    Define Ne to be the “noise” distribution over Fn obtained by repeating the following process e times:
Pick a uniformly random position from [n], and set it to a uniform symbol in F. Now, define De to be the
distribution on n log|F| bits obtained from adding Ne to C, and we have the following fact.

Fact 1.12. De is (1 − d ⊥ /n)e -biased.

Proof. If a test is on less than d ⊥ field elements, De has zero bias because it is (d ⊥ − 1)-wise independent.
Otherwise, the bias is nonzero only if none of the symbols touched by the test are hit by random noise,
which happens with probability (1 − d ⊥ /n)e .

   Our main observation is that the XOR of two noisy codewords is also a noisy codeword, with the
number of errors injected to the codeword doubled. That is,

                              De + De = C + Ne +C + Ne = C + N2e = D2e .

Definition 1.13. An algorithm is a threshold-e discriminator for the code C if it decides whether a string
is within Hamming distance e of the code.

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                5
                                  C HIN H O L EE AND E MANUELE V IOLA

    Now suppose an algorithm is a threshold-2e discriminator for C. Then it can be used to distinguish
De + De from uniform. More generally, if an algorithm is a threshold-ke discriminator for C, then it can
distinguish the XOR of k independent copies of De from uniform. Contrapositively, if De + De fools f ,
then f is not a threshold-2e discriminator for C. Thus, to obtain counterexamples we only have to exhibit
an appropriate threshold discriminator. We achieve this by drawing from results in coding theory. This is
explained below after two remarks.
Remark 1.14. Our threshold discriminator is only required to tell apart noisy codewords and uniform
random strings. This is a weaker condition than decoding. In fact, similar threshold discriminators have
been considered in the context of tolerant property testing [17, 19, 30], where tolerant testers are designed
to decide if the input is close to being a codeword or far from every codeword, by looking at as few
positions of the input as possible.
Remark 1.15. We note that our connection between ε-bias distributions and linear codes is different
from the well-known connection in [24], which shows that for a binary linear code with relative minimum
and maximum distance ≥ 1/2 − ε and ≤ 1/2 + ε, respectively, the columns of its k × n generator matrix
form the support of an ε-biased distribution over {0, 1}k . However, the connection to codes is lost once
we consider the sum of the same distributions. In contrast, the sum of our distributions bears the code
structure of a single copy.
     As hinted before Fact 1.12, the small-bias property is established through a case analysis based on the
weight of the test. This paradigm goes back at least to the original work by Naor and Naor [24]. It was
used again more recently in [23, 3]. Our reasoning is especially close to [23, 3] because in both papers
small tests are handled by local independence but large tests by sum of independent biased bits.
     For general circuits (Theorem 1.4.i), we consider the asymptotically good binary linear code with
constant dual relative distance, based on algebraic geometry and exhibited by Guruswami in [32]. We
conjecture that the corresponding threshold discriminator can be implemented in NC2 . However, we are
unable to verify this. Instead, for NC2 circuits (Theorem 1.4.ii), we use Reed–Solomon codes and the
Peterson–Gorenstein–Zierler syndrome-decoding algorithm [27, 16] which we note is in NC2 . Under the
assumption that NC2 is contained in AC0 circuits of size 2n , by scaling the NC2 result down to polylog n
                                                                δ


bits followed by a depth reduction, we obtain our results for AC0 circuits (Theorem 1.6). This result could
also be obtained by scaling down a result in [7].
     Our counterexample for one-way log-space computation (Theorem 1.4.iii) also uses Reed–Solomon
codes. The threshold discriminator is simply syndrome decoding: To decode from e errors it can be
realized by computing the syndrome in a one-way fashion using space O(e log q), where q is the size of the
underlying field of the code. For a given constant c, setting q = n, message length k = d ⊥ − 1 = n − O(c),
and e = O(c) we obtain a one-way space O(c log n) distinguisher for the sum of two distributions with
bias n−c .
     Naturally, one might try to eliminate the dependence on c in the O(c log n) space bound with a different
choice of e and q, which would answer the “RL question” in the negative. In Claim 2.2 however we show
that to obtain n−c bias, the space O(e log q) for syndrome decoding must be of Ω(c log n), regardless
of the code and the alphabet. Thus our result is the best possible that can be obtained using syndrome
decoding. We raise the question of whether syndrome decoding is optimal for one-way decoding in this
setting of parameters, and specifically if it is possible to devise a one-way decoding algorithm using space

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                               6
                     S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

o(e log q). There do exist alternative one-way decoding algorithms, cf. [30], but apparently not for our
setting of parameters of e = O(1) and k = n − O(1).
     Our conditional result for depth-3 circuits and DNF formulas (Theorem 1.5) follows from scaling
down to barely superlogarithmic input length, and a depth reduction [35] (cf. [38, Chapter 3]) of the
counterexample for general circuits (Theorem 1.4.i). We note that the 2−Ω(n) -bias in Theorem 1.4.i
is essential for this result, in the sense that 2−n/ log n -bias would be insufficient to obtain Theorem 1.5.
We also remark that since O(log2 n)-wise independence suffices to fool DNF formulas [4], one must
consider linear codes with dual distance less than log2 n in our construction, and so De has bias at
                                  2                                                 2
least (1 − log2 n/n)e = 2−O(log n) . On the other hand, [14] shows that 2−O(log n log log n) -bias fools DNF
formulas.
    The connection between codes and small-bias distributions motivate us to study further the complexity
of decoding. [37, Chapter 6] and [31], cf. [37, Chapter 6], show that list-decoding requires computing
the majority function. In Claim 4.2 we extend their ideas and prove that the same requirement holds
even for decoding up to half of the minimum distance. This gives some new results for AC0 and for
branching programs. Finally, since logO(1) n-wise independence fools AC0 [10, 34], we obtain that AC0
cannot distinguish a codeword from a code with logΩ(1) n dual distance from uniform random strings.
This also gives some explanation of why scaling is necessary to obtain Theorem 1.6 from Theorem 1.4.ii.

A different approach. We now explain the high-level ideas in proving Theorem               1.4.iv. Meka and
Zuckerman [22] construct the following constant-bias distribution D over n := d5 bits with mod 3 rank
          √
less than n. Each output bit is the square of the mod 3 sum of 5 out of the d uniform random bits, which
can be written as a degree-5 polynomial over F2 . Since any parity of the output bits is also a degree-5
polynomial over {0, 1}d , D has constant bias. To show that a mod 3 linear function is always 0 on the
                                                                                                          √
support of D + D, they observe that for sufficiently large n, D has mod 3 rank at most d 2 < n, and
D + D has mod 3 rank at most (d 2 )2 + d 2 = d 4 + d 2 < n. (See Fact 2.4.)
     We extend their construction using ideas from the Nisan generator [25]: We pick a pseudo-design
consisting of n sets where each set has size nβ (we will choose β to be a small constant), and the
intersection of any two sets has size O(log n). Such pseudo-design exists provided the universe has size
n2β . The output distribution is again the square of the mod 3 sum on each set.
     For any test of size at least C log n bits, let J be any C log n bits of the test. We fix the intersections of
their corresponding sets in the universe to make them independent. After we do this, every bit in J is still
a mod 3 function on nβ − |J| log n ≥ 0.9nβ bits.
     We further fix every bit outside the |J| sets in the universe. This will not affect the bits in J. Now
consider any bit b in the test that is not in J, it corresponds to a set which has intersection at most
log n with each of the sets that correspond to the bits in J. Thus, b is now a mod 3 function on at most
|J| log n = log2 n input bits and thus can be written as a degree-log2 n polynomial over F2 . Hence, the
parity of the bits outside J is also an F2 -polynomial of the same degree, and we call this polynomial p.
     Now observe that the bias of the test equals to the correlation between the parity of the bits in J
and p. Since each bit in J is a mod 3 function on nβ bits, by Smolensky’s theorem [33], it has constant
correlation with p. In Lemma 2.9 we prove a variant of Impagliazzo’s XOR lemma [18] to show that the
XOR of log n independent such bits makes the correlation drop from constant to ε = n−β /4 . This variant
of XOR lemma may be folklore, but we are not aware of any reference.

                         T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                   7
                                   C HIN H O L EE AND E MANUELE V IOLA

    This handles tests of size at least C log n. For smaller tests, the above distribution could have constant
bias, and hence we XOR it with an 1/nΩ(1) -almost C log n-wise independent distribution, which gives us
ε bias for tests of size less than C log n and has sufficiently small rank. We then show that the XOR of the
                                        √
two distributions has rank less than n and conclude as in the previous paragraph.
    We refer the reader to [40] for background on XOR lemmas.


Organization. In Section 2 we describe our counterexamples and prove Theorem 1.4 and 1.5, and
Claim 1.10. In Section 3 we prove our lower bounds and upper bounds on the mod 3 rank of k-wise
independence. As a bonus, in Section 4 we include some results on the complexity of decoding. For
example, we show that for codes with large minimum distance, AC0 circuits and read-once branching
programs cannot decode when the number of errors is close to half of the minimum distance of a code
(Claim 4.2). We obtained these results while attempting to devise low-complexity algorithms that can
decode (which, by our connection, would have consequences for the sum of small-bias generators).


2     Our counterexamples
We are now ready to prove Theorem 1.4 and 1.5, and Claim 1.10. We consider linear codes with different
parameters, the bias of D follows from Fact 1.12. Then we present our distinguishers. In the end, we
explain how our results hold for k copies instead of 2.


2.1   General circuits
Venkatesan Guruswami [32] exhibits the following family of constant-rate binary linear codes whose
primal and dual relative minimum distance are both constant.

Theorem 2.1 (Theorem 4 in [32]). For infinitely many n, there exists a binary linear code C with block
length n and dimension n/2, which can be constructed, encoded, and decoded from n/60 errors in time
poly(n). Moreover, the dual of C has minimum distance at least n/30.

Proof of Theorem 1.4.i. Applying Fact 1.12 with e = n/120 to the code in Theorem 2.1, we obtain a
distribution D that is 2−n/3600 -biased. Our threshold-2e discriminator f for the code C decodes and
encodes the input, and accepts if and only if the input and the re-encoded string differ by at most 2e
positions. Since both the encoding and decoding algorithms run in polynomial time, so does f .
    Note that f accepts at most

                                        2e  
                                             n
                                  2n/2 · ∑     ≤ 2n/2 · 2nH(1/60) ≤ 20.75n
                                       i=0   i

possible strings, where H(·) is the binary entropy function (cf. [13, Example 11.1.3] for the first
inequality). Hence, f distinguishes D + D from the uniform distribution with probability at least 1 −
2−0.25n .

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                8
                     S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

2.2 NC2 circuits
Proof of Theorem 1.4.ii. Let q be a power of 2. Consider the Reed–Solomon code C over Fq with block
length q − 1, dimension q/2 and minimum distance q/2. C has dual minimum distance q/2 + 1 and
can decode from q/4 errors. Applying Fact 1.12 to C with e = q/12, we obtain a distribution D over
n := (q − 1) log q bits that is 2−Ω(n/ log n) -biased.
    Let α be a primitive element of Fq . Let H be a parity-check matrix for C. We first recall the
Peterson–Gorenstein–Zierler syndrome-decoding algorithm [27, 16].
    Given a corrupted codeword y, let (s1 , . . . , sq/2 )T := Hy be the syndrome of y. Suppose y has v < q/2
errors. Let E denote the set of its corrupted positions. Let
                                                                         v
                                     Λv (x) := ∏(1 − α i x) = 1 + ∑ λi xi
                                                i∈E                     i=1

be the error locator polynomial. The syndromes and the coefficients of Λv are linearly related by

                                λv s j−v + λv−1 s j−v+1 + · · · + λ1 s j−1 + s j = 0 ,

for j > v. This forms a linear system with unknowns λi . The algorithm decodes by attempting to solve
the corresponding linear systems with v errors, where v ranges from 2e to 1.
    Note that the system has a unique solution if and only if y and some codeword differ by exactly v
positions, for some v between 1 and 2e. Thus, f computes the determinants of the 2e < q/4 systems
and accepts if and only if one of them is nonzero. Since computing determinant is in NC2 [6], f can be
computed by an NC2 circuit. The system always has a solution when inputs are under D + D and so f
always accepts. On the other hand, f accepts at most
                                2e      
                                    q−1
                          q/2
                         q ·∑              (q − 1)i ≤ qq/2 · 2qhq (1/6) ≤ 22n/3+o(n)
                               i=0    i

possible strings, where

                           hq (x) := x logq (q − 1) − x logq x − (1 − x) logq (1 − x)

is the q-ary entropy function. Therefore, f distinguishes D + D from the uniform distribution with
probability at least 1 − 2−n/4 .

2.3   One-way log-space computation
Proof of Theorem 1.4.iii. Let q be a power of 2. Consider the [q − 1, q − 6c, 6c]2log q Reed–Solomon code
C over F2log q , which has dual minimum distance q − 6c + 1 and can decode from 3c errors. Applying
Fact 1.12 to C with e = c, we obtain a distribution D over n := (q − 1) log q bits that is O(c log n/n)c -
biased.
    Let H be a parity-check matrix of C. On input y ∈ Fq2log q , our distinguisher f computes s2e+1 , . . . , s4e
from the syndrome s := Hy. Clearly this can be implemented in one-pass and space (2e + O(1)) log q.
Finally, using the Peterson–Gorenstein–Zierler syndrome-decoding algorithm, f accepts if and only if y
differs from a codeword of C by at most 2e positions.

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                  9
                                      C HIN H O L EE AND E MANUELE V IOLA

      Since f accepts at most
                                     2e      
                                           q−1
                          q   q−6c
                                     ·∑        (q − 1)i ≤ qq−6c · 2q4c ≤ O(qq−2c )
                                     i=0    i

strings, f distinguishes D + D from uniform with probability 1 − O(log n/n)2c .

   Computing the input for syndrome decoding requires space (2e + O(1)) log q. We now show that in
order to obtain n−c bias via our construction, we always have 2e log q = Ω(c log n). Thus, one cannot
answer the “RL question” in the negative via syndrome decoding.

Claim 2.2. For every q ≥ n + 1, let C be an [n, k, d] code over Fq which decodes from e errors, and d ⊥
be its dual minimum distance. If C satisfies (1 − d ⊥ /n)e < q−c for sufficiently large c, then we have
e log q = Ω(c log n).

Proof. If d ⊥ > (1 − 1/q)n, then by the Plotkin bound on the dual code, n − k = O(1). By the Singleton
bound, e ≤ d ≤ n − k and so we have e = O(1). Hence, (1 − d ⊥ /n)e = (1/q)e ≥ q−c for sufficiently large
c, and therefore the condition is not satisfied.
    On the other hand, suppose d ⊥ ≤ (1 − 1/q)n. Then (1 − d ⊥ /n)e ≥ (1/q)e . The condition (1 −
d /n)e < q−c implies e log q > c log q > c log n.
 ⊥



2.4     Depth 3 circuits, DNF formulas and AC0 circuits
Proof of Theorem 1.5. We will use Valiant’s depth reduction [35, 36] (cf. [38, Theorem 25]).

Theorem 2.3 ([35, 36]). Let C : {0, 1}n → {0, 1} be a circuit of size cn, depth c log n and fan-in 2. The
                                                                                              0
function computed by C can also be computed by an unbounded fan-in circuit of size 2c n/ log log n and
depth 3 with inputs x1 , x2 , . . . , xn , x1 , x2 , . . . , xn , where c0 depends only on c.

    By the assumption that P has fan-in 2 circuits of linear size and logarithmic depth and the fact that f
in Theorem 1.4.i is in P, we can apply Theorem 2.3 to f and obtain an unbounded fan-in depth-3 circuit f 0
of size 2O(n/ log log n) that computes the same function. Then we scale down n to n0 = log n log log log log n
bits (we set the rest of the n − n0 bits uniformly at random) to get an n−ω(1) -biased distribution Dn0 and
a circuit fn0 0 of size no(1) and depth 3 that distinguishes Dn0 + Dn0 from uniform with probability at least
1 − n−ω(1) . This proves Theorem 1.5.i.
    To prove Theorem 1.5.ii, note that fn0 0 accepts with probability 1 under Dn0 + Dn0 and without loss of
generality we can assume fn0 0 is an AND-OR-AND circuit. Hence, it contains a DNF f 00 such that (1) f 00
accepts under Dn0 + Dn0 with probability 1, and (2) f 00 rejects with probability at least 1/2no(1) under the
uniform distribution.

Proof of Theorem 1.6. Let D and f be the distribution and distinguisher in Theorem 1.4.ii, respectively.
Let Dn0 and fn0 be the scaled distribution and distinguisher of D and f on n0 = logc+1 n bits, respectively.
                                                                              0    0            c
(We set the rest of the n − n0 bits uniformly at random.) Dn0 has bias 2−Ω(n /log n ) = n−Ω(log n) . By our
                                                                                                  c
assumption, fn0 is in AC0 and distinguishes Dn0 + Dn0 from uniform with probability 1 − n− log n/4 .

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                               10
                      S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

2.5     Mod 3 linear functions
Recall the definition of mod p rank in Definition 1.8.
Fact 2.4 (Lemma 7.1 and 7.2 in [22]). Let S, T be two sets of vectors in Fn3 . Define S2 to be the set
{x ×3 x : x ∈ S}, where x ×3 y denote the pointwise product of two vectors x and y (over F3 ). Then
  (1) rank3 (S2 ) ≤ rank3 (S)2 ;
  (2) when S and T are subsets in {0, 1}n ⊆ Fn3 ,

                             rank3 (S +2 T ) ≤ rank3 (S) + rank3 (T ) + rank3 (S) rank3 (T ) .

        If S = T , then rank3 (S +2 S) ≤ rank3 (S) + rank3 (S)2 .
Proof. Let dS := rank3 (S) and dT := rank3 (T ). Let {β1 , . . . , βdS } be a basis of S and {γ1 , . . . , γdT } be a
basis of T . Let
                                           dS                              dT
                                     x = ∑ ci βi ∈ S            and    y = ∑ r jγ j ∈ T
                                          i=1                              j=1

be any two vectors. We have
                                           x ×3 x =         ∑ ci c j (βi ×3 β j ) .
                                                          i, j∈[dS ]

Thus {βi ×3 β j }i, j∈[dS ] spans S2 , proving (1). For (2), observe that for any a, b ∈ {0, 1} ⊆ F3 , we have
a +2 b = a +3 b +3 a ×3 b. Hence we have
                                                 dS            dT
                       x +3 y +3 x ×3 y = ∑ ci βi + ∑ r j γ j +                 ∑           ci r j (βi ×3 γ j ) ,
                                                i=1            j=1       i∈[dS ], j∈[dT ]

and thus
                                  {βi }i∈[dS ] ∪ {γ j } j∈[dT ] ∪ {βi ×3 γ j }i∈[dS ], j∈[dT ]
spans S +2 T .

      The following lemma is well-known (cf. [25]). We include a proof here for completeness.
Lemma 2.5. There exists a pseudo-design (S1 , . . . , Sn ) over the universe [d] such that
   1. |Si | = t for every i ∈ [d], and
   2. |Si ∩ S j | ≤ tˆ for every i 6= j ∈ [d],
where d = n2β ,t = nβ , and tˆ = log n for any β < 0.5.
      We will use the following Chernoff bound in the proof.
Claim 2.6 (Chernoff bound). Let X1 , . . . , Xn ∈ {0, 1} be n independent and identically distributed vari-
ables with E[Xi ] = µ for each i. We have
                                       "                   #
                                                      n
                                                                                 2
                                         Pr      ∑ Xi − µn ≥ t ≤ e−t /4n .
                                                 i=1



                         T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                      11
                                    C HIN H O L EE AND E MANUELE V IOLA

Proof of Lemma 2.5. It suffices to show that given S1 , . . . , Si−1 , there exists a set S such that |S| ≥ t and
|S ∩ S j | ≤ tˆ for j < i. Consider picking each element in [d] to be in S with probability p = 0.1 log n/nβ .
We have E[|S|] = pd ≥ 2nβ . By the Chernoff bound,

                                         Pr[|S| < t = nβ ] ≤ 2−n /4 < 1/2 .
                                                                   β




We also have E[|S ∩ S j |] = pt = 0.1 log n. Again by the Chernoff bound,

                                 Pr[|S ∩ S j | > tˆ = log n] ≤ 2−4 log n < 1/2n .

It follows by a union bound that with nonzero probability there is an S which satisfies the two conditions
above.

Proof of Theorem 1.4.iv. Let α < 1/36 and β = 4α. Also let d,t, tˆ be the parameters and S1 , . . . , Sn be
the pseudo-design specified in Lemma 2.5. Define the function L : {0, 1}d → {0, 1}n whose i-th output
bit yi equals
                                                       2
                                  mod 23 (xSi ) := ∑ x j mod 3 .
                                                            j∈Si

Let T1 be the image set of L. Without the square, this set has mod 3 rank d and so by Fact 2.4,
rank3 (T1 ) = O(d 2 ) = O(n16α ). Let T2 be an ε-almost k-wise independent set, where ε = 1/nα and
k = 2 log n. Known constructions [2, Theorem 2] (see also [24]) produce such a set of size O((k log n)/ε)2
and therefore rank3 (T2 ) is at most O(n2α log4 n).
    Consider the set T := T1 +2 T2 . By Fact 2.4, T has rank at most O(n18α log4 n). By the same fact,
T +2 T has rank at most O(n36α log8 n) < n because α < 1/36. Therefore, there is a non-zero mod 3
linear function ` such that `(y) ≡ 0 (mod 3) for any y ∈ T , while Pr[`(y) = 0] ≤ 1/2 for a uniform y in
{0, 1}n . It remains to show that T is O(1/n0.99α )-biased. For any test on I ⊆ [n], we consider the cases
(1) when |I| ≤ k, and (2) when |I| > k separately.
    Write y = y1 + y2 , where y1 ∈ T1 and y2 ∈ T2 . Case (1) follows from the fact that T2 is 1/nα -almost
k-wise independent. Case (2) follows from the following claim.

Claim 2.7. For any |I| > k, we have |Ey1 ∈T1 [χI (y1 )]| ≤ O(1/n0.99α ), where χI (z) := (−1)∑i∈I zi .

Proof. Pick a subset J ⊆ I of size k. Define f , p : {0, 1}n → {0, 1} to be

                          f (x) := ∑ mod 23 (xSi ) and          p(x) := ∑ mod 23 (xSi ) ,
                                   i∈J                                    i∈I\J

respectively. Observe that

                                Exi :i∈[d] [χI (y1 )] = Exi :i∈[d] [(−1)( f (x)+p(x)) ] ,

which is the correlation between f and p.
    Consider the sets S j ⊆ [d] with j ∈ J. Let B1 be the set of indices appearing in their pairwise
intersections. That is, B1 := {` ∈ [d] : ` ∈ Si ∩ S j for some distinct i, j ∈ J}. Fixing the value of every
x` ∈ B1 , each mod 23 (xS j ) in f becomes a function on m := nβ − tˆ · k ≥ 0.9nβ bits.

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                  12
                     S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

    Let B2 be the set of indices in [d] outside the S j for j ∈ J. The bits in B2 do not affect the outputs in
J. Fixing their values, each mod 23 (xS j ) in p is a function of at most tˆ · k = O(log2 n) bits and so can be
written as a polynomial of degree O(log2 n) over F2 . Since p is a parity of values mod 23 (xS j ), it can also
be written as a polynomial of degree O(log2 n) over F2 .
    We will use the following theorem by Smolensky [33] (cf. [38, Chapter 1]). The proof in [38] has the
condition that n is divisible by 3. This condition can be removed. For example, when n = 3` + 1, we can
set a random bit of the uniform distribution to zero. This distribution is close to uniform, but now we can
apply [38] as stated.
Theorem 2.8 ([33]). There exists an absolute constant ε > 0 such that for every n that is divisible by 3
                                                         √
and for every polynomial p : Fn2 → F2 of degree at most ε n, we have
                                                        2
                                       Ex∼{0,1}n (−1)mod3 (x)+p(x) ≤ 0.9 .
                                                                 


   To build intuition, note that after fixing the input bits in B1 and B2 , for each of the mod 23 (xS j ) in f ,
by Theorem 2.8 we have
                                                    (mod23 (xS j )+p(x))
                                        h                    i
                                  Exi :i∈[d] (−1)                          ≤ 1 − Ω(1) .

In the following lemma we prove a variant of Impagliazzo’s XOR Lemma [18] to show that

                       Exi :i∈[d] (−1)( f (x)+p(x)) ≤ O(1/m0.249 ) = O(1/n0.99α ) .
                                                  


Averaging over the values of the xk in B1 and B2 finishes the proof.

Lemma 2.9. Let k = 2 log m, define f : {0, 1}m×k → {0, 1} by

                           f (x(1) , . . . , x(k) ) := mod 23 (x(1) ) +2 · · · +2 mod 23 (x(k) ) .

Let p : {0, 1}m×k → {0, 1} be any polynomial of degree O(log2 m). We have

                        Cor( f , p) := Ex∼{0,1}m×k (−1) f (x)+p(x) ≤ O(1/m0.249 ) .
                                                                 


Proof. We will use the fact that Theorem 2.8 holds for degree nΩ(1) polynomials to get correlation 1/nΩ(1)
for polynomials of degree polylog(n).
    As in the proof in [18] we first show the existence of a measure M : {0, 1}m → [0, 1] of size at most
|M| := ∑x M(x) = 2m /4 such that with respect to its induced distribution D(x) := M(x)/|M|, the function
mod 23 is 1/2m0.249 -hard for any polynomial p of degree O(log2 m), i. e.,

                                   Pr [mod23 (x) = p(x)] ≤ 1/2 + 1/4m0.249 .
                                  x∼D

Suppose not. Lemma 1 in [18] implies that one can obtain a function q by taking the majority of O(m0.498 )
polynomials of degree O(log2 m) such that

                                           Pr      [mod23 (x) = q(x)] > 0.99 .
                                        x∼{0,1}m


                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                  13
                                          C HIN H O L EE AND E MANUELE V IOLA

Note that q can be represented as a degree O(m0.498 log2 m) polynomial. By Theorem 2.8,
                                                 Pr       [mod23 (x) = p(x)] ≤ 0.95
                                               x∼{0,1}m

for any degree εm1/2 polynomial p, a contradiction.
    Now we show that there is a set S ⊆ {0, 1}m of size 2m /8 such that mod 23 is 1/m0.249 -hard-core on S
for any polynomial p of degree O(log2 m), i. e.,
                                         Pr [mod32 (x) = p(x)] ≤ 1/2 + 1/2m0.249 .
                                         x∼S

Let p be any degree-O(log2 m) polynomial. For any measure M : {0, 1}m → [0, 1], define Adv p (M) by
                                                                             2
                                          Adv p (M) := ∑ M(x)(−1)mod3 (x)+p(x) .
                                                             x
We construct S probabilistically by picking each x to be in S with probability M(x). Let MS be the
indicator function of S. Then ES [Adv p (MS )] = Adv p (M) ≤ |M|/(2m0.249 ). Note that Adv p (MS ) is the
sum of 2m independent random variables, where each variable is over [−1, 0] or [0, 1]. By Hoeffding’s
inequality,
                                                         2   m   0.498     m   0.498
                   Pr[Adv p (MS ) > |M|/m0.249 ] ≤ 2−2|M| /(2 ·4m ) = 2−2 /32m .
                         S
                             O(log2 m)
Note that there are 2m          polynomials of degree log2 m. Moreover, since ES [|S|] = 2m /4, again by
Hoeffding’s inequality, PrS [|S| < 2m /8] ≤ 1/2. Hence, by a union bound, the required S exists.
    It follows that there exists a set of inputs S ⊆ {0, 1}m of size at most 2m /8 such that mod 23 is 1/m0.249 -
hard-core on S for any polynomial of degree O(log2 m). Now we apply the following lemma, which is
stated in [18] for circuits, but the same proof applies to polynomials.
Lemma 2.10 (Lemma 4 in [18]). If g is ε-hard-core for some set of δ 2n inputs for polynomials of degree
d, then the function f (x(1) , . . . , x(k) ) := g(x(1) ) +2 · · · +2 g(x(k) ) is ε + (1 − δ )k -hard-core for polynomials
of the same degree.
    Applying this lemma with our choice of k, we have for any polynomial p of degree O(log2 m),
                     Pr[ f (x) = p(x)] ≤ (1 + 1/m0.249 + (7/8)k )/2 = 1/2 + O(1/m0.249 ) .
                     x

Hence f is O(1/m0.249 )-hard for any polynomial of degree O(log2 m), and the lemma follows.
Proof of Claim 1.10. We replace the pseudo-design in the proof of Theorem 1.4.iv with one that has set
size t = O(log4 n) and intersection size tˆ = O(log n). Using the same idea as in the proof of Lemma 2.5
one can show that such pseudo-design exists provided the universe is of size d = O(log8 n). Now, using
the same argument, for tests of size larger than c log n, we apply (1) to f and p, which are the parity of
c log n copies of mod 3 function on m = O(log4 n) bits and a polynomial of degree O(log2 n), respectively.
This gives bias O(1/nc ). Note that the image set T1 now has mod 3 rank d 2 = O(log16 n).
    For tests of size at most c log n, we replace the almost k-wise independent set with the n−c -almost
k-wise independent distribution given by (2), which has bias n−c , and we denote the support of the
distribution by T2 .
    By Fact 2.4, T := T1 +2 T2 has mod 3 rank O(n0.49 log16 n) = o(n0.5 ). Hence, T +2 T has rank less
than n and the claim follows.

                             T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                     14
                    S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

2.6   Sum of k copies of small-bias distributions
We now show that the results hold for k copies when ε is replaced by ε 2/k , proving the “Moreover” part
in Theorem 1.4, 1.5 and 1.6.

Proof of “Moreover” part in Theorem 1.4, 1.5 and 1.6. To prove Theorem 1.4.i, 1.4.ii and 1.4.iii, we
can replace e by 2e/k in their proofs to obtain distributions D0 that are ε 2/k -biased. Since we have to
throw in at least one error, 2e/k ≥ 1. The rest follows by noting the sum of k copies of D0 is identical to
D + D.
     By scaling down the above small-bias distributions D0 for Theorem 1.4.i and 1.4.ii to n0 bits as in the
proofs of Theorem 1.5 and 1.6, respectively, we obtain ε 2/k -biased distributions D0n0 so that the sum of k
copies of D0n0 is identical to Dn0 + Dn0 in Theorem 1.6 and 1.5. Moreover, k scales from k(n) to k(n0 ).
     For Theorem 1.4.iv, let α := log(1/ε)/ log n and so ε 2/k = n−2α/k . We set β = 8α/k instead of 4α
in the construction of T1 and replace T2 by an n−2α/k -almost 2 log n-wise independent set in the proof, and
call them T10 and T20 , respectively. We now have rank3 (T10 ) = O(n32α/k ) and rank3 (T20 ) = O(n4α/k log4 n).
Thus, the set T 0 := T10 +2 T20 has rank at most O(n36α/k log4 n) and therefore the sum of k copies has rank
at most rank3 (T 0 )k = O(n36α log4k n) < n, for k < O(log n/ log log n). The bias of T 0 follows from the
facts that T20 has bias n−2α/k against tests of size at most 2 log n, and T10 has bias O(n−2α/k ) for tests of
size greater than 2 log n.


3     Mod 3 rank of k-wise independence
In this section, we begin a systematic investigation on the mod 3 rank of k-wise independent distributions.
    Recall Definition 1.8 of mod p rank. We also define the mod p rank of a matrix over the integers to
be its rank over F p . We also write rank p for mod p rank.
    We will sometimes work with vectors over {−1, 1} instead of {0, 1}. Note that the map (1 − x)/2
convert the values 1 and −1 to 0 and 1, respectively, and so the mod 3 rank of a set will differ by at most
1 when we switch vector values from {−1, 1} to {0, 1}, and vice versa.
    While we state our results for mod 3, all the results in this section can be extended to mod p for any
odd prime p naturally.


3.1   Lower bound for almost k-wise independence
In the following claim we give a rank lower bound on almost k-wise independent distributions. Here
“almost” is measured with respect to statistical distance. (Another possible definition is the max bias of
any parity.)

Claim 3.1. Let D be any subset {0, 1}n . If rank3 (D) = t, then D is not 1/10-almost ct/ log(n/t)-wise
independent, for a universal constant c.

   This gives an exponential separation between seed length and rank for almost k-wise independence.
Indeed, for k = O(1), the seed length is Θ(log log n), whereas the rank must be Ω(log n).

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                15
                                  C HIN H O L EE AND E MANUELE V IOLA

Proof. Let C be the span of D over F3 and C⊥ be its orthogonal complement. C⊥ has dimension n − t.
We view C⊥ as a linear code over F3 and let d ⊥ be its minimum distance. Since C⊥ is linear, d ⊥ equals
the minimum Hamming weight of its non-zero elements. Moreover, by the Singleton bound, d ⊥ − 1 ≤ t.
By the Hamming bound, that is,
                                               ⊥
                                           b d 2−1 c  
                                                      n i
                                      3n−t ∑            2 ≤ 3n ,
                                              i=0     i
we have
                                               d⊥ −1            
                                                 b   2   c 
                                                          n
                              t log2 3 ≥ log2  ∑             2i 
                                                  i=0      i
                                                                        
                                                      n             ⊥
                                                                 b d 2−1 c
                                       ≥ log2        ⊥        ·2
                                                  b d 2−1 c
                                                          !b d⊥2−1 c 
                                                                           d⊥ − 1
                                                                                  
                                                    n
                                       ≥ log2      ⊥                 +
                                                b d 2−1 c                    2
                                          ⊥       
                                           d −1              n
                                       ≥             log2 t
                                              2             b c
                                                      n  2
                                       = Ω d ⊥ log2        ,
                                                       t

where we use the fact that d ⊥ − 1 ≤ t in the last inequality. Hence, d ⊥ ≤ O(t/ log(n/t)). Now, let y be
a codeword in C⊥ with Hamming weight d ⊥ . Let I := {i | yi 6= 0}. Note that for every x ∈ D, we have
hy, xi3 = 0 on I. On the other hand, for a uniformly distributed x in {0, 1}I we have hy, xi3 = 0 with
probability at most 1/2. Therefore, D is constant bounded away from uniform on the d ⊥ bits indexed by
I, and thus cannot be close to d ⊥ -wise independent.

3.2   Pairwise independence
We now show that the mod 3 rank of a pairwise independent set can be as small as n0.73 . Then we give
evidence that our approach cannot do any better.

Definition 3.2. We say H is an Hadamard matrix of order n if its entries are ±1 and it satisfies HH T = nIn ,
where In is the n × n identity matrix.

    It is well-known that by removing the all-ones row of an Hadamard matrix H, which can always be
created by multiplying each column by its first element, the uniform distribution over the columns of the
truncated matrix is pairwise independent.
    Henceforth we will work with vectors whose entries are from {−1, 1} = {2, 1} ⊆ F3 . The following
two claims show that certain Hadamard matrices cannot have dimension smaller than n/2. They are taken
from [41], and here we give a self-contained proof for completeness. First, we would give a lower bound
to the mod p rank from the determinant of any square matrix.

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                              16
                     S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

Claim 3.3 (Theorem 1 in [41]). Let A be an n × n matrix over the integers. Assume pe+1 - det(A). Then
rank p (A) ≥ n − e.

Proof. Suppose nullity p (A) = n − r. Let (β1 , . . . , βn−r ) be a basis of the null space of A over F p . Extend
the basis to (β1 , . . . , βn ) so that it forms a basis of Fnp . Let B be the matrix whose columns are the βi . Note
that det(B) 6≡ 0 (mod p) and det(AB) = det(A) det(B). Thus, ps | det(A) if and only if ps | det(AB). By
construction, β1 , . . . , βn−r are in the null space of A over F p and so the first n − r columns of AB are zero
mod p. Hence pn−r | det(A).

Claim 3.4 (Theorem 2 in [41]). Let H be an n × n Hadamard matrix. Let p be an odd prime such that
p2 - n. Then rank p (H) ≥ n/2.

Proof. Since H is an Hadamard matrix, we have HH T = nI and so det(H) det(H T ) = det(H)2 = nn .
Hence |det(H)| = nn/2 . By the condition on p we have that pn/2+1 - nn/2 . Hence, it follows from
Claim 3.3 that rank p (H) ≥ n/2.

    The following claim characterizes Hadamard matrices with mod p rank at most n/2.

Claim 3.5. Let H be an n × n Hadamard matrix. If p | n, then rank p (H) ≤ n/2. Otherwise, rank p (H) = n.

Proof. If p | n, then by Sylvester’s rank inequality, we have

                       rank p (H) + rank p (H T ) − n ≤ rank p (HH T ) = rank p (nIn ) = 0 .

Hence, rank p (H) ≤ n/2, proving the first part. For the second part, if p - n then

                                  det(H)2 = det(H) det(H T ) 6≡ 0 (mod p)

and so det(H) 6≡ 0 (mod p). Hence rank p (H) = n.

    Now we give a construction of Hadamard matrices whose orders do not satisfy the condition in
Claim 3.4. These matrices have much smaller mod p ranks than the lower bound stated in Claim 3.4.
Note that the affine bijection L : {−1, 1}n → {0, 1}n defined by L(v) = (1 − v)/2, where 1 is the all-ones
vector, maps vectors from {−1, 1}n to {0, 1}n . We have the following facts.

Fact 3.6. Let S ⊆ {−1, 1}n be a set containing the all-ones vector. Then rank3 (L(S)) ≤ rank3 (S).

Fact 3.7. If A and B are two Hadamard matrices then A ⊗ B is also an Hadamard matrix, where ⊗
indicates the Kronecker product.

    The following is well-known.

Fact 3.8. Let A, B be two matrices over any field. Then we have rank(A ⊗ B) = rank(A) · rank(B).

Claim 3.9. For infinitely many values of n, there exists a pairwise independent distribution over {0, 1}n
with mod 3 rank at most n0.73 .

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                     17
                                   C HIN H O L EE AND E MANUELE V IOLA

Proof. Paley [26] constructed a (q − 1) × (q − 1) Hadamard matrix for every prime power q ≡ 1 (mod 4).
Starting with an Hadamard matrix H12 over {−1, 1} = {2, 1} ⊆ F3 using Paley’s construction, for every
                                                                  ⊗r
n that is a power of 12, we construct the Hadamard matrix Hn := H12  , where r = log12 n. It follows from
Claim 3.4 and 3.5 that rank3 (H12 ) = 6. Hence, by Fact 3.8, Hn has rank 6log12 n ≤ n0.73 . As discussed
above, we can assume H12 contains an all-ones row. Thus, Hn also contains an all-ones row. and the
claim follows from Fact 3.6.

    We start from an m × m Hadamard matrix with mod 3 rank m/2, for some m. The smaller m we
start from, the better exponent we get. Since Hadamard matrices must be of order 1, 2, or multiple of 4,
Claim 3.4 implies that 12 is indeed the smallest possible m.


4    Complexity of decoding
In this section we prove some negative results on the complexity of decoding.
    In [37, Chapter 6] and [31] it is shown that list-decoding binary codes from error rate 1/2 − ε requires
computing the majority function on 1/ε bits, which implies lower bounds for list-decoding over several
computational models. Using a similar approach, we give lower bounds on the decoding complexity for
AC0 circuits and read-once branching programs. We give a reduction from ε-approximating the majority
function to decoding (1/2 − ε)d errors of a code, where d is the minimum distance.
    Define ε-MAJ to be the promise problem on {0, 1}n , where the YES and NO instances are strings of
Hamming weight at least (1/2 + ε)n and at most (1/2 − ε)n, respectively. We say that a probabilistic
circuit solves ε-MAJ if it accepts every YES instance with probability at least 2/3 and accepts every NO
instance with probability at most 1/3.
    Let C ⊆ {0, 1}n be a code with minimum distance d and let the codewords x and y differ by exactly d
positions, respectively. Define ε-DECODE to be the promise problem on {0, 1}n , where the YES and NO
instances are strings that differ from x and y by at most (1/2 − ε)d, respectively.
    The results in this section have been obtained while attempting to bound the complexity of the
threshold discriminator explained in Section 1.2: distinguish a random noisy codeword from uniform.
The ε-DECODE problem is different, as it asks to distinguish between two noisy codewords.

Lemma 4.1. If a function D : {0, 1}n → {0, 1} solves ε-DECODE, then a restriction of D solves ε-MAJ
on d bits.

Proof. Let x, y ∈ C be the codewords at Hamming distance d. Without loss of generality, we assume x
and y differ in the first d positions. We further assume xi = 0 and yi = 1 for i ∈ [d]. Given an ε-MAJ
instance w of length d, let z be the n-bit string where zi = wi for i ∈ [d] and zi = xi (= yi ) otherwise. If w
has weight at most (1/2 − ε)d, then w and x disagree in at most (1/2 − ε)d positions and therefore D
accepts. Similarly, if w has weight at least (1/2 + ε)d then D rejects.

    Shaltiel and Viola [31] show that depth-c AC0 [⊕] circuits can solve ε-MAJ only if ε is at least
1/O(log n)(c+2) . Brody and Verbin [11] show that ε-MAJ can be solved by a read-once width-w branching
program whenever ε is at least 1/(log n)Θ(w) . Combining these results with Lemma 4.1, we have the
following claim.

                       T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                                18
                    S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

Claim 4.2. Let D : {0, 1}n → {0, 1} be a function.
    1. If D is computable by an AC0 [⊕] circuit of depth c, then it can only solve ε-DECODE with
       ε ≥ 1/O(log d)c+2 .
    2. If D is computable by a read-once width-w branching program, then it can only solve ε-DECODE
       with ε ≥ 1/(log d)Θ(w) .

    We also note the following negative result for decoding by low-degree polynomials.

Claim 4.3. Let C ⊆ {0, 1}n be an [n, k, d] code with dual minimum distance d ⊥ . If
                                                                e/2t−1
                                                             d⊥
                                                        
                                               −t
                                           2        > 16 1 −
                                                              n

for some constant t and e ≤ b d−1
                               2 c, then no degree-t polynomial over F2 can be a threshold-te discrimi-
nator for C.

Proof. Suppose on the contrary a polynomial P is a threshold-te discriminator for C. By Fact 1.12 and the
Schwartz–Zippel Lemma, there exists an ε := (1 − d ⊥ /n)e -biased distribution D such that P distinguishes
the sum of t independent copies of D from uniform with probability at least 2−t . But by [39], the sum of t
                                            t−1
copies of D fools P with probability 16ε 1/2 , a contradiction.


5    Fooling read-once DNF formulas
In this section we state and prove that an m−O(log(1/δ )) bound on the bias suffices to δ -fool any read-once
DNF formulas with m terms. This directly follows from Lemma 5.2 in [12].

Claim 5.1. Let φ be a read-once DNF formula with m terms. For 1 ≤ k ≤ m, every ε-biased distribution
D fools φ with error O(2−Ω(k) + εmk ).
                        Wm
Proof. Write φ (x) :=      i=1 Ci . By Lemma 5.2 in [12], |Prx∼D [φ (x)] − Prx∼{0,1}n [φ (x)]| is upper bounded
by
                                           k                     h^ i            h^ i
                          −k      −k/2e
                      2        +e·e       +∑          ∑       Pr   Ci −     Pr n   Ci .
                                                            x∼D           x∼{0,1}
                                          `=1 S⊆[m]:|S|=`         i∈S               i∈S
                                                          V
The rest follows from the fact that D fools each            i∈S Ci with error ε because it is an AND of AND
terms.

Acknowledgments. A previous version of this paper made an unconditional claim about AC0 . Specifi-
cally, we claimed Theorem 1.6 without the assumption in the first sentence. We are grateful to Andrej
Bogdanov for pointing out this mistake.
    We are grateful to Xue Chen and David Zuckerman for telling us an alternative proof of Theorem 1.4.iii
mentioned in Section 1.1. We also thank Xue Chen for pointing out that the proof of Theorem 1.4.iv was
written with the wrong pseudo-design parameters. We are also grateful to Ravi Boppana for the detailed

                        T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                               19
                                 C HIN H O L EE AND E MANUELE V IOLA

feedback on a preliminary version of this paper, in particular pointing out several inaccuracies in our
definitions and in our empirical results.
    We wish to thank the referees for the detailed and useful feedback, and in particular for pointing out
several inaccuracies.


References
 [1] N OGA A LON , J EHOSHUA B RUCK , J OSEPH NAOR , M ONI NAOR , AND RON M. ROTH:
     Construction of asymptotically good low-rate error-correcting codes through pseudo-random
     graphs. IRE Trans. Inform. Theory, 38(2):509–516, 1992. Preliminary version in ISIT’91.
     [doi:10.1109/18.119713] 2

 [2] N OGA A LON , O DED G OLDREICH , J OHAN H ÅSTAD , AND R ENÉ P ERALTA: Simple construction
     of almost k-wise independent random variables. Random Structures & Algorithms, 3(3):289–304,
     1992. Preliminary version in FOCS’90. [doi:10.1002/rsa.3240030308] 2, 12

 [3] B ENNY A PPLEBAUM , A NDREJ B OGDANOV, AND A LON ROSEN: A dichotomy for local
     small-bias generators. J. Cryptology, 29(3):577–596, 2016. Preliminary version in TCC’12.
     [doi:10.1007/s00145-015-9202-8] 6

 [4] L OUAY M. J. BAZZI: Polylogarithmic independence can fool DNF formulas. SIAM J. Comput.,
     38(6):2220–2272, 2009. Preliminary version in FOCS’07. [doi:10.1137/070691954] 4, 7

 [5] AVRAHAM B EN -A ROYA AND A MNON TA -S HMA: Constructing small-bias sets from algebraic-
     geometric codes. Theory of Computing, 9(3):253–272, 2013. Preliminary version in FOCS’09.
     [doi:10.4086/toc.2013.v009a005] 2

 [6] S TUART J. B ERKOWITZ: On computing the determinant in small parallel time using a small number
     of processors. Inform. Process. Lett., 18(3):147–150, 1984. [doi:10.1016/0020-0190(84)90018-8] 9

 [7] A NDREJ B OGDANOV, Z EEV DVIR , E LAD V ERBIN , AND A MIR Y EHUDAYOFF: Pseudo-
     randomness for width-2 branching programs. Theory of Computing, 9(7):283–293, 2013.
     [doi:10.4086/toc.2013.v009a007] 2, 3, 6

 [8] A NDREJ B OGDANOV AND E MANUELE V IOLA: Pseudorandom bits for polynomials. SIAM
     J. Comput., 39(6):2464–2486, 2010. Preliminary versions in FOCS’07 and ECCC.
     [doi:10.1137/070712109] 2

 [9] R AVI B. B OPPANA , J OHAN H ÅSTAD , C HIN H O L EE , AND E MANUELE V IOLA: Bounded
     independence vs. moduli. In Proc. 20th Internat. Workshop on Randomization and Com-
     putation (RANDOM’16), LIPIcs, pp. 24:1–24:9. Leibniz-Zentrum fuer Informatik, 2016.
     [doi:10.4230/LIPIcs.APPROX-RANDOM.2016.24] 5

[10] M ARK B RAVERMAN: Poly-logarithmic independence fools AC0 circuits. J. ACM, 57(5):28:1–28:10,
     2010. Preliminary version in CCC’09. [doi:10.1145/1754399.1754401] 4, 7

                      T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                            20
                   S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

[11] J OSHUA B RODY AND E LAD V ERBIN: The coin problem and pseudorandomness for branching pro-
     grams. In Proc. 51st FOCS, pp. 30–39. IEEE Comp. Soc. Press, 2010. [doi:10.1109/FOCS.2010.10]
     18
[12] S URESH C HARI , PANKAJ ROHATGI , AND A RAVIND S RINIVASAN: Improved algorithms via
     approximations of probability distributions. J. Comput. System Sci., 61(1):81–107, 2000. Preliminary
     version in STOC’94. [doi:10.1006/jcss.1999.1695] 3, 19
[13] T HOMAS M. C OVER AND J OY A. T HOMAS: Elements of Information Theory. Wiley-Interscience,
     2006. 8
[14] A NINDYA D E , O MID E TESAMI , L UCA T REVISAN , AND M ADHUR T ULSIANI: Improved pseudo-
     random generators for depth 2 circuits. In Proc. 14th Internat. Workshop on Randomization and Com-
     putation (RANDOM’10), volume 6302 of LNCS, pp. 504–517. Springer, 2010. [doi:10.1007/978-3-
     642-15369-3_38] 3, 7
[15] G UY E VEN , O DED G OLDREICH , M ICHAEL L UBY, N OAM N ISAN , AND B OBAN V ELICKOVIC:
     Efficient approximation of product distributions. Random Structures & Algorithms, 13(1):1–16,
     1998. [doi:10.1002/(SICI)1098-2418(199808)13:1<1::AID-RSA1>3.0.CO;2-W] 3
[16] DANIEL G ORENSTEIN AND N EAL Z IERLER: A class of error-correcting codes in pm symbols. J.
     Soc. Indust. Appl. Math., 9(2):207–214, 1961. [doi:10.1137/0109020] 6, 9
[17] V ENKATESAN G URUSWAMI AND ATRI RUDRA: Tolerant locally testable codes. In Proc. 9th
     Internat. Workshop on Randomization and Computation (RANDOM’05), volume 3624 of LNCS, pp.
     306–317. Springer, 2005. [doi:10.1007/11538462_26] 6
[18] RUSSELL I MPAGLIAZZO: Hard-core distributions for somewhat hard problems. In Proc. 36th
     FOCS, pp. 538–545. IEEE Comp. Soc. Press, 1995. [doi:10.1109/SFCS.1995.492584] 7, 13, 14
[19] S WASTIK KOPPARTY AND S HUBHANGI S ARAF: Tolerant linearity testing and locally testable
     codes. In Proc. 13th Internat. Workshop on Randomization and Computation (RANDOM’09),
     volume 5687 of LNCS, pp. 601–614. Springer, 2009. [doi:10.1007/978-3-642-03685-9_45] 6
[20] C HIN H O L EE AND E MANUELE V IOLA: Some limitations of the sum of small-bias distributions.
     Electron. Colloq. on Comput. Complexity (ECCC). TR15-005, 2016. 1, 5
[21] S HACHAR L OVETT: Unconditional pseudorandom generators for low degree polynomials. Theory of
     Computing, 5(3):69–82, 2009. Preliminary version in STOC’08. [doi:10.4086/toc.2009.v005a003]
     2
[22] R AGHU M EKA AND DAVID Z UCKERMAN: Small-bias spaces for group products. In Proc. 13th
     Internat. Workshop on Randomization and Computation (RANDOM’09), volume 5687 of LNCS, pp.
     658–672. Springer, 2009. [doi:10.1007/978-3-642-03685-9_49] 2, 3, 4, 7, 11
[23] E LCHANAN M OSSEL , A MIR S HPILKA , AND L UCA T REVISAN: On epsilon-biased generators in
     NC0 . Random Structures & Algorithms, 29(1), 2005. Preliminary versions in FOCS’03 and ECCC.
     [doi:10.1109/SFCS.2003.1238188] 6

                      T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                           21
                                C HIN H O L EE AND E MANUELE V IOLA

[24] J OSEPH NAOR AND M ONI NAOR: Small-bias probability spaces: efficient constructions and
     applications. SIAM J. Comput., 22(4):838–856, 1993. Preliminary version in STOC’90.
     [doi:10.1137/0222053] 2, 6, 12

[25] N OAM N ISAN: Pseudorandom bits for constant depth circuits. Combinatorica, 11(1):63–70, 1991.
     Preliminary version in FOCS’88. [doi:10.1007/BF01375474] 7, 11

[26] R AYMOND E. A. C. PALEY: On orthogonal matrices. J. Math. Phys., 12(1-4):311–320, 1933.
     [doi:10.1002/sapm1933121311] 18

[27] W ILLIAM W ESLEY P ETERSON: Encoding and error-correction procedures for the Bose-Chaudhuri
     codes. IRE Trans. Inform. Theory, 6(4):459–470, 1960. [doi:10.1109/TIT.1960.1057586] 6, 9

[28] A LEXANDER A. R AZBOROV: Lower bounds on the dimension of schemes of bounded depth in a
     complete basis containing the logical addition function. Mat. Zametki Akad. Nauk SSSR, 41(4):598–
     607, 1987. English translation in Math. Notes of the Acad. Sci. of the USSR, 41(4):333-338, 1987.
     [doi:10.1007/BF01137685] 2

[29] A LEXANDER A. R AZBOROV: A simple proof of Bazzi’s theorem. ACM Trans. Comput. Theory,
     1(1):3:1–3:5, 2009. [doi:10.1145/1490270.1490273] 4

[30] ATRI RUDRA AND S TEVE U URTAMO: Data stream algorithms for codeword testing. In Proc. 37th
     Internat. Colloq. on Automata, Languages and Programming (ICALP’10), pp. 629–640. Springer,
     2010. [doi:10.1007/978-3-642-14165-2_53, arXiv:1004.4601] 6, 7

[31] RONEN S HALTIEL AND E MANUELE V IOLA: Hardness amplification proofs require majority. SIAM
     J. Comput., 39(7):3122–3154, 2010. Preliminary version in STOC’08. [doi:10.1137/080735096] 7,
     18

[32] A MIR S HPILKA: Constructions of low-degree and error-correcting ε-biased generators. Comput.
     Complexity, 18(4):495–525, 2009. Preliminary version found in CCC’06. [doi:10.1007/s00037-
     009-0281-5] 6, 8

[33] ROMAN S MOLENSKY: Algebraic methods in the theory of lower bounds for boolean circuit
     complexity. In Proc. 19th STOC, pp. 77–82. ACM Press, 1987. [doi:10.1145/28395.28404] 7, 13

[34] AVISHAY TAL: Tight bounds on the Fourier spectrum of AC0 . In Proc. 32nd IEEE Conf. on
     Computational Complexity (CCC’17), volume 79 of LIPIcs, pp. 15:1–15:31. Leibniz-Zentrum fuer
     Informatik, 2017. Preliminary version in ECCC’14. [doi:10.4230/LIPIcs.CCC.2017.15] 4, 7

[35] L ESLIE G. VALIANT: Graph-theoretic arguments in low-level complexity. In 6thSymp. Math.
     Found. Comp. Sci. (MFCS’77), volume 53 of LNCS, pp. 162–176. Springer, 1977. [doi:10.1007/3-
     540-08353-7_135] 3, 7, 10

[36] L ESLIE G. VALIANT: Exponential lower bounds for restricted monotone circuits. In Proc. 15th
     STOC, pp. 110–117. ACM Press, 1983. [doi:10.1145/800061.808739] 10

                     T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                         22
                  S OME L IMITATIONS OF THE S UM OF S MALL -B IAS D ISTRIBUTIONS

[37] E MANUELE V IOLA: The Complexity of Hardness Amplification and Derandomization. Ph. D.
     thesis, Harvard University, 2006. 7, 18
[38] E MANUELE V IOLA: On the power of small-depth computation. Found. Trends Theor. Comput. Sci.,
     5(1):1–72, 2009. [doi:10.1561/0400000033] 2, 7, 10, 13
[39] E MANUELE V IOLA: The sum of d small-bias generators fools polynomials of degree d. Comput.
     Complexity, 18(2):209–217, 2009. Preliminary version in CCC’08. [doi:10.1007/s00037-009-0273-
     5] 2, 19
[40] E MANUELE V IOLA AND AVI W IGDERSON: Norms, XOR lemmas, and lower bounds for polyno-
     mials and protocols. Theory of Computing, 4(7):137–168, 2008. Preliminary version in CCC’07.
     [doi:10.4086/toc.2008.v004a007] 8
[41] R ICHARD M. W ILSON: Combinatorial analysis lecture notes. 2012. 16, 17


AUTHORS
     Chin Ho Lee
     Ph. D. student
     College of Computer and Information Science
     Northeastern University
     Boston, MA
     chlee ccs neu edu
     http://www.ccs.neu.edu/~chlee

     Emanuele Viola
     Associate professor
     College of Computer and Information Science
     Northeastern University
     Boston, MA
     viola ccs neu edu
     http://www.ccs.neu.edu/~viola


ABOUT THE AUTHORS
     C HIN H O L EE is a Ph. D. student at the College of Computer and Information Science of
        Northeastern University, under the supervision of Emanuele Viola. He obtained his
        M. Phil. in Computer Science and Engineering (advised by Andrej Bogdanov) and B. S.
        in Mathematics from the Chinese University of Hong Kong.

     E MANUELE V IOLA moved four times in the academic year 2015–2016. He also has a blog!
        Check it out at https://emanueleviola.wordpress.com/.


                    T HEORY OF C OMPUTING, Volume 13 (16), 2017, pp. 1–23                       23