New Error Correcting Codes from Lifting Alan Xinyu Guo

New Error Correcting Codes from Lifting
by
Alan Xinyu Guo
B.S., Duke University (2011)
S.M., Massachusetts Institute of Technology (2013)
Submitted to the Department of Electrical Engineering and Computer
Science
in partial fulfillment of the requirements for the degree of
Doctor of Philosophy in Electrical Engineering and Computer Science
at the
MASSACHUSETTS INSTITUTE OF TECHNOLOGY
June 2015
c Massachusetts Institute of Technology 2015. All rights reserved.
β—‹
Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Department of Electrical Engineering and Computer Science
May 1, 2015
Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Madhu Sudan
Adjunct Professor
Thesis Supervisor
Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Leslie A. Kolodziejski
Chair, Department Committee on Graduate Students
2
New Error Correcting Codes from Lifting
by
Alan Xinyu Guo
Submitted to the Department of Electrical Engineering and Computer Science
on May 1, 2015, in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy in Electrical Engineering and Computer Science
Abstract
Error correcting codes have been widely used for protecting information from noise. The
theory of error correcting codes studies the range of parameters achievable by such codes, as
well as the efficiency with which one can encode and decode them. In recent years, attention
has focused on the study of sublinear-time algorithms for various classical problems, such
as decoding and membership verification. This attention was driven in part by theoretical
developments in probabilistically checkable proofs (PCPs) and hardness of approximation.
Locally testable codes (codes for which membership can be verified using a sublinear number
of queries) form the combinatorial core of PCP constructions and thus play a central role
in computational complexity theory. Historically, low-degree polynomials (the Reed-Muller
code) have been the locally testable code of choice. Recently, “affine-invariant” codes have
come under focus as providing potential for new and improved codes.
In this thesis, we exploit a natural algebraic operation known as “lifting” to construct new
affine-invariant codes from shorter base codes. These lifted codes generically possess desirable
combinatorial and algorithmic properties. The lifting operation preserves the distance of the
base code. Moreover, lifted codes are naturally locally decodable and testable. We tap
deeper into the potential of lifted codes by constructing the “lifted Reed-Solomon code”,
a supercode of the Reed-Muller code with the same error-correcting capabilities yet vastly
greater rate.
The lifted Reed-Solomon code is the first high-rate code known to be locally decodable up
to half the minimum distance, locally list-decodable up to the Johnson bound, and robustly
testable, with robustness that depends only on the distance of the code. In particular, it is
the first high-rate code known to be both locally decodable and locally testable. We also
apply the lifted Reed-Solomon code to obtain new bounds on the size of Nikodym sets, and
also to show that the Reed-Muller code is robustly testable for all field sizes and degrees up
to the field size, with robustness that depends only on the distance of the code.
Thesis Supervisor: Madhu Sudan
Title: Adjunct Professor
3
4
Acknowledgments
First and foremost, I thank my advisor, Madhu Sudan, for his patient guidance, support, and
encouragement throughout my four years at MIT. Our shared taste in finding clean, general
solutions to algebraic problems has led to many collaborations. I thank Scott Aaronson and
Dana Moshkovitz for serving on my thesis committee.
I am also indebted to Ezra Miller, who guided my undergraduate research at Duke, and
to Vic Reiner and Dennis Stanton who mentored me at their REU in Minnesota.
I thank my co-authors during graduate school: Greg Aloupis, Andrea Campagna, Erik
Demaine, Elad Haramaty, Swastik Kopparty, Ronitt Rubinfeld, Madhu Sudan, and Giovanni
Viglietta. I especially thank Ronitt and Piotr Indyk for their guidance early on, Erik for his
excitement and willingness to entertain my more fun research ideas (i.e. hardness of video
games), and Swastik for hosting my visit to Rutgers and teaching me about codes and PCPs.
I also thank friends and faculty with whom I have shared conversations at MIT: Pablo
Azar, Mohammad Bavarian, Eric Blais, Adam Bouland, Mahdi Cheraghchi, Henry Cohn,
Matt Coudron, Michael Forbes, Badih Ghazi, Pritish Kamath, Ameya Velingker, Henry
Yuen, and so many others.
I thank my friends outside of the field for their friendship and the good times during the
past four years. I especially thank Vivek Bhattacharya, Sarah Freitas, Henry Hwang, Steven
Lin, Ann Liu, Lakshya Madhok, and Roger Que.
I am grateful to my family. Without my parents, I would not exist, and neither would
this thesis. Moreover, they always supported my education and encouraged me to pursue
my dreams. I thank my younger sister Julia for encouraging me to set a good example.
Finally, my biggest thanks goes to Lisa — my fiancée, best friend, and companion for the
rest of my life. Graduate school was not always easy, but you were always there to support
me and pick me up when I was down, and to share in my triumphs. Thank you for your
unwavering love and loyalty, without which I do not believe I would have survived through
graduate school. Only you know every twist and turn my journey has taken. Without
hesitation, I dedicate this thesis to you.
5
6
Contents
1 Introduction
1.1
1.2
11
Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11
1.1.1
Error Correcting Codes . . . . . . . . . . . . . . . . . . . . . . . . . .
11
1.1.2
PCPs and Local Algorithms . . . . . . . . . . . . . . . . . . . . . . .
13
1.1.3
Affine-Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
16
This Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
17
1.2.1
Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
17
1.2.2
Lifting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18
1.2.3
Robust Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18
1.2.4
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
23
1.2.5
Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
25
2 Preliminaries
27
2.1
Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
27
2.2
Probability and Concentration bounds . . . . . . . . . . . . . . . . . . . . .
29
3 Error Correcting Codes
31
3.1
Classical Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
31
3.2
Local decoding, correcting, and list-decoding . . . . . . . . . . . . . . . . . .
32
3.3
Local testing and robust testing . . . . . . . . . . . . . . . . . . . . . . . . .
34
3.4
Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
35
3.4.1
36
Reed-Solomon code . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7
3.4.2
Reed-Muller code . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4 Affine-invariance
36
39
4.1
Affine-invariant Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
39
4.2
Equivalence of Invariance under Affine Transformations and Permutations . .
40
5 Lifting Codes
47
5.1
The Lift Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
47
5.2
Algebraic and Combinatorial Properties
. . . . . . . . . . . . . . . . . . . .
48
5.2.1
Algebraic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . .
48
5.2.2
Distance of Lifted Codes . . . . . . . . . . . . . . . . . . . . . . . . .
50
Local Decoding and Correcting . . . . . . . . . . . . . . . . . . . . . . . . .
53
5.3.1
Local Correcting up to 1/4 Distance . . . . . . . . . . . . . . . . . .
53
5.3.2
Local Correcting up to 1/2 Distance . . . . . . . . . . . . . . . . . .
54
5.3.3
Local Decoding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
57
Local Testing and Robust Testing . . . . . . . . . . . . . . . . . . . . . . . .
59
5.4.1
Local Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
59
5.4.2
Robust Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
60
5.3
5.4
6 Robust Testing of Lifted Codes
6.1
6.2
61
Robustness of Lifted Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . .
61
6.1.1
Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
61
6.1.2
Robustness for Small Dimension . . . . . . . . . . . . . . . . . . . . .
62
6.1.3
Robustness of Special Tensor Codes . . . . . . . . . . . . . . . . . . .
66
6.1.4
Robustness for Large Dimension . . . . . . . . . . . . . . . . . . . . .
74
Technical Algebraic Results . . . . . . . . . . . . . . . . . . . . . . . . . . .
81
6.2.1
Degree Lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
81
6.2.2
Analysis of Subspace Restrictions . . . . . . . . . . . . . . . . . . . .
85
8
7 Applications
7.1
91
Lifted Reed-Solomon Code . . . . . . . . . . . . . . . . . . . . . . . . . . . .
91
7.1.1
Relationship to Reed-Muller . . . . . . . . . . . . . . . . . . . . . . .
92
7.1.2
Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
95
7.1.3
Global List-Decoding . . . . . . . . . . . . . . . . . . . . . . . . . . .
97
7.1.4
Local List-Decoding . . . . . . . . . . . . . . . . . . . . . . . . . . .
98
7.1.5
Main Result: The Code That Does It All . . . . . . . . . . . . . . . . 102
7.2
Robust Low-Degree Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.3
Nikodym Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
A Algebra Background
107
A.1 Arithmetic over finite fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
A.2 Tensor codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
B Finite field geometry
111
B.1 Affine maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
B.2 Affine subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
9
10
Chapter 1
Introduction
1.1
1.1.1
Background
Error Correcting Codes
Error correcting codes arise as a solution to the problem of communicating over a noisy
channel. The sender first encodes the message using an error correcting code, which adds
redundancy to the message, into a codeword. The codeword is then sent over the noisy
channel. The receiver receives a word which is a corruption of the codeword. The receiver
then decodes the received word and hopefully retrieves the original message. The actual code
is the set of possible codewords.
Two classical parameters of interest are the rate and distance of the code. The rate is
the ratio of the message length to the codeword length, and measures the efficiency of the
encoding. The distance measures the error-correcting capability of the code. The Hamming
distance between two strings is the number of symbols in which they differ. The distance of
a code is the minimum Hamming distance between two distinct codewords. If the distance
of a code is 𝑑, then in principle one can detect up to 𝑑 − 1 errors and correct up to ⌊𝑑/2⌋ − 1
errors: if the codeword has been corrupted in at most 𝑑−1 locations, then it cannot have been
corrupted into a different codeword, so to detect errors one “merely” checks if the received
word is a codeword; if the codeword has been corrupted in at most ⌊𝑑/2⌋ − 1 locations, then
11
there is at most one codeword within Hamming distance ⌊𝑑/2⌋ − 1 of the received word, so
to correct errors one “merely” finds the nearest codeword to the received word. We often
prefer to work with the relative distance of a code, which is simply its distance divided by
the length of the codewords. There is a fundamental tradeoff between rate and distance,
which is still not fully understood. In addition, there is the problem of designing codes which
support efficient algorithms for encoding and decoding.
Another notion of decoding is list-decoding. We just showed that if a code π’ž has distance
𝑑, then for any word, there is at most one codeword within Hamming distance ⌊𝑑/2⌋ − 1. If
we wish to correct more than ⌊𝑑/2⌋ − 1 errors, we cannot guarantee that there is a unique
codeword. However, if the radius is not too large, then we can hope that there are not too
many codewords within the radius from the received word. Instead of outputting the correct
codeword or message, a list-decoding would output a list of potential codewords or messages.
Another important feature of an error correcting code is the alphabet. Designing a code
is easier using a larger alphabet. If the alphabet size is allowed to grow with the code
length 𝑛, then the Singleton bound asserts that the rate 𝑅 and relative distance 𝛿 satisfies
𝑅 + 𝛿 ≤ 1 + 1/𝑛. This bound is tight, as demonstrated by the Reed-Solomon code.
The Reed-Solomon code is perhaps the most ubiquitous code in the literature. The idea
is simple. Let π‘˜ ≥ 1 and let π‘ž ≥ π‘˜ be a prime power. Let Fπ‘ž be the finite field of size π‘ž. Each
message π‘š = (π‘š0 , . . . , π‘šπ‘˜−1 ) of length π‘˜ over the alphabet Fπ‘ž is interpreted as a degree π‘˜ − 1
polynomial π‘š(𝑋) = π‘š0 + π‘š1 𝑋 + · · · + π‘šπ‘˜−1 𝑋 π‘˜−1 . Let 𝛼1 , . . . , π›Όπ‘ž be the elements of Fπ‘ž .
The encoding of π‘š is simply the evaluation of π‘š at every point: (π‘š(𝛼1 ), . . . , π‘š(π›Όπ‘ž )). The
rate of this code is clearly 𝑅 = π‘˜π‘ž , and it follows from the Fundamental Theorem of Algebra
that 𝛿 = 1 −
π‘˜−1
,
π‘ž
so that 𝑅 + 𝛿 = 1 + 1/π‘ž, meeting the Singleton bound. Furthermore,
the Reed-Solomon code can be efficiently decoded up to half its distance using, for instance,
the Welch-Berlekamp algorithm [WB86] (see [GS92] for an exposition). Guruswami and
Sudan [GS99] showed that the Reed-Solomon code with distance 𝛿 > 0 can be list-decoded
√
up to the “Johnson bound” (1 − 1 − 𝛿 fraction errors). More precisely, they gave an
efficient algorithm which, on input a received word, outputs a list of 𝑂(1/πœ–2 ) codewords that
12
√
are within (relative) distance 1 − (1 + πœ–) 1 − 𝛿 of the received word.
A related code is the Reed-Muller code. This code, parameterized by a degree 𝑑 and a
number π‘š of variables, consists of polynomials 𝑓 : Fπ‘š
π‘ž → Fπ‘ž of degree at most 𝑑 (or rather,
(︀𝑑+π‘š)οΈ€
their evaluations). Its message length is π‘˜ = π‘š and its code length is 𝑛 = π‘ž π‘š . When
π‘š = 1, this is simply the Reed-Solomon code. If 𝑑 < π‘ž, it follows from the SchwartzZippel lemma that the distance of the Reed-Muller code is 1 − π‘‘π‘ž . Although the Reed-Muller
code’s rate-distance tradeoff is worse than that of the Reed-Solomon code, the Reed-Muller
code offers locality, which we discuss in Section 1.1.2. Like the Reed-Solomon code, the
Reed-Muller code can be list-decoded up to the Johnson bound [PW04].
1.1.2
PCPs and Local Algorithms
In the late 1980s and early 1990s, there was an explosion of interest in sublinear-time algorithms. Blum, Luby, and Rubinfeld [BLR93] showed that one can probabilistically test
whether a given function 𝑓 : {0, 1}𝑛 → {0, 1} is linear — that is, 𝑓 (x + y) = 𝑓 (x) + 𝑓 (y) for
every x, y ∈ {0, 1}𝑛 — by querying 𝑓 at only 3 points. If 𝑓 is linear, then their test accepts
with probability one, while if 𝑓 is “πœ–-far” from linear (it disagrees with every linear function
in at least πœ–-fraction of the domain), then their test rejects with probability Ω(πœ–). The space
of linear functions 𝑓 : {0, 1}𝑛 → {0, 1} forms an error correcting code with rate 𝑛/2𝑛 and
distance 1/2 — this code is known as the Hadamard code. This was in fact the first code
shown to be locally testable — that is, using a sublinear number of queries to the received
word, one can verify membership in the code. It is also easy to show that the Hadamard
code is locally correctable — one can correct any given symbol of the received word with high
probability using a sublinear number of queries. To correct 𝑓 at x ∈ {0, 1}𝑛 , select random
y ∈ {0, 1}𝑛 and output the value 𝑓 (x + y) − 𝑓 (y). A related notion if local decodability —
one can correct any given symbol of the message with high probability by making a sublinear
number of queries to the received word.
We will often consider codes π’ž ⊆ Fπ‘›π‘ž that are linear (i.e. π’ž forms a vector space over
Fπ‘ž ). The Reed-Solomon code, Reed-Muller code, and Hadamard code are all linear codes.
13
Rather than thinking of words in Fπ‘›π‘ž as sequences of length 𝑛, we view them as functions
from some fixed set 𝑆 of cardinality |𝑆| = 𝑛 to the range Fπ‘ž . The structure of the set 𝑆 and
symmetries will play a role later. We use {𝑆 → Fπ‘ž } to denote the set of all such functions.
We say a function 𝑓 is 𝜏 -far from π’ž if 𝛿(𝑓, π’ž) , min𝑔∈π’ž 𝛿(𝑓, 𝑔) ≥ 𝜏 .
Given a code π’ž ⊆ {𝑆 → Fπ‘ž } and integer β„“, an β„“-local tester 𝒯 is a distribution π’Ÿ on
β„“
(𝑆 β„“ , 2Fπ‘ž ) with the semantics as follows: given oracle access to 𝑓 : 𝑆 → Fπ‘ž , the tester 𝒯
samples (πœ‹, 𝑉 ) ← π’Ÿ, where πœ‹ = (πœ‹1 , . . . , πœ‹β„“ ) ∈ 𝑆 β„“ and 𝑉 ⊆ Fβ„“π‘ž , and accepts 𝑓 if and only if
𝑓 |πœ‹ , (𝑓 (πœ‹1 ), . . . , 𝑓 (πœ‹β„“ )) ∈ 𝑉 . The tester is πœ–-sound if 𝒯 accepts 𝑓 ∈ π’ž with probability one,
while rejecting 𝑓 that is 𝛿-far from π’ž with probability at least πœ– · 𝛿.
We will also be interested in a stronger property of testers known as their robustness,
formally defined by Ben-Sasson and Sudan [BSS06], based on analogous notions in complexity
theory due to Ben-Sasson et al. [BSGH+ 04] and Dinur and Reingold [DR04]. The hope with
a robust tester is that, while it may make a few more queries than the minimum possible,
the rejection is “more emphatic” in that functions that are far from π’ž typically yield views
that are far from acceptable, i.e. if 𝛿(𝑓, π’ž) is large, then so is 𝛿(𝑓 |πœ‹ , 𝑉 ) for typicaly choices of
(πœ‹, 𝑉 ) ← π’Ÿ. Formally, a tester π’Ÿ is 𝛼-robust if E(πœ‹,𝑉 )←π’Ÿ [𝛿(𝑓 |πœ‹ , 𝑉 )] ≥ 𝛼 · 𝛿(𝑓, π’ž). Robustness
can be a much stronger property than mere soundnesss since it allows for composition with
other local testers. In particular, if there is an 𝛼-robust tester for 𝑓 with distribution π’Ÿ and
if for every (πœ‹, 𝑉 ) in the support of π’Ÿ, the property of being in 𝑉 has an β„“′ -local tester that
is πœ–-sound, then π’ž has an β„“′ -local tester that is 𝛼 · πœ–-sound. The hope that membership in 𝑉
has a nice local tester for every 𝑉 in the support of π’Ÿ may seem overly optimistic, but for
many symmetric codes (such as affine-invariant codes, to be discussed later), all the 𝑉 ’s are
isomorphic — so this is really just one hope.
The interest in sublinear-time algorithms is obvious from a practical perspective. As the
amount of data stored and transmitted by society continues its explosive growth, even lineartime algorithms may be too slow, and also unnecessary. Indeed, statisticians understood this
decades ago when estimating the population averages. If we want an approximate answer,
often 𝑂(1) queries suffice to give a good approximation with high confidence. Moreover, in
14
the context of error correction, if we have a very large encoded file and only wish to decode
a small portion of it, then we need our code to be locally decodable.
Surprisingly, sublinear-time algorithms for algebraic codes have played a prominent role
in computational complexity theory. In particular, locally testable codes form the “combinatorial core” of probabilistically checkable proofs (PCPs). A PCP for a language 𝐿 is a
protocol involving two parties, a Prover and a Verifier, and an input π‘₯, whereby the Prover
supplies a proof, depending on π‘₯, in an attempt to convince the Verifier that π‘₯ ∈ 𝐿. The
Verifier makes a small number of random queries to the proof and then either accepts or
rejects the proof based on what it sees. A valid PCP for 𝐿 satisfies the following: if π‘₯ ∈ 𝐿,
then there is some proof that the Prover can provide such that 𝑉 accepts with probability
1; if π‘₯ ∈
/ 𝐿, then regardless of the proof provided by the Prover, the Verifier will reject
with high probability (over its random queries). The celebrated PCP Theorem, proved
in [AS98, ALM+ 98], characterizes the complexity class NP as the class of languages with
PCPs where the Verifier makes only 𝑂(1) queries to the proof and uses only 𝑂(log 𝑛) bits of
randomness in determining its random queries, where 𝑛 is the length of the input π‘₯. Not only
did this theorem elucidate the class of NP a bit more, but it also paved the way for proving that, for many combinatorial optimization problems, even approximating the solution is
NP-hard.
At the heart of the proof of the PCP theorem lies the problem of low-degree testing —
the problem of testing whether a given function 𝑓 : Fπ‘š
π‘ž → Fπ‘ž has total degree deg(𝑓 ) ≤ 𝑑
or is far from any such function. Low-degree testing has been the most extensively studied
algebraic property testing problem. First studied in the work of Rubinfeld and Sudan [RS96],
low-degree testing and many variations have been analyzed in many subsequent works — a
partial list includes [ALM+ 98, FS95, AS03, RS97, MR06, AKK+ 05, KR06, JPRZ09, BKS+ 10,
HSS11]. When 𝑑 β‰ͺ π‘ž, low-degree tests making as few as 𝑑 + 2 queries are known, that have
1/ poly(𝑑)-soundness (see, for instance, Friedl-Sudan [FS95]). However, tests that make 𝑂(𝑑)
queries achieve constant soundness (a universal constant independent of π‘š, 𝑑, π‘ž provided π‘ž is
sufficiently larger than 𝑑), and even constant robustness. This constant robustness is central
15
to the PCP construction of Arora et al. [ALM+ 98]. In all cases with 𝑑 β‰ͺ π‘ž, low-degree
tests operate by considering the restriction of a function to a random line, or sometimes
plane, in the domain, and accepting a function if its restriction to the chosen subspace is
a polynomial of degree at most 𝑑. Thus, the different restrictions πœ‹ are different affine
subspaces of low dimension (one or two) and the acceptable pattern 𝑉 is the same for all
πœ‹. In particular, the robust analysis of the low-degree test allows for low-query tests, or
even proofs, of membership in 𝑉 in constant dimensional spaces to be composed with the
low-degree test in high dimensions to yield low-query PCPs. Robustness turns out to be
much more significant as a parameter to analyze in these results than the query complexity
of the outer test. Indeed, subsequent strengthenings of the PCP theorem in various senses
(e.g. in [AS03, RS97, MR06]) rely on improving the robustness to a quantity close to 1, and
this leads to PCPs of arbitrarily small constant, and then even π‘œ(1), error.
1.1.3
Affine-Invariance
In an attempt to understand what exactly makes codes like the Reed-Muller code testable,
Kaufman and Sudan set out to systematically study affine-invariant codes [KS08]. By then,
it was known that symmetry in properties contributed to their testability, as in graph property testing. The hope was to find the analogous notion of symmetry for algebraic properties
such as linearity or low-degree, and affine-invariance seemed to be a promising abstraction. A
code π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } is affine-invariant if, for every codeword 𝑓 ∈ π’ž, the codeword 𝑓 ∘ 𝐴 obπ‘š
tained by first applying an affine permutation 𝐴 : Fπ‘š
π‘ž → Fπ‘ž to the domain before evaluating
𝑓 , is also a codeword. Kaufman and Sudan showed in [KS08] that any linear affine-invariant
code that is “π‘˜-single-orbit characterized” is testable with π‘˜ queries and soundness Ω(π‘˜ −2 ).
For example, the Hadamard code is 3-single-orbit characterized by 𝑓 (x)+𝑓 (y)−𝑓 (x+y) = 0
for any x, y ∈ Fπ‘š
π‘ž (the characterization is by constraints on 3 points, and the constraints
are all in a single orbit by the group action of affine permutations). Their proof simplifies,
unifies, and generalizes the proofs found in [BLR93, RS96, AKK+ 05, KR06, JPRZ09] and
unearths some of the fundamental underlying reasons for why testing works. Additionally,
16
Kaufman and Sudan initiated the systematic study of affine-invariant properties, and laid
the groundwork for our structural understanding of affine-invariant properties. The hope
was that eventually the study of affine-invariance would lead to constructions of new affineinvariant codes with desirable testability properties.
While the locality properties (testability and correctability) of Reed-Muller codes are
well-studied, they are essentially the only rich class of symmetric codes that are well-studied.
The only other basic class of symmetric codes that are studied seem to be sparse ones (codes
with few codewords).
1.2
This Thesis
In this thesis, we use an algebraic operation, known as “lifting”, to construct new affineinvariant codes. This thesis includes results from [GKS13, GK14, GHS15], though we omit
some results and generalize other results from these papers. [GKS13] is joint work with
Swastik Kopparty and Madhu Sudan. [GK14] is joint work with Swastik Kopparty. [GHS15]
is joint work with Elad Haramaty and Madhu Sudan.
1.2.1
Main Result
The main result of our thesis is the construction of high-rate LCCs and LTCs.
Theorem 1.2.1 (Main theorem, informal). ∀πœ–, 𝛽 > 0 ∃𝛿, 𝛼 > 0 such that for infinitely many
𝑛 there exists π‘ž = π‘ž(𝑛) = 𝑂(π‘›πœ– ) and a linear code π’ž ⊆ Fπ‘›π‘ž of distance 𝛿 and rate 1 − 𝛽 such
that
βˆ™ π’ž is locally correctable and decodable from Ω(𝛿) fraction errors with 𝑂(π‘›πœ– ) queries;
√
βˆ™ π’ž is list-decodable from 1 − 1 − 𝛿 fraction errors in polynomial time, and is locally
√
list-decodable from 1 − 1 − 𝛿 fraction errors with 𝑂(𝑛3πœ– ) queries;
βˆ™ π’ž has an 𝛼-robust tester using 𝑂(𝑛2πœ– ) queries.
Theorem 1.2.1 is proved in Section 7.1.5, and is the culmination of the work in this thesis.
17
1.2.2
Lifting
The lifting operation was first defined and used in [BSMSS11] to prove negative results
— in particular, to construct “symmetric LDPC codes” that are not testable. The work
of [GKS13] initiated the systematic study of lifting, and also was the first work to use lifting
to prove positive results — in particular, to construct new high-rate locally correctable and
locally testable codes. Our definition of lifting is more general and somewhat cleaner than
that of [BSMSS11]. Starting from a base linear affine-invariant code π’ž ⊆ {Fπ‘žπ‘‘ → Fπ‘ž }, we
π‘š
define the π‘š-dimensional lift π’ž π‘‘β†—π‘š ⊆ {Fπ‘š
π‘ž → Fπ‘ž } to be the code consisting of all 𝑓 : Fπ‘ž → Fπ‘ž
satisfying the following: 𝑓 is in the lift if and only if, for every 𝑑-dimensional affine subspace
𝐴 ⊆ Fπ‘š
π‘ž , the restriction 𝑓 |𝐴 ∈ π’ž.
Lifting is a natural algebraic operation on affine-invariant codes, which is evident in the
generic properties that lifted codes naturally possess. We show that if the base code π’ž has
(relative) distance 𝛿, then the lifted code π’ž π‘‘β†—π‘š has distance 𝛿 − π‘ž −𝑑 , and in fact if π‘ž ∈ {2, 3},
then the lifted distance is also 𝛿. So, lifting approximately preserves distance. Moreover,
π’ž π‘‘β†—π‘š is naturally π‘ž 𝑑 -single-orbit characterized by construction, and so, by [KS08], the natural
𝑑-dimensional test — choose a random 𝑑-dimensional affine subspace 𝐴 ⊆ Fπ‘š
π‘ž and accept if
and only if 𝑓 |𝐴 ∈ π’ž — is Ω(π‘ž −2𝑑 )-sound. Finally, lifted codes are naturally locally correctable
π‘š
— to correct 𝑓 at a point x ∈ Fπ‘š
π‘ž , choose a random 𝑑-dimensional subspace 𝐴 ⊆ Fπ‘ž passing
through x, use the correction algorithm for π’ž to correct 𝑓 |𝐴 to a codeword 𝑐 ∈ π’ž, and then
output 𝑐(x).
1.2.3
Robust Testing of Lifted Codes
In [GHS15], we consider robust testing of lifted codes. We propose and analyze the following
𝑑↗2𝑑
test for π’ž π‘‘β†—π‘š : Pick a random 2𝑑-dimensional subspace 𝐴 in Fπ‘š
.
π‘ž and accept if 𝑓 |𝐴 ∈ π’ž
Our main theorem relates the robustness of this test to the distance of the code π’ž.
Theorem 1.2.2. ∀𝛿 > 0 ∃𝛼 > 0 such that the following holds: For every finite field Fπ‘ž , for
every pair of positive integers 𝑑 and π‘š, and for every affine-invariant code π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž }
satisfying 𝛿(π’ž) ≥ 𝛿, the code π’ž π‘‘β†—π‘š has a π‘ž 2𝑑 -local test that is 𝛼-robust.
18
Theorem 1.2.2 is proved in Section 6.1. As we elaborate below, Theorem 1.2.2 immediately implies a robust analysis for low-degree tests. Whereas almost all previous robust
analyses of low-degree tests had more complex conditions on the relationship between the
robustness, the degree, and the field size, our relationship is extremely clean. The dependence of 𝛼 on 𝛿 that we prove is polynomial but of fairly high degree 𝛼 = Ω(𝛿 74 ). We do not
attempt to improve this relationship in this thesis and choose instead to keep the intermediate statements simple and general. We note that a significant portion of this complexity
arises due to our desire to lift 𝑑-dimensional codes for general 𝑑, and here the fact that the
robustness lower-bound is independent of 𝑑 is itself significant.
Comparing with other testing results for lifted codes, there are only two prior works to
compare with: Kaufman and Sudan [KS08] analyze a tester for a broader family of codes
that they call “single-orbit” codes. Their result would yield a robustness of Θ(π‘ž −3𝑑 ). (See
Corollary 6.1.2.)
Haramaty et al. [HRS13] also give a tester for lifted codes. They do not state their results
in terms of robustness but their techniques would turn into a robustness of πœ–π‘ž · 𝛿, where the
πœ–π‘ž is a positive constant for constant π‘ž but goes to zero extremely quickly as π‘ž → ∞. Thus
for growing π‘ž (and even slowly shrinking 𝛿) our results are much stronger.
Proof approach and some technical contributions
In order to describe our test and analysis techniques, we briefly review the two main tests
proposed in the literature for “low-degree testing”, when the field size is much larger than
the degree. The most natural test for this task is the one that picks a random line in Fπ‘š
π‘ž
and computes the proximity of the function restricted to this line to the space of univariate
degree 𝑑 polynomials. This is the test proposed by Rubinfeld and Sudan [RS96] and analyzed
in [RS96, ALM+ 98, AS03]. A second low-degree test is somewhat less efficient in its query
complexity (quadratically so) but turns out to have a much simpler analysis — this test
would pick a random two-dimensional (affine) subspace in Fπ‘š
π‘ž and verify that the function
is a bivariate polynomial of degree at most 𝑑 on this subspace. This is the test proposed
19
by Raz and Safra [RS97] and analyzed in [RS97, MR06]. Both tests can be analyzed by
first reducing the testing problem to that of testing constant variate functions (at most four
variate functions) and then analyzing the constant dimensional problem as a second step.
The first step is completely generic or at least it was sensed to be so. However there was
no prior formalization of the fact that it is generic. The only class of functions to which
it has been applied is the class of low-degree polynomials and a priori it is not clear how
to even justify the claim of genericity. Here we show that the first step applies to all lifted
codes, thus giving the first justification of the presumed genericity of this step, which we
consider to be a conceptual contribution.
For the second step, the robust analyses in [ALM+ 98, AS03] are quite algebraic and there
seems to be no hope to use them on general lifted codes. The test and analysis of Raz and
Safra [RS97] on the other hand feels much more generic. In this work we use their test, and
extend it to general lifted codes and show that it is robust. Even the extension of the test is
not completely obvious. In particular, to test low-degree polynomials they look at restrictions
of the given function to 2-dimensional “planes”. When lifting 𝑑-dimensional properties, it
is not immediate what would be the dimension of the restrictions the test should look at:
Should it be 𝑑 + 1? Or 2𝑑 or maybe 3𝑑 − 1 (each of which does make logical sense)? We show
that the 2𝑑 dimensional tests are robust, with robustness being independent of 𝑑.
Next we turn to our analysis. In showing robustness of their test, applied to generic lifted
codes there is a major barrier: Almost all analyses of low-degree tests, for polynomials of
degree at most 𝑑, attempt to show first that a function passing the test with high probability
is close to a polynomial of degree twice the degree, i.e., at most 2𝑑, with some additional
features. They then use the distance of the space of polynomials of degree 2𝑑 and the
additional features to establish that the function being tested is really close to a degree
𝑑 polynomial. In extending such analyses to our setting we face two obstacles: In the
completely generic setting, there is no nice notion corresponding to the set of degree 2𝑑
polynomials. One approach might be to consider the linear space spanned by products of
functions in our basic space and work with them, but the algebra gets hairy to understand and
20
analyze. Even if we abandon the complete genericity and stick to the space of polynomials
of degree 𝑑, but now allow 𝑑 > π‘ž/2 we hit a second obstacle: The space of polynomials of
degree 2𝑑 have negligible relative distance compared to the space of polynomials of degree 𝑑.
Thus we need to search for a new proof technique and we find one by unearthing a new
connection between “lifted codes” and “tensor product” codes. The tensor product is a
natural operation in linear algebra and when applied to two linear codes, it produces a new
linear code in a natural way. Tensor products of codes are well-studied in the literature
on coding theory. The testing of tensor product codes was initiated by Ben-Sasson and
Sudan [BSS06] and subsequently has been well-studied [DSW06, Val05, BSV09b, BSV09a,
GGR09]. Specifically, a recent result of Viderman [Vid12] gives a powerful analysis which
we are able to reproduce in a slightly different setting to get our results. In particular this
is the ingredient that allows us to work with base codes whose distance is less than 1/2.
Also, for the sake of the exposition we pretend that this test can test two-dimensional tensor
products of one dimensional codes, with one-dimensional tests. (Actually, the test works
with three dimensional tensors and tests them by looking at two-dimensional planes, but by
suppressing this difference, our exposition becomes a little simpler.)
To explain the connection between lifted codes and tensor product codes, and the idea
that we introduce to test the former, we turn to the simple case of testing a bivariate lift of
a univariate Reed-Solomon code. Specifically, let π’ž be the family of univariate polynomials
of degree at most 𝑑 mapping Fπ‘ž to Fπ‘ž . Let π’ž2 be the family of bivariate polynomials that
become a univariate polynomial of degree at most 𝑑 on every restriction to a line. The tensor
product of π’ž with itself, which we denote π’ž ⊗2 corresponds to the set of bivariate polynomials
of degree at most 𝑑 in each variable. Clearly π’ž2 ⊆ π’ž ⊗2 but such subset relationships are
not of immediate use in testing a code. (Indeed locally testable codes contain many nonLTCs.) To get a tighter relationship, now fix two “directions” 𝑑1 and 𝑑2 and let π’žπ‘‘1 ,𝑑2 be
the code containing all bivariate polynomials over Fπ‘ž that on every restriction to lines in
directions 𝑑1 and 𝑑2 form univariate degree 𝑑 polynomials. On the one hand the code π’žπ‘‘1 ,𝑑2
is just isomorphic to the tensor product code π’ž ⊗2 which is testable by the natural test,
21
by our assumption. On the other hand, we now have π’ž2 = ∩𝑑1 ,𝑑2 π’žπ‘‘1 ,𝑑2 so we now have a
characterization of the lifted codes in terms of the tensor product. One might hope that
one could use this characterization to get a (robust) analysis of the lifted test since it tests
membership in π’žπ‘‘1 ,𝑑2 for random choices of 𝑑1 and 𝑑2 , but unfortunately we do not see a
simple way to implement this hope.
Our key idea is look instead at a more complex family of codes π’žπ‘‘1 ,𝑑2 ,𝑑3 that consists
of functions of degree 𝑑 in directions 𝑑1 , 𝑑2 and 𝑑3 . (Of course now 𝑑1 , 𝑑2 , 𝑑3 are linearly
dependent and so π’žπ‘‘1 ,𝑑2 ,𝑑3 is not a tensor product code. We will return to this issue later.) We
still have π’ž2 = ∩𝑑1 ,𝑑2 ,𝑑3 π’žπ‘‘1 ,𝑑2 ,𝑑3 . Indeed we can even fix 𝑑1 , 𝑑2 arbitrarily (only requiring them
to be linearly independent) and we have π’ž2 = ∩𝑑3 π’žπ‘‘1 ,𝑑2 ,𝑑3 . This view turns out to be more
advantageous since we now have that for any 𝑑3 and 𝑑′3 we have π’žπ‘‘1 ,𝑑2 ,𝑑3 ∪ π’žπ‘‘1 ,𝑑2 ,𝑑′3 ⊆ π’žπ‘‘1 ,𝑑2
which is a code of decent distance. This allows us to show that if the function being tested
is close to π’žπ‘‘1 ,𝑑2 ,𝑑3 for many choices of 𝑑3 then the nearest codewords for all these choices of
𝑑3 are the same. An algebraic analysis of lifted codes tells us that a codeword of π’žπ‘‘1 ,𝑑2 can
not be in π’žπ‘‘1 ,𝑑2 ,𝑑3 for many choices of 𝑑3 without being a codeword of the lifted code and
this lends promise to our idea. But we are not done, since we still need to test the given
function for proximity to π’žπ‘‘1 ,𝑑2 ,𝑑3 and this is no longer a tensor product code so Viderman’s
result does not apply directly. Fortunately, we are able to develop the ideas from Viderman’s
analysis for tensor product codes [Vid12] and apply them also to our case and this yields
our test and analysis. We note that this extension is not immediate — indeed one of the
central properties of tensor product codes is that they are decodable from some clean erasure
patterns and this feature is missing in our codes. Nevertheless the analysis can be modified
to apply to our codes and this suffices to complete the analysis.
In the actual implementation, as noted earlier, we can’t work with univariate tests even
for the simple case above, and work instead by using a bivariate test for trivariate and 4variate functions. (This is similar to the reasons why Raz and Safra used a bivariate test.)
This complicates the notations a bit, but the idea remains similar to the description above.
Our task gets more complicated when the base code being lifted is 𝑑-dimensional for 𝑑 > 1.
22
The most natural adaptation of our analysis leads to dependencies involving 𝛿 (the distance
of the base code) and 𝑑. We work somewhat harder in this case to eliminate any dependence
on 𝑑 while working within the framework described above.
1.2.4
Applications
Lifted Reed-Solomon Code
The most interesting construction arising from lifting is the “lifted Reed-Solomon code”.
As its name suggests, the lifted Reed-Solomon code is obtained by simply lifting the ReedSolomon code, a univariate code, to π‘š dimensions. The lifted Reed-Solomon, by definition,
contains the Reed-Muller code. However, if the degree 𝑑 of the Reed-Solomon code is sufficiently large relative to the field size — in particular, if 𝑑 ≥ π‘ž − π‘ž/𝑝, where Fπ‘ž has characteristic 𝑝 — then the lifted Reed-Solomon code contains polynomials of high degree as well.
This fact follows from the characterization of low-degree polynomials as proven in [KR06].
In fact, using structural results about affine-invariance and lifts, we easily re-prove a special
case of the characterization of low-degree polynomials in Theorem 7.1.6. The lifted ReedSolomon code is arguably more natural than the Reed-Muller code, the latter of which is an
unnecessarily sparse subcode of the former. We therefore expect the lifted Reed-Solomon
code to exhibit the same versatility as that of the Reed-Muller code, and indeed since the
lifted Reed-Solomon code is a lifted code, it generically has good distance and is locally
decodable and testable. We show that it is also in fact (locally) list-decodable up to the
Johnson radius, just like the Reed-Muller code. What sets the lifted Reed-Solomon code
apart from the Reed-Muller code is its vastly greater rate. If one insists on having positive
distance 𝛿 > 0, then the π‘š-variate Reed-Muller code has rate bounded by 1/π‘š!, whereas
the rate of the lifted Reed-Solomon code approaches 1 as 𝛿 → 0.
The first family of high-rate locally correctable codes known were the multiplicity codes of
Kopparty, Saraf, and Yekhanin [KSY14]. Kopparty [Kop12] showed that multiplicity codes
are also locally list-decodable up to the Johnson radius. The only prior construction of
high-rate codes that are robustly testable are the tensor product codes of Viderman [Vid12].
23
Thus, the lifted Reed-Solomon code is the first high-rate code known to be locally correctable,
locally list-decodable up to the Johnson radius, and robustly testable. This is the code in
Theorem 1.2.1.
Robust Low-Degree Testing
An almost direct corollary of our robustness result for lifted codes is a π‘ž 4 -local robust lowdegree test for the setting 𝑑 ≤ (1 − 𝛿)π‘ž. To see why we get π‘ž 4 queries, note that when
𝑑 ≥ π‘ž − π‘ž/𝑝, the π‘š-variate Reed-Muller code of degree 𝑑 is not equal to the π‘š-dimensional
lift of the degree 𝑑 Reed-Solomon code. But the latter turns out to be the π‘š-dimensional
lift of the bivariate Reed-Muller code of degree 𝑑. Applying our robust testing result to this
lifted family yields a robust test making π‘ž 4 queries. But with some slight extra work we can
get a better tester that makes only π‘ž 2 queries and this yields the following theorem.
Theorem 1.2.3. ∀𝛿 > 0 ∃𝛼 > 0 such that the following holds: For every finite field Fπ‘ž ,
for every integer 𝑑 ≤ (1 − 𝛿)π‘ž and every positive integer π‘š, there is a π‘ž 2 -query 𝛼-robust
low-degree test for the class of π‘š-variate polynomials of degree at most 𝑑 over Fπ‘ž .
We note that previous works on low-degree testing worked only when 𝑑 < π‘ž/2. This
ratio seems to be achieved by Friedl and Sudan (see [FS95, Theorem 13]). Other works
[RS96, ALM+ 98, RS97, AS03, MR06] seem to achieve weaker ratios for a variety of reasons
discussed above.
Nikodym Set Size Bounds
One of the applications of lifted codes is to bounding, from below, the size of “Nikodym
sets” over finite fields (of small characteristic). A set 𝑆 ⊆ Fπ‘š
π‘ž is a Nikodym set if every point
x has a line passing through it such that all points of the line, except possibly the point x
itself, are elements of 𝑆. Nikodym sets are closely related to “Kakeya sets” — the latter
contain a line in every direction, while the former contain almost all of a line through every
point. A lower bound for Kakeya sets over finite fields was proved by Dvir [Dvi08] using the
polynomial method and further improved by using the “method of multiplicites” by Saraf and
24
Sudan [SS08] and Dvir et al. [DKSS09]. Kakeya sets have seen applications connecting its
study to the study of randomness extractors, especially [DS07, DW11]. Arguably, Nikodym
sets are about as natural in this connection as Kakeya sets.
Previous lower bounds on Kakeya sets were typically also applicable to Nikodym sets
and led to bounds of the form |𝑆| ≥ (1 − π‘œ(1))π‘ž π‘š /2π‘š where the π‘œ(1) term goes to zero as
π‘ž → ∞. In particular, previous lower bounds failed to separate the growth of Nikodym sets
from those of Kakeya sets. In our work, we present a simple connection that shows that the
existence of the lifted Reed-Solomon code yields a large lower bound on the size of Nikodym
sets, thereby significantly improving the known lower bound on the size of Nikodym sets
over fields of constant characteristic.
Theorem 1.2.4. For every prime 𝑝, and every integer π‘š, there exists πœ– = πœ–(𝑝, π‘š) > 0
such that for every finite field Fπ‘ž of characteristic 𝑝, if 𝑆 ⊆ Fπ‘š
π‘ž is a Nikodym set, then
|𝑆| ≥ π‘ž π‘š − π‘ž (1−πœ–)π‘š . In particular, if π‘ž → ∞, then |𝑆| ≥ (1 − π‘œ(1)) · π‘ž π‘š .
Thus, whereas previous lower bounds on the size of Nikodym sets allowed for the possibility that the density of the Nikodym sets vanishes as π‘š grows, ours show that Nikodym
sets occupy almost all of the space. One way to view our results is that they abstract the
polynomial method in a more general way, and thus lead to stronger lower bounds in some
cases.
1.2.5
Organization
In Chapter 2, we establish notation and preliminary definitions. In Chapter 3, we formally
define error correcting codes and relevant models: local decoding, correcting, list-decoding,
and testing. In Chapter 4, we review some structural properties of affine-invariant codes. In
Chapter 5, we formally define the lifting operation and prove generic properties of lifted codes.
In Chapter 6, we prove that lifted codes are robustly testable with robustness parameter
that depends only on the distance of the base code. In Chapter 7, we discuss applications
of lifted codes: construction of the lifted Reed-Solomon code, robust low-degree testing, and
new lower bounds on the size of Nikodym sets.
25
26
Chapter 2
Preliminaries
2.1
Notation
Letters. We will typically use lower-case italic letters (e.g. π‘Ž, 𝑏, 𝑐) to denote scalar values, lower-case bold letters (e.g. a, b, c) to denote vectors, and upper-case bold letters (e.g.
A, B, C) to denote matrices. If A is a matrix, then we denote by A𝑖* the 𝑖-th row of A and
by A*𝑗 the 𝑗-th row of A.
Sets and functions. For a set 𝑆 and 𝑛 ∈ N, let
(︀𝑆 )οΈ€
𝑛
be the collection of subsets 𝑇 ⊆ 𝑆
with |𝑇 | = 𝑛. Let 2𝑆 be the collection of all subsets of 𝑆. For a positive integer 𝑛, define
[𝑛] , {1, . . . , 𝑛} and J𝑛K , {0, 1, . . . , 𝑛 − 1}. For sets 𝐴 and 𝐡, let {𝐴 → 𝐡} denote the set
of functions from 𝐴 to 𝐡.
Vectors and Hamming distance. If a ∈ Nπ‘š , let β€–aβ€– ,
∑οΈ€π‘š
𝑖=1
π‘Žπ‘– . If Σ is a finite set,
𝑛 ∈ N, and s ∈ Σ𝑛 is a string, then let 𝑠𝑖 denote the 𝑖-th component of s, for 𝑖 ∈ [𝑛], so that
s = (𝑠1 , . . . , 𝑠𝑛 ). For two vectors a, b ∈ Σ𝑛 , denote their Hamming distance by Δ(a, b) ,
#{𝑖 ∈ [𝑛] | π‘Žπ‘– ΜΈ= 𝑏𝑖 } and their normalized Hamming distance by 𝛿(a, b) ,
|Δ(a,b)|
.
𝑛
If 𝑆 ⊆ Σ𝑛
is a set, then 𝛿(a, 𝑆) , minb∈𝑆 𝛿(a, b) is the distance from a to 𝑆. If Σ is a field, then the
(resp. normalized) Hamming weight of a ∈ Σ𝑛 is (resp. 𝛿(a, 0)) Δ(a, 0). We will frequently
think of functions 𝑓 : 𝐴 → 𝐡 as vectors in 𝐡 𝐴 and so we extend the vector notations to
27
functions as well. In particular, if 𝑓, 𝑔 : 𝐴 → 𝐡, then Δ(𝑓, 𝑔) , #{π‘₯ ∈ 𝐴 | 𝑓 (π‘₯) ΜΈ= 𝑔(π‘₯)} and
𝛿(𝑓, 𝑔) ,
Δ(𝑓,𝑔)
.
|𝐴|
Note that 𝛿(𝑓, 𝑔) = Prπ‘₯∈𝐴 [𝑓 (π‘₯) ΜΈ= 𝑔(π‘₯)].
Minkowski sums and affine subspaces. For sets 𝐴, 𝐡 ⊆ Fπ‘š , their Minkowski sum
is denoted 𝐴 + 𝐡 , {a + b | a ∈ 𝐴, b ∈ 𝐡}. For 𝐴 ⊆ Fπ‘š , the span of 𝐴 is denoted
∑οΈ€
span(𝐴) , { a∈𝐴 𝑐a · a | 𝑐a ∈ F}. For x ∈ Fπ‘š and 𝐴 ⊆ Fπ‘š , let (x, 𝐴) , {x} + span(𝐴) be
the affine subspace through x in directions 𝐴.
Shadows. Let 𝑝, π‘Ž, 𝑏 ∈ N and π‘Ž =
∑οΈ€
𝑖≥0
π‘Ž(𝑖) 𝑝𝑖 , 𝑏 =
∑οΈ€
𝑖≥0
𝑏(𝑖) 𝑝𝑖 with π‘Ž(𝑖) , 𝑏(𝑖) ∈ J𝑝K for 𝑖 ≥ 0.
Then π‘Ž is in the 𝑝-shadow of 𝑏, denoted by π‘Ž ≤𝑝 𝑏, if π‘Ž(𝑖) ≤ 𝑏(𝑖) for 𝑖 ≥ 0. If a, b ∈ Nπ‘š , then
a ≤𝑝 b means that π‘Žπ‘– ≤𝑝 𝑏𝑖 for every 𝑖 ∈ [π‘š]. If a ∈ N𝑛 and 𝑏 ∈ N, then a ≤𝑝 𝑏 means that
∑οΈ€
π‘š×𝑛
and b ∈ Nπ‘š , then A ≤𝑝 b means that
𝑖∈𝑆 π‘Žπ‘– ≤𝑝 𝑏 for any subset 𝑆 ⊆ [𝑛]. If A ∈ N
A𝑖* ≤𝑝 b for every 𝑖 ∈ [π‘š].
Finite fields and polynomials. If 𝑝 is a prime and π‘ž is a power of 𝑝, then Fπ‘ž denotes
the finite field of size π‘ž. If X = (𝑋1 , . . . , π‘‹π‘š ) are variables and d ∈ Nπ‘š , then Xd denotes
∏οΈ€
𝑑𝑖
the monomial π‘š
𝑖=1 𝑋𝑖 . Any polynomial β„Ž(X) ∈ Fπ‘ž [X] with such that β„Ž(X) = 𝑓 (X) +
∏οΈ€
π‘ž
π‘š
𝑔(X) π‘š
𝑖=1 (𝑋𝑖 − 𝑋𝑖 ) for some 𝑔(X) ∈ Fπ‘ž [X] defines the same function Fπ‘ž → Fπ‘ž , since the
polynomial 𝑋 π‘ž − 𝑋 is identically zero on Fπ‘ž . Observe that every function 𝑓 : Fπ‘š
π‘ž → Fπ‘ž can
∑οΈ€
be expressed uniquely as a linear combination 𝑓 (X) = d∈Jπ‘žKπ‘š 𝑓d ·Xd of monomials Xd with
d ∈ Jπ‘žKπ‘š . Throughout this thesis, when we refer to a “polynomial” 𝑓 : Fπ‘š
π‘ž → Fπ‘ž , we mean
the unique 𝑓 (X) defined above. The support of 𝑓 is supp(𝑓 ) , {d ∈ Jπ‘žKπ‘š | 𝑓d ΜΈ= 0}. The
degree of a polynomial 𝑓 : Fπ‘š
π‘ž → Fπ‘ž is deg(𝑓 ) , max{β€–dβ€– | d ∈ supp(𝑓 )}. When we refer
π‘š
to a polynomial 𝑓 : Fπ‘š
π‘ž → Fπ‘ž in the context of codewords, we refer to the π‘ž -dimensional
vector (𝑓 (x))x∈Fπ‘š
.
π‘ž
28
Mod-star. For π‘Ž ∈ N and 𝑏 > 1, define the operation mod* by
π‘Ž mod* 𝑏 =
⎧
βŽͺ
βŽ¨π‘Ž
π‘Ž<𝑏
βŽͺ
βŽ©π‘Ž mod (𝑏 − 1) π‘Ž ≥ 𝑏
*
so that 𝑋 π‘Ž ≡ 𝑋 π‘Ž mod
2.2
𝑏
(mod 𝑋 𝑏 − 𝑋).
Probability and Concentration bounds
If 𝑋 is a random variable, we use E[𝑋] and Var[𝑋] to denote the expectation and variance
of 𝑋, respectively. If the probability space of 𝑋 is not clear from context, we use subscripts,
e.g. if 𝑋 = 𝑋(π‘Ž, 𝑏), then Eπ‘Ž [𝑋] is the average over π‘Ž with 𝑏 fixed.
Proposition 2.2.1 (Markov inequality). Let 𝑋 ≥ 0 be a random variable with E[𝑋] < ∞
and let π‘Ž > 0. Then Pr[𝑋 ≥ π‘Ž] ≤
E[𝑋]
.
π‘Ž
Proposition 2.2.2 (Chebyshev inequality). Let 𝑋 be a random variable with E[𝑋] < ∞
and Var[𝑋] < ∞, and let π‘Ž > 0. Then Pr[|𝑋 − E[𝑋]| ≥ π‘Ž] ≤
Var[𝑋]
.
π‘Ž2
Proposition 2.2.3 (Hoeffding inequality). Let 𝑋1 , . . . , 𝑋𝑛 ∈ [0, 1] be independent random
βƒ’
[οΈ€βƒ’
]οΈ€
∑οΈ€
variables and let 𝑋 , 𝑛1 𝑛𝑖=1 𝑋𝑖 . Let πœ– > 0. Then Pr ⃒𝑋 − E[𝑋]βƒ’ > πœ– ≤ 2 exp(−2π‘›πœ–2 ).
ΜƒοΈ€
Proposition 2.2.4. Let 𝑓, 𝑔 : Fπ‘š
π‘ž → Fπ‘ž . Let 𝛿(𝑓, 𝑔) be the estimate of 𝛿(𝑓, 𝑔) by random
sampling, i.e. independent uniformly random x ∈ Fπ‘š
π‘ž are chosen and the average of 1𝑓 (x)ΜΈ=𝑔(x)
is output. If Θ(ln(1/πœ‚)/πœ–2 ) queries are used in the sample, then with probability at least 1−πœ‚,
ΜƒοΈ€ 𝑔) has additive error at most πœ–.
the estimate 𝛿(𝑓,
Proof. Let 𝑛 ≥ ln(2/πœ‚)/(2πœ–2 ). For each 𝑖 ∈ [𝑛], let 𝑋𝑖 , 1𝑓 (x𝑖 )ΜΈ=𝑔(x𝑖 ) , and define 𝑋 ,
∑︀𝑛
1
ΜƒοΈ€
𝑖=1 𝑋𝑖 so that 𝛿(𝑓, 𝑔) = 𝑋, and 𝛿(𝑓, 𝑔) = E[𝑋]. By Proposition 2.2.3,
𝑛
βƒ’
[︁⃒
]︁
βƒ’ΜƒοΈ€
βƒ’
Pr ⃒𝛿(𝑓,
𝑔) − 𝛿(𝑓, 𝑔)βƒ’ > πœ– ≤ 2 exp(−2π‘›πœ–2 ) ≤ πœ‚.
29
30
Chapter 3
Error Correcting Codes
3.1
Classical Parameters
Error correcting codes are schemes for encoding messages as codewords to protect them from
noise. The code itself is the subset of valid codewords. The rate of a code is the ratio of
the message length to the encoding length, and measures the efficiency of the encoding. The
distance of a code measures the minimum distance between valid codewords, and indicates
the error-correcting capability of the code. In this section, we formally define these notions.
Definition 3.1.1 (Code). Let Σ be a finite set and let 𝑛 be a natural number. A code over
Σ of block length 𝑛 is a subset π’ž ⊆ Σ𝑛 . If there exists a set Σ0 , integer π‘˜ ≤ 𝑛, and injective
function Enc : Σπ‘˜0 → Σ𝑛 such that Enc(Σπ‘˜0 ) = π’ž, then Enc is an encoding function for π’ž.
Definition 3.1.2 (Rate of a code). The rate of a code π’ž ⊆ Σ𝑛 is
log |π’ž|
.
𝑛 log |Σ|
Definition 3.1.3 (Distance of a code). The (normalized) distance of a code π’ž ⊆ Σ𝑛 is
𝛿(π’ž) , minxΜΈ=y∈π’ž 𝛿(x, y).
The following proposition shows that, algorithmic efficiency aside, any code supports
unique decoding up to half its minimum distance.
Proposition 3.1.4. For every r ∈ Σ𝑛 , there exists at most one c ∈ π’ž with 𝛿(r, c) <
31
𝛿(π’ž)
.
2
Proof. Suppose c, c′ ∈ π’ž and 𝛿(r, c), 𝛿(r, c′ ) <
𝛿(π’ž)
.
2
Then, by the triangle inequality,
𝛿(c, c′ ) ≤ 𝛿(c, r) + 𝛿(r, c′ ) < 𝛿(π’ž), so c = c′ .
Linear codes are codes whose alphabet is a (finite) field and whose codewords form a
vector space over the field. Every code of interest to us in this thesis is a linear code.
Definition 3.1.5 (Linear code). A code π’ž ⊆ Σ𝑛 is linear if Σ = F is a field and π’ž is a linear
subspace of F𝑛 .
Proposition 3.1.6. If π’ž ⊆ F𝑛 is a linear code, then 𝛿(π’ž) is equal to the minimal normalized
Hamming weight of nonzero c ∈ π’ž.
Proof. Let c ∈ π’ž be nonzero of minimal Hamming weight. By definition, 𝛿(π’ž) ≤ 𝛿(c, 0).
On the other hand, for any two distinct x, y ∈ π’ž, 𝛿(x, y) = 𝛿(x − y, 0) ≥ 𝛿(c, 0), and so
minimizing over x ΜΈ= y, we have 𝛿(π’ž) ≥ 𝛿(c, 0).
3.2
Local decoding, correcting, and list-decoding
In this section, we formally define the models of local decoding, local correcting, and local
list-decoding. Intuitively, local decoding entails recovering a symbol of the original message
using few queries, while local correcting entails recovering a symbol of the original codeword
using few queries.
Definition 3.2.1 (Local decoding). A code π’ž ⊆ Σ𝑛 with encoding function Enc : Σπ‘˜0 → Σ𝑛
is (β„“, 𝜏, πœ–)-locally decodable if there exists a randomized oracle π’œ : [π‘˜] → Σ0 with oracle access
to a received word r ∈ Σ𝑛 such that
1. π’œπ‘Ÿ queries at most β„“ symbols of r;
2. if there is m ∈ Σπ‘˜0 with 𝛿(Enc(m), r) ≤ 𝜏 , then Pr [π’œπ‘Ÿ (𝑖) = π‘šπ‘– ] ≥ 1−πœ– for every 𝑖 ∈ [π‘˜].
Definition 3.2.2 (Local correcting). A code π’ž ⊆ Σ𝑛 is (β„“, 𝜏, πœ–)-locally correctable if there
exists a randomized oracle π’œ : [𝑛] → Σ with oracle access to a received word r ∈ Σ𝑛 such
that
32
1. π’œπ‘Ÿ queries at most β„“ symbols of r;
2. if there is c ∈ π’ž with 𝛿(c, r) ≤ 𝜏 , then Pr [π’œπ‘Ÿ (𝑖) = 𝑐𝑖 ] ≥ 1 − πœ– for every 𝑖 ∈ [𝑛].
If π’ž is a linear code, then it is possible to encode π’ž in systematically, i.e. such that the
original message is part of the codeword. Of course, this is not algorithmically satisfying
unless the systematic encoding function is explicit, i.e. computable in polynomial time. When
π’ž is linear, we can think of it as a space of functions {𝑆 → Fπ‘ž }. A systematic encoding is
then equivalent to finding an interpolating set for π’ž in 𝑆.
Definition 3.2.3. If π’ž ⊆ {𝑆 → Fπ‘ž } is linear, then 𝐼 ⊆ 𝑆 is an interpolating set for π’ž if, for
every 𝑓 : 𝐼 → Fπ‘ž , there exists a unique extension 𝑔 ∈ π’ž such that 𝑔|𝐼 = 𝑓 .
Remark 3.2.4. If π’ž ⊆ {𝑆 → Fπ‘ž } has an interpolating set 𝐼 ⊆ 𝑆, then |π’ž| = π‘ž |𝐼| . In
particular, if π’ž is linear, then |𝐼| = dimFπ‘ž (π’ž).
Proposition 3.2.5. If π’ž ⊆ {𝑆 → Fπ‘ž } is linear and has an explicit interpolating set 𝐼 ⊆ 𝑆,
and π’ž is a (β„“, 𝜏, πœ–)-locally correctable code, then π’ž is a (β„“, 𝜏, πœ–)-locally decodable code.
Proof. Let Enc be the map which takes 𝑓 : 𝐼 → Fπ‘ž to its unique extension 𝑔 ∈ π’ž such that
𝑔|𝐼 = 𝑓 , guaranteed by the fact that 𝐼 is an interpolating set. Then the local correcting
algorithm for π’ž also serves as the local decoding algorithm, when restricted to 𝐼.
A local list-decoding algorithm outputs a list of oracles, such that each valid codeword
within the given radius is computed by some oracle in the output list.
Definition 3.2.6 (Local list-decoding). A code π’ž ⊆ Σ𝑛 is (β„“1 , β„“2 , 𝜏, 𝐿, πœ–, πœ‚)-locally listdecodable if there exists a randomized algorithm π’œ with oracle access to a received word
r ∈ Σ𝑛 that outputs a list 𝑀1 , . . . , 𝑀𝐿 : [𝑛] → Σ of randomized oracles with oracle access to
r, such that
1. π’œπ‘Ÿ queries at most β„“1 symbols of r;
2. for each 𝑗 ∈ [𝐿], 𝑀𝑗 queries at most β„“2 symbols of r;
3. with probability at least 1 − πœ‚, the following holds: if there is c ∈ π’ž with 𝛿(c, r) ≤ 𝜏 ,
then there is some 𝑗 ∈ [𝐿] such that Pr[𝑀 π‘Ÿ (𝑖) = 𝑐𝑖 ] ≥ 1 − πœ– for every 𝑖 ∈ [𝑛].
33
3.3
Local testing and robust testing
In this section, we formally define the model of testing. Since we will be solely interested in
the testing of linear codes, we only present the definition of local testing in the context of
linear codes. We also define the notions of soundness and robustness, and prove some simple
relationships between the two.
Definition 3.3.1 (Local testing). A β„“-local tester for a code π’ž ⊆ {𝑆 → Fπ‘ž } is a randomized
algorithm 𝒯 with oracle access to a received word 𝑓 : 𝑆 → Fπ‘ž , which randomly samples
β„“
(πœ‹, 𝑉 ) according to some distribution π’Ÿ on (𝑆 β„“ , 2Fπ‘ž ), with πœ‹ = (πœ‹1 , . . . , πœ‹β„“ ) ∈ 𝑆 β„“ and 𝑉 ⊆ Fβ„“π‘ž
and accepts if and only if 𝑓 |πœ‹ , (𝑓 (πœ‹1 ), . . . , 𝑓 (πœ‹β„“ )) ∈ 𝑉 .
The tester is πœ–-sound if 𝒯 accepts 𝑓 ∈ π’ž with probability one, and rejects 𝑓 ∈
/ π’ž with
probability at least πœ– · 𝛿(𝑓, π’ž).
The tester is 𝛼-robust if E(πœ‹,𝑉 )←π’Ÿ [𝛿 (𝑓 |πœ‹ , 𝑉 )] ≥ 𝛼 · 𝛿(𝑓, π’ž).
Proposition 3.3.2. Let 𝒯 be an β„“-local tester for π’ž. If 𝒯 is πœ–-sound, then 𝒯 is (πœ–/β„“)-robust.
If 𝒯 is 𝛼-robust, then 𝒯 is 𝛼-sound.
Proof. Suppose 𝒯 is πœ–-sound. Observe that if 𝑓 |πœ‹ ∈
/ 𝑉 , then 𝛿(𝑓 |πœ‹ , 𝑉 ) ≥ 1/β„“. Therefore,
E(πœ‹,𝑉 )←π’Ÿ [𝛿 (𝑓 |πœ‹ , 𝑉 )] = E(πœ‹,𝑉 )←π’Ÿ [𝛿 (𝑓 |πœ‹ , 𝑉 ) | 𝑓 |πœ‹ ∈
/ 𝑉]·
≥ (1/β„“) ·
Pr
(πœ‹,𝑉 )←π’Ÿ
Pr
(πœ‹,𝑉 )←π’Ÿ
[𝑓 |πœ‹ ∈
/ 𝑉]
[𝑓 |πœ‹ ∈
/ 𝑉]
(3.1)
(3.2)
≥ (1/β„“) · πœ– · 𝛿(𝑓, π’ž).
(3.3)
Now suppose 𝒯 is 𝛼-robust. Then
Pr
(πœ‹,𝑉 )←π’Ÿ
[𝑓 |πœ‹ ∈
/ 𝑉 ] ≥ E(πœ‹,𝑉 )←π’Ÿ [𝛿 (𝑓 |πœ‹ , 𝑉 ) | 𝑓 |πœ‹ ∈
/ 𝑉]·
Pr
(πœ‹,𝑉 )←π’Ÿ
[𝑓 |πœ‹ ∈
/ 𝑉]
(3.4)
= E(πœ‹,𝑉 )←π’Ÿ [𝛿 (𝑓 |πœ‹ , 𝑉 )]
(3.5)
≥ 𝛼 · 𝛿(𝑓, π’ž).
(3.6)
34
Proposition 3.3.3. Let π’―π’ž be an β„“1 -local tester for π’ž with distribution π’Ÿπ’ž , and suppose
for every (πœ‹, 𝑉 ) in the support of π’Ÿ, 𝑉 has an β„“2 -local tester 𝒯𝑉 with distribution π’Ÿπ‘‰ , and
β„“2 ≤ β„“1 .
1. If π’―π’ž is 𝛼1 -robust and 𝒯𝑉 is 𝛼2 -robust for every (πœ‹, 𝑉 ) in the support of π’Ÿπ’ž , then π’ž has
an β„“2 -local tester that is (𝛼1 · 𝛼2 )-robust.
2. If π’―π’ž is 𝛼-robust and 𝒯𝑉 is πœ–-sound for every (πœ‹, 𝑉 ) in the support of π’Ÿπ’ž , then π’ž has
an β„“2 -local tester that is (𝛼 · πœ–)-sound.
Proof. Let π’Ÿ be the following distribution: choose (πœ‹, 𝑉 ) ← π’Ÿπ’ž and then choose and output
(πœ‹ ′ , 𝑉 ′ ) ← π’Ÿπ‘‰ . Let 𝒯 be the β„“2 -tester for π’ž with distribution π’Ÿ.
1. If π’―π’ž is 𝛼1 -robust and 𝒯𝑉 is 𝛼2 -robust, then
[οΈ€
]οΈ€
E(πœ‹′ ,𝑉 ′ )←π’Ÿ [𝛿 (𝑓 |πœ‹′ , 𝑉 ′ )] = E(πœ‹,𝑉 )←π’Ÿπ’ž E(πœ‹′ ,𝑉 ′ )←π’Ÿπ‘‰ [𝛿 (𝑓 |πœ‹′ , 𝑉 ′ )]
(3.7)
≥ 𝛼2 · E(πœ‹,𝑉 )←π’Ÿπ’ž [𝛿 (𝑓 |πœ‹ , 𝑉 )]
(3.8)
= 𝛼1 · 𝛼2 · 𝛿(𝑓, π’ž).
(3.9)
2. If π’―π’ž is 𝛼-robust and 𝒯𝑉 is πœ–-sound, then
Pr
(πœ‹ ′ ,𝑉 ′ )←π’Ÿ
3.4
[οΈ‚
′
[𝑓 |πœ‹′ ∈
/ 𝑉 ] = E(πœ‹,𝑉 )←π’Ÿπ’ž
Pr
(πœ‹ ′ ,𝑉 ′ )←π’Ÿπ‘‰
]οΈ‚
[𝑓 |πœ‹′ ∈
/𝑉 ]
′
(3.10)
≥ πœ– · E(πœ‹,𝑉 )←π’Ÿπ’ž [𝛿 (𝑓 |πœ‹ , 𝑉 )]
(3.11)
≥ 𝛼 · πœ– · 𝛿(𝑓, π’ž).
(3.12)
Codes
In this section, we present two of the most ubiquitous linear codes: the Reed-Solomon code
and the Reed-Muller code. We will directly use the Reed-Solomon code in our constructions,
35
whereas the Reed-Muller code serves as a benchmark for comparison.
3.4.1
Reed-Solomon code
The Reed-Solomon code consists of evaluations of low-degree univariate polynomials over a
finite field Fπ‘ž .
Definition 3.4.1. Let π‘ž be a prime power, and let 𝑑 ∈ N. The Reed-Solomon code RS(π‘ž, 𝑑)
of degree 𝑑 over Fπ‘ž is the code RS(π‘ž, 𝑑) , {𝑓 : Fπ‘ž → Fπ‘ž | deg(𝑓 ) ≤ 𝑑}.
Proposition 3.4.2. If 𝑑 < π‘ž, then the Reed-Solomon code RS(π‘ž, 𝑑) has distance 1 −
rate
𝑑
π‘ž
and
𝑑+1
.
π‘ž
In [GS99], Guruswami and Sudan showed that the Reed-Solomon code of distance 𝛿 > 0
√
can be efficiently list-decoded up to the Johnson radius 1− 1 − 𝛿. We will use this algorithm
as a subroutine in our list-decoding and local list-decoding algorithms for the lifted ReedSolomon code in Sections 7.1.3 and 7.1.4, respectively.
Theorem 3.4.3 (Guruswami-Sudan list-decoding [GS99]). For every 𝛿, πœ– > 0, there is a
polynomial time algorithm taking as input a function π‘Ÿ : Fπ‘ž → Fπ‘ž and outputs a list β„’ of size
√
|β„’| = 𝑂(1/πœ–2 ) satisfying the following: if 𝑐 ∈ RS(π‘ž, (1 − 𝛿)π‘ž) and 𝛿(π‘Ÿ, 𝑐) < 1 − 1 − 𝛿 − πœ–,
then 𝑐 ∈ β„’.
3.4.2
Reed-Muller code
The Reed-Muller code consists of evaluations of low-degree multivariate polynomials over a
finite field Fπ‘ž .
Definition 3.4.4. Let π‘ž be a prime power, and let 𝑑, π‘š ∈ N. The π‘š-variate Reed-Muller
code RM(π‘ž, 𝑑, π‘š) of degree 𝑑 over Fπ‘ž is the code RM(π‘ž, 𝑑, π‘š) , {𝑓 : Fπ‘š
π‘ž → Fπ‘ž | deg(𝑓 ) ≤ 𝑑}.
Remark 3.4.5. It follows immediately from definitions that RM(π‘ž, 𝑑, 1) = RS(π‘ž, 𝑑).
Proposition 3.4.6. If 𝑑 < π‘ž, then the Reed-Muller code RM(π‘ž, 𝑑, π‘š) has distance 1 − π‘‘π‘ž and
(𝑑+π‘š)
rate π‘žπ‘šπ‘š .
36
Reed-Muller codes play a prominent role in complexity theory due to their locality features. Reed-Muller codes are locally decodable/correctable, and are also list-decodable
and locally list-decodable up to the Johnson radius [PW04, STV99, BK09]. Moreover,
Reed-Muller codes are testable, even robustly [RS96, ALM+ 98, FS95, AS03, RS97, MR06,
AKK+ 05, KR06, JPRZ09, BKS+ 10, HSS11].
Note that if we want a family of Reed-Muller codes with positive distance 𝛿 > 0, we
need the degree 𝑑 = (1 − 𝛿)π‘ž. The rate is therefore roughly
(1−𝛿)π‘š
π‘š!
<
1
.
π‘š!
In particular, the
rate never exceeds 12 . The multiplicity codes of [KSY14] were the first locally correctable
codes with rate close to 1. The highlight of our work is the construction of codes with
the same distance as that of the Reed-Muller code and same locality features (decodability,
correctability, list-decodability, and testability), but with rate close to 1.
37
38
Chapter 4
Affine-invariance
4.1
Affine-invariant Codes
In [KS08], Kaufman and Sudan examine the role of symmetry in algebraic property testing
(testing of linear codes). The type of symmetry they focus on is affine-invariance. Viewing
codewords as functions 𝑓 : Fπ‘š
𝑄 → Fπ‘ž , where F𝑄 is an extension field of Fπ‘ž , one can permute
π‘š
the symbols of 𝑓 by applying a permutation πœ‹ : Fπ‘š
𝑄 → F𝑄 to the domain, resulting in a new
word 𝑓 ∘ πœ‹. Affine-invariance is simply the property of being closed under applying affine
permutations to the domain, i.e. if 𝑓 is a codeword, then 𝑓 ∘ 𝐴 is a codeword for any affine
permutation 𝐴.
Definition 4.1.1 (Affine-invariance). A code π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } is affine-invariant if 𝑓 ∘𝐴 ∈ π’ž
π‘š
whenever 𝑓 ∈ π’ž and 𝐴 : Fπ‘š
π‘ž → Fπ‘ž is an affine permutation.
Affine-invariance appears to be the right abstraction of the low-degree property. In fact,
when 𝑄 = π‘ž is prime, then the only affine-invariant codes are the Reed-Muller codes [KS08].
However, when π‘ž is a prime power or if 𝑄 is a power of π‘ž, then there is a richer collection of
affine-invariant codes. Affine-invariant codes are particularly appealing because they possess
rich structure. The main structural feature of affine-invariant codes we will use is that they
are spanned by monomials.
39
π‘š
Definition 4.1.2 (Degree set). A code π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } has a degree set Deg(π’ž) ⊆ Jπ‘žK
if π’ž = {𝑓 : Fπ‘š
π‘ž → Fπ‘ž | supp(𝑓 ) ⊆ Deg(π’ž)}. The degree set Deg(π’ž) is 𝑝-shadow-closed if,
whenever d ∈ Deg(π’ž) and e ≤𝑝 d, we have e ∈ Deg(π’ž).
Proposition 4.1.3 ([KS08]). If π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } is linear affine-invariant, where Fπ‘ž has
characteristic 𝑝, then π’ž has a 𝑝-shadow-closed degree set.
Proposition 4.1.4 ([BGM+ 11a]). If π’ž ⊆ {Fπ‘žπ‘š → Fπ‘ž } is a Fπ‘ž -linear affine-invariant code,
then dimFπ‘ž (π’ž) = | Deg(π’ž)|.
4.2
Equivalence of Invariance under Affine Transformations and Permutations
In their work initiating the study of the testability of affine-invariant properties (codes),
Kaufman and Sudan [KS07] studied properties closed under general affine transformations
and not just permutations. While affine transformations are nicer to work with when available, they are not mathematical elegant (they do not form a group under composition).
Furthermore in the case of codes they also do not preserve the code — they only show that
every codeword stays in the code after the transformation. Among other negative features
affine transformations do not even preserve the weight of non-zero codewords, which can
lead to some rude surprises. Here we patch the gap by showing that families closed under
affine permutations are also closed under affine transformations. So one can assume the
latter, without restricting the class of properties under consideration. We note that such
a statement was proved in [BGM+ 11b] for the case of univariate functions. Unfortunately
their proof does not extend to the multivariate setting and forces us to rework many steps
from [KS08].
Theorem 4.2.1. If π’ž ⊆ {Fπ‘š
𝑄 → Fπ‘ž } is an Fπ‘ž -linear code invariant under affine permutations,
then π’ž is invariant under all affine transformations.
40
The central lemma (Lemma 4.2.2) that we prove is that every non-trivial function can
be split into more basic ones. This leads to a proof of Theorem 4.2.1 fairly easily.
We first start with the notion of a basic function. For 𝑄 = π‘ž 𝑛 , let Tr : F𝑄 → Fπ‘ž denote
the trace function Tr(π‘₯) = π‘₯ + π‘₯π‘ž + · · · + π‘₯π‘ž
𝑛−1
. We say that 𝑓 : Fπ‘š
𝑄 → Fπ‘ž is a basic function
if 𝑓 (X) = Tr(πœ†Xd ) for some d ∈ J𝑄Kπ‘š . For π’ž ⊆ {Fπ‘š
𝑄 → Fπ‘ž } and 𝑓 ∈ π’ž we say 𝑓 can be split
(in π’ž) if there exist functions 𝑔, β„Ž ∈ π’ž such that 𝑓 = 𝑔 + β„Ž and supp(𝑔), supp(β„Ž) ( supp(𝑓 ).
Lemma 4.2.2. If π’ž ⊆ {Fπ‘š
𝑄 → Fπ‘ž } is an Fπ‘ž -linear code invariant under affine permutations,
then for every function 𝑓 ∈ π’ž, 𝑓 is either basic or 𝑓 can be split.
We first prove Theorem 4.2.1 from Lemma 4.2.2.
Proof of Theorem 4.2.1. First we assert that it suffices to prove that for every function 𝑓 ∈ π’ž
the function 𝑓˜ = 𝑓 (𝑋1 , . . . , π‘‹π‘š−1 , 0) is also in π’ž. To see this, consider 𝑓 ∈ π’ž and 𝐴 : Fπ‘š
𝑄 →
π‘š
π‘š
Fπ‘š
𝑄 which is not a permutation. Then there exists affine permutations 𝐡, 𝐢 : F𝑄 → F𝑄 such
that 𝐴(X) = 𝐡(𝐢(X)1 , . . . , 𝐢(X)π‘Ÿ , 0, . . . , 0) where π‘Ÿ < π‘š is the dimension of the image of 𝐴.
By closure under affine permutations, it follows 𝑓 ∘𝐡 ∈ π’ž. Applying the assertion above π‘š−π‘Ÿ
times we have that 𝑓 ′ (X) , (𝑓 ∘ 𝐡)(𝑋1 , . . . , π‘‹π‘Ÿ , 0, . . . , 0) is also in π’ž. Finally 𝑓 ∘ 𝐴 = 𝑓 ′ ∘ 𝐢
is also in π’ž. So we turn to proving that for every 𝑓 ∈ π’ž the function 𝑓˜ = 𝑓 (𝑋1 , . . . , π‘‹π‘š−1 , 0)
is also in π’ž.
d
d
˜
˜(X) = ∑οΈ€
𝑐
X
.
Notice
𝑓
d
d|π‘‘π‘š =0 𝑐d X . Writing 𝑓 = 𝑓 + 𝑓1 , we use
d
∑οΈ€
Lemma 4.2.2 to split 𝑓 till we express it as a sum of basic functions 𝑓 = 𝑁
𝑖=1 𝑏𝑖 , where
Let 𝑓 (X) =
∑οΈ€
each 𝑏𝑖 is a basic function in π’ž. Note that for every 𝑏𝑖 , we have supp(𝑏𝑖 ) ⊆ supp(𝑓˜) or
supp(𝑏𝑖 ) ⊆ supp(𝑓1 ) (since the trace preserves π‘‘π‘š = 0). By reordering the 𝑏𝑖 ’s assume the
∑οΈ€
first 𝑀 𝑏𝑖 ’s have their support in the support of 𝑓˜. Then we have 𝑓˜ = 𝑀
𝑖=1 𝑏𝑖 ∈ π’ž.
We thus turn to the proof of Lemma 4.2.2. We prove the lemma in a sequence of cases,
based on the kind of monomials that 𝑓 has in its support.
We say that d and e are equivalent (modulo π‘ž), denoted d ≡π‘ž e if there exists a 𝑗 such
that for every 𝑖, 𝑑𝑖 = π‘ž 𝑗 𝑒𝑖 mod* 𝑄. The following proposition is immediate from previous
works (see, for example, [BGM+ 11b]). We include a proof for completeness.
41
Proposition 4.2.3. If every pair d, e in the support of 𝑓 : Fπ‘š
𝑄 → Fπ‘ž are equivalent, then 𝑓
is a basic function.
Proof. We first note that since the Tr : F𝑄 → Fπ‘ž is a (𝑄/π‘ž)-to-one function, we have
in particular that for every 𝛽 ∈ Fπ‘ž there is an 𝛼 ∈ F𝑄 such that Tr(𝛼) = 𝛽. As an
immediate consequence we have that every function 𝑓 : Fπ‘š
𝑄 → Fπ‘ž can be expressed Tr βˆ˜π‘”
where 𝑔 : Fπ‘š
𝑄 → F𝑄 . Finally we note that we can view 𝑔 as an element of F𝑄 [X], to conclude
that 𝑓 = Tr βˆ˜π‘” for some polynomial 𝑔.
Now fix 𝑓 : Fπ‘š
𝑄 → Fπ‘ž all of whose monomials are equivalent. By the above we can express
𝑓 = Tr βˆ˜π‘” for some polynomial 𝑔. By inspection we can conclude that all monomials in the
support of 𝑔 are equivalent to the monomials in the support of 𝑓 . Finally, using the fact
*
that Tr(𝛼Xd ) = Tr(π›Όπ‘ž Xπ‘žd mod
𝑄
) we can assume that 𝑔 is supported on a single monomial
and so 𝑓 = Tr(πœ†Xd ) for some πœ† ∈ F𝑄 .
So it suffices to show that every function that contains non-equivalent degrees in its
support can be split. We first prove that functions with “non-weakly-equivalent” monomials
can be split.
We say that d and e are weakly equivalent if there exists a 𝑗 such that for every 𝑖,
𝑑𝑖 = π‘ž 𝑗 𝑒𝑖 (mod 𝑄 − 1).
Lemma 4.2.4. If β„± ⊆ {Fπ‘š
𝑄 → Fπ‘ž } is an Fπ‘ž -linear code invariant under affine permutations
and 𝑓 ∈ π’ž contains a pair of non-weakly equivalent monomials in its support, then 𝑓 can be
split.
Proof. Let d and e be two non weakly-equivalent monomials in the support of 𝑓 . Fix 𝑗 and
∑οΈ€
∏οΈ€
𝑗𝑑
𝑖
consider the function 𝑓𝑗 (X) = a∈(F* )π‘š π‘Ž−π‘ž
𝑓 (π‘Ž1 𝑋1 , . . . , π‘Žπ‘š π‘‹π‘š ). We claim that (1) the
𝑖
𝑄
support of 𝑓𝑗 is a subset of the support of 𝑓 , (2) π‘ž 𝑗 d is in the support of 𝑓𝑗 , (3) d is in the
support of 𝑓𝑗 only if for every 𝑖 𝑓𝑖 = π‘ž 𝑗 𝑑𝑖 (mod 𝑄 − 1) and in particular (4) e is not in the
support of 𝑓𝑗 .
Now let 𝑏 = 𝑏(d) be the smallest positive integer such that π‘ž 𝑏 𝑑𝑖 = 𝑑𝑖 mod* 𝑄 for every 𝑖.
∑︀𝑏−1
Now consider the function 𝑔 =
𝑗=0 𝑓𝑗 . We have that 𝑔 ∈ π’ž since it is an Fπ‘ž -linear
42
combination of linear transforms of functions in π’ž. By the claims about the 𝑓𝑗 ’s we also
have that d is in the support of 𝑔, the support of 𝑔 is contained in the support of 𝑓 and e is
not in the support of 𝑓 . Expressing 𝑓 = 𝑔 + (𝑓 − 𝑔) we now have that 𝑓 can be split.
The remaining cases are those where some of coordinates of d are zero or 𝑄 − 1 for every
d in the support of 𝑓 . We deal with a special case of such functions next.
Lemma 4.2.5. Let π’ž be a linear affine-invariant code. Let 𝑓 ∈ π’ž be given by 𝑓 (X, Y) =
Tr(Yd 𝑝(X)) where every variable in 𝑝(X) has degree in {0, 𝑄 − 1} in every monomial, and
d is arbitrary. Further, let degree of 𝑝(X) be π‘Ž(𝑄 − 1). Then for every 0 ≤ 𝑏 ≤ π‘Ž and for
every πœ† ∈ F𝑄 , the function (𝑋1 · · · 𝑋𝑏 )𝑄−1 Tr(πœ†Yd ) ∈ π’ž.
Note that in particular the lemma above implies that such 𝑓 ’s can be split into basic
functions.
Proof. We prove the lemma by a triple induction, first on π‘Ž, then on 𝑏, and then on the
number of monomials in 𝑝. The base case is π‘Ž = 0 and that is trivial. So we consider general
π‘Ž > 0.
First we consider the case 𝑏 < π‘Ž. Assume without loss of generality that the monomial
(𝑋1 · · · π‘‹π‘Ž )𝑄−1 is in the support of 𝑝 and write 𝑝 = 𝑝0 +𝑋1𝑄−1 𝑝1 where 𝑝0 , 𝑝1 do not depend on
𝑋1 . Note that 𝑝1 ΜΈ= 0 and deg(𝑝1 ) = (π‘Ž − 1)(𝑄 − 1). We will prove that − Tr(Yd 𝑝1 (X)) ∈ π’ž
∑οΈ€
and this will enable us to apply the inductive hypothesis to 𝑝1 . Let 𝑔(X, Y) = 𝛽∈F𝑄 𝑓 (𝑋1 +
𝛽, 𝑋2 , . . . , π‘‹π‘š , Y). By construction 𝑔 ∈ π’ž. By linearity of the Trace we have
βŽ›
βŽ›
𝑔 = Tr ⎝Yd ⎝
⎞⎞
∑︁
𝑝0 + (𝑋1 + 𝛽)𝑄−1 𝑝1 ⎠⎠ = Tr(Yd (−𝑝1 (X))),
𝛽∈F𝑄
where the second equality follows from the fact that
∑οΈ€
𝛽∈F𝑄 (𝑧
+ 𝛽)𝑄−1 = −1. Thus we can
now use induction to claim (𝑋1 . . . 𝑋𝑏 )𝑄−1 Tr(πœ†Yd ) ∈ π’ž.
Finally we consider the case 𝑏 = π‘Ž. Now note that since the case 𝑏 < π‘Ž is known, we can
assume without loss of generality that 𝑝 is homogeneous (else we can subtract off the lower degree terms). Now if π‘Ž = π‘š there is nothing to be proved since 𝑝 is just a single monomial. So
43
assume π‘Ž < π‘š. Also if 𝑝 has only one monomial then there is nothing to be proved, so assume
𝑝 has at least two monomials. In particular assume 𝑝 is supported on some monomial that
depends on 𝑋1 and some monomial that does not depend on 𝑋1 . Furthermore, assume without loss of generality that a monomial depending on 𝑋1 does not depend on 𝑋2 . Write 𝑝 =
𝑋1𝑄−1 𝑝1 + 𝑋2𝑄−1 𝑝2 + (𝑋1 𝑋2 )𝑄−1 𝑝3 + 𝑝4 where the 𝑝𝑖 ’s don’t depend on 𝑋1 or 𝑋2 . By assumption on the monomials of 𝑝 we have that 𝑝1 ΜΈ= 0 and at least one of 𝑝2 , 𝑝3 , 𝑝4 ΜΈ= 0. Now consider
the affine transform 𝐴 that sends 𝑋1 to 𝑋1 + 𝑋2 and preserves all other 𝑋𝑖 ’s. We have 𝑔 =
(︁
)︁
𝑓 ∘ 𝐴 = Tr Yd (𝑋1𝑄−1 𝑝1 + 𝑋2𝑄−1 (𝑝1 + 𝑝2 ) + (𝑋1 𝑋2 )𝑄−1 𝑝3 + 𝑝4 + π‘Ÿ) where the 𝑋1 -degree of
∑οΈ€
every monomial in π‘Ÿ is in [𝑄 − 2]. Now consider 𝑔 ′ (x, y) = 𝛼∈F* 𝑔(𝛼π‘₯1 , π‘₯2 , . . . , π‘₯π‘š , y). The
𝑄
′
terms of π‘Ÿ vanish in 𝑔 leaving
(︁
(︁
)︁)︁
𝑔 ′ = −(𝑓 ∘ 𝐴 − π‘Ÿ) = Tr Yd −𝑋1𝑄−1 𝑝1 − 𝑋2𝑄−1 (𝑝1 + 𝑝2 ) − (𝑋1 𝑋2 )𝑄−1 𝑝3 − 𝑝4 .
Finally we consider the function 𝑔˜ = 𝑓 + 𝑔 ′ = Tr(Yd (−𝑋2𝑄−1 𝑝1 )) which is a function in π’ž
of degree π‘Ž(𝑄 − 1) supported on a smaller number of monomials than 𝑓 , so by applying the
inductive hypothesis to 𝑔˜ we have that π’ž contains the monomial (𝑋1 · · · π‘‹π‘Ž )𝑄−1 .
The following lemma converts the above into the final piece needed to prove Lemma 4.2.2.
Lemma 4.2.6. If β„± ⊆ {Fπ‘š
𝑄 → Fπ‘ž } is an Fπ‘ž -linear code invariant under affine permutations
and all monomials in 𝑓 ∈ π’ž are weakly equivalent, then 𝑓 can be split.
Proof. First we describe the structure of a function 𝑓 : Fπ‘š
𝑄 → Fπ‘ž that consists only of weakly
equivalent monomials. First we note that the π‘š variables can be separated into those in
which every monomial has degree in [𝑄 − 2] and those in which every monomial has degree
in {0, 𝑄 − 1} (since every monomial is weakly equivalent). Let us denote by X the variables
in which the monomials of 𝑓 have degree in {0, 𝑄 − 1} and Y be the remaining variables.
Now consider some monomial of the form 𝑀 = 𝑐Xe Yd in 𝑓 . Since 𝑓 maps to Fπ‘ž we must
𝑗
𝑗
have that the coefficient of (Xe Yd )π‘ž is π‘π‘ž . Furthermore, we have every other monomial 𝑀 ′
𝑗
′
in the support of 𝑓 is of the form 𝑐′ Yπ‘ž d Xπ‘₯e . Thus 𝑓 can be written as Tr(Yd 𝑝(X)) where
𝑄−1
𝑝(𝑋1 , . . . , π‘‹π‘š ) = 𝑝˜(𝑋1𝑄−1 , . . . , π‘‹π‘š
). But, by Lemma 4.2.5, such an 𝑓 can be split.
44
Proof of Lemma 4.2.2. If 𝑓 contains a pair of non-weakly equivalent monomials then 𝑓 can
be split by Lemma 4.2.4. If not, then 𝑓 is either basic or, by Lemma 4.2.6 is can be split.
We also prove an easy consequence of Lemma 4.2.2.
d
Lemma 4.2.7. Let π’ž ⊆ {Fπ‘š
𝑄 → Fπ‘ž } be affine invariant. If d ∈ Deg(π’ž), then Tr(πœ†X ) ∈ π’ž
for all πœ† ∈ F𝑄 .
Proof. We first claim that Lemma 4.2.2 implies that there exists 𝛽 ∈ F𝑄 such that Tr(𝛽Xd )
is a non-zero function in π’ž. To verify this, consider a “minimal” function (supported on
fewest monomials) 𝑓 ∈ π’ž with d ∈ supp(𝑓 ). Since 𝑓 can’t be split in π’ž (by minimality), by
Lemma 4.2.2 𝑓 must be basic and so equals (by definition of being basic) Tr(𝛽Xd ).
Now let 𝑏 = 𝑏(d) be the smallest positive integer such that π‘ž 𝑏 d mod* 𝑄 = d. If 𝑄 = π‘ž 𝑛 ,
note that 𝑏 divides 𝑛 and so one can write Tr : F𝑄 → Fπ‘ž as Tr1 ∘ Tr2 where Tr1 : Fπ‘žπ‘ → Fπ‘ž
is the function Tr1 (𝑧) = 𝑧 + 𝑧 π‘ž + · · · + 𝑧 π‘ž
𝑏
𝑏−1
and Tr2 : F𝑄 → Fπ‘žπ‘ is the function Tr2 (𝑧) =
𝑏
𝑧 + 𝑧 π‘ž + · · · + 𝑧 𝑄/π‘ž . (Both Tr1 and Tr2 are trace functions mapping the domain to the
range.) It follows that Tr(𝛽Xd ) = Tr1 (Tr2 (𝛽)Xd ).
We first claim that Tr1 (𝜏 Xd ) ∈ π’ž for every 𝜏 ∈ Fπ‘žπ‘ . Let 𝑆 = {
∑οΈ€
𝛼∈(F*𝑄 )π‘š
π‘Žπ›Ό ·π›Όd | π‘Žπ›Ό ∈ Fπ‘ž }.
We note that by linearity and affine-invariance of π’ž, we have that Tr1 (Tr2 (𝛽) · πœ‚Xd ) ∈ π’ž
for every πœ‚ ∈ 𝑆. By definition 𝑆 is closed under addition and multiplication and so is a
𝑏
subfield of F𝑄 . In fact, since every πœ‚ ∈ 𝑆 satisfies πœ‚ π‘ž = πœ‚ (which follows from the fact
𝑏
that 𝛼d = π›Όπ‘ž d ), we have that 𝑆 ⊆ Fπ‘žπ‘ . It remains to show 𝑆 = Fπ‘žπ‘ . Suppose it is a
strict subfield of size π‘ž 𝑐 for 𝑐 < 𝑏. Consider 𝛾 𝑑𝑖 for 𝛾 ∈ F𝑄 and 𝑖 ∈ [π‘š]. Since 𝛾 𝑑𝑖 ∈ 𝑆,
𝑐
we have that 𝛾 𝑑𝑖 π‘ž = 𝛾 𝑑𝑖 for every 𝛾 ∈ F𝑄 and so we get π‘‹π‘–π‘ž
conclude that Xπ‘ž
𝑐d
𝑐𝑑
𝑖
= 𝑋𝑖 (mod 𝑋𝑖𝑄 − 𝑋𝑖 ). We
= Xd (mod X𝑄 − X) which contradicts the minimality of 𝑏 = 𝑏(d). We
conclude that 𝑆 = Fπ‘žπ‘ . Since Tr2 (𝛽) ∈ F*π‘žπ‘ , we conclude that the set of coefficients 𝜏 such
that Tr1 (𝜏 Xd ) ∈ π’ž is all of Fπ‘žπ‘ as desired.
Finally consider any πœ† ∈ F𝑄 . since Tr2 (πœ†) ∈ Fπ‘žπ‘ , we have that Tr1 (Tr2 (πœ†)Xd ) ∈ π’ž (from
the previous paragraph), and so Tr(πœ†Xd ) = Tr1 (Tr2 (πœ†)Xd ) ∈ π’ž
45
46
Chapter 5
Lifting Codes
5.1
The Lift Operator
First defined and used in [BSMSS11] to prove the existence of certain codes that are not
locally testable, the lift operator takes short codes and creates longer codes from them. The
original lift operator took codes defined over the domain Fπ‘ž and “lifted” them to codes
defined over the domain Fπ‘žπ‘š . We generalize the definition, allowing one to lift codes over Fπ‘‘π‘ž
to codes over Fπ‘š
π‘ž , for any π‘š ≥ 𝑑 (in particular, 𝑑 does not need to divide π‘š).
Definition 5.1.1 (Lift). Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be a linear affine-invariant code. Let π‘š ≥ 𝑑 be
an integer. The π‘š-dimensional lift of π’ž, denoted by π’ž π‘‘β†—π‘š , is the code
βƒ’
{οΈ€
}οΈ€
βƒ’ 𝑓 |𝐴 ∈ π’ž for every 𝑑-dimensional affine subspace 𝐴 ⊆ Fπ‘š
π’ž π‘‘β†—π‘š , 𝑓 : Fπ‘š
.
π‘ž → Fπ‘ž
π‘ž
Proposition 5.1.2. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant, and let π‘š ≥ 𝑑. If
𝑓 ∈ π’ž π‘‘β†—π‘š and 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž is an affine map, then 𝑓 ∘ 𝐴 ∈ π’ž.
Proof. By Proposition B.1.1, there exists an invertible affine map 𝐴′′ : Fπ‘‘π‘ž → Fπ‘š
π‘ž and a linear
map 𝐴′ : Fπ‘‘π‘ž → Fπ‘‘π‘ž such that 𝐴 = 𝐴′′ ∘ 𝐴′ . By the definition of lift, 𝑔 , 𝑓 ∘ 𝐴′′ ∈ π’ž. Since π’œ
47
is affine-invariant, it follows from Theorem 4.2.1 that 𝑔 ∘ 𝐴′ ∈ π’ž. Therefore,
𝑓 ∘ 𝐴 = 𝑓 ∘ (𝐴′′ ∘ 𝐴′ ) = (𝑓 ∘ 𝐴′′ ) ∘ 𝐴′ = 𝑔 ∘ 𝐴′ ∈ π’ž.
5.2
Algebraic and Combinatorial Properties
We begin by exploring the algebraic and combinatorial (as opposed to algorithmic) properties
of codes lifted from linear affine-invariant codes. First, we show that the lifting operator is
a natural algebraic operation, preserving linearity and affine-invariance and composing well.
We then present the degree set of a lifted code in terms of the degree set of the base code.
Finally, we conclude by showing that lifting preserves distance.
5.2.1
Algebraic Properties
Proposition 5.2.1. Let 𝑑 ≤ π‘š be integers. If π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } is linear affine-invariant,
then π’ž π‘‘β†—π‘š is linear affine-invariant.
Proof. First, we show that π’ž π‘‘β†—π‘š is linear. Let 𝑓, 𝑔 ∈ π’ž π‘‘β†—π‘š and let 𝛼 ∈ Fπ‘ž . If 𝐴 ⊆ Fπ‘š
π‘ž is a
𝑑-dimensional affine subspace, then, since π’ž is linear,
(𝛼𝑓 + 𝑔)|𝐴 = 𝛼 · (𝑓 |𝐴 ) + (𝑔|𝐴 ) ∈ π’ž
Next, we show that π’ž π‘‘β†—π‘š is affine-invariant. Let 𝑓 (X) ∈ π’ž π‘‘β†—π‘š , and let M ∈ Fπ‘š×π‘š
and
π‘ž
π‘š×𝑑
c ∈ Fπ‘š
and b ∈ Fπ‘š
π‘ž . Let 𝑔(X) = 𝑓 (MX + c). Let Y = (π‘Œ1 , . . . , π‘Œπ‘‘ ), let A ∈ Fπ‘ž
π‘ž . Then
𝑔(AX + b) = 𝑓 (M(AX + b) + c) = 𝑓 ((MA)X + (Mb + c)) ∈ π’ž.
Since A, b were arbitrary, this implies that 𝑔 ∈ π’ž π‘‘β†—π‘š . Since M, c were arbitrary, this implies
that π’ž π‘‘β†—π‘š is affine-invariant.
48
Proposition 5.2.2 (Composition of lift). Let 𝑑 ≤ π‘š ≤ 𝑛 be integers. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž }
be linear affine-invariant. Then
(οΈ€
)οΈ€π‘šβ†—π‘›
π’ž 𝑑↗𝑛 = π’ž π‘‘β†—π‘š
.
Proof. This follows immediately from the fact that choosing a 𝑑-dimensional affine subspace
𝐴 ⊆ Fπ‘›π‘ž is equivalent to first choosing an π‘š-dimensional affine subspace 𝐡 ⊆ Fπ‘›π‘ž and then
choosing a 𝑑-dimensional affine subspace 𝐴 ⊆ 𝐡.
Proposition 5.2.3 (Degree set of lift). Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant with
degree set Deg(π’ž). If π‘š ≥ 𝑑 is an integer, then the lift π’ž π‘‘β†—π‘š has degree set
Deg(π’ž π‘‘β†—π‘š ) = {d ∈ Jπ‘žKπ‘š | E ≤𝑝 d =⇒ (β€–E*1 β€–, . . . , β€–E*𝑑 β€–) mod* π‘ž ∈ Deg(π’ž)}
where E has rows [π‘š] and columns J𝑑 + 1K.
Proof. Let 𝑓 ∈ π’ž π‘‘β†—π‘š . Let 𝐷 = Deg(π’ž π‘‘β†—π‘š ). Let X = (𝑋1 , . . . , π‘‹π‘š ) and let Y = (π‘Œ1 , . . . , π‘Œπ‘‘ ).
∑οΈ€
d
π‘š×𝑑
Write 𝑓 (X) =
and b ∈ Fπ‘š
π‘ž , it follows from
d∈𝐷 𝑓d · X . For any matrix A ∈ Fπ‘ž
Proposition A.1.4 that
(οΈƒ
)οΈƒ 𝑑
𝑑
π‘š
∏︁
∑︁ (οΈ‚ d )οΈ‚ ∏︁
∏︁ β€–E β€–
𝑒
𝑏𝑒𝑖 𝑖0
π‘Žπ‘–π‘—π‘–π‘— ·
π‘Œπ‘— *𝑗
𝑓 (AY + b) =
𝑓d ·
E
𝑗=1
𝑗=1
𝑖=1
d∈𝐷
E≤ d
∑︁
𝑝
Since 𝑓 (AY + b) ∈ π’ž for any A, b, and π’ž is affine-invariant, it follows by Proposition 4.1.3
that for every d ∈ 𝐷 and E ≤𝑝 d, with columns J𝑑 + 1K, it holds that the monomial
∏︀𝑑
β€–E*𝑗 β€–
∈ π’ž, hence (β€–E*1 β€–, . . . , β€–E*𝑑 β€–) mod* π‘ž ∈ Deg(π’ž). Conversely, if d satisfies that
𝑗=1 π‘Œπ‘—
for every E ≤𝑝 d with columns J𝑑 + 1K that (β€–E*1 β€–, . . . , β€–E*𝑑 β€–) mod* π‘ž ∈ Deg(π’ž), then it is
easy to see that d ∈ 𝐷 by considering 𝑓 (X) = Xd .
49
5.2.2
Distance of Lifted Codes
Proposition 5.2.4 (Distance of lift). Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant, and let
π‘š ≥ 𝑑. The following hold:
1. 𝛿(π’ž π‘‘β†—π‘š ) ≤ 𝛿(π’ž);
2. 𝛿(π’ž π‘‘β†—π‘š ) ≥ 𝛿(π’ž) − π‘ž −𝑑 ;
3. if π‘ž ∈ {2, 3} and 𝛿(π’ž) > π‘ž −𝑑 , then 𝛿(π’ž π‘‘β†—π‘š ) = 𝛿(π’ž).
We prove several lemmas, which in turn will help us prove Proposition 5.2.4. The first
lemma proves Proposition 5.2.4 (1).
Lemma 5.2.5. If π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } is linear affine-invariant, and π‘š ≥ 𝑑, then 𝛿(π’ž π‘‘β†—π‘š ) ≤ 𝛿(π’ž).
Proof. The case π‘š = 𝑑 is trivial, so assume π‘š > 𝑑. By Proposition 5.2.2 and induction, it
suffices to consider the case π‘š = 𝑑 + 1. Let 𝑓 ∈ π’ž and let 𝛿 , 𝛿(𝑓, 0). Let X = (𝑋1 , . . . , 𝑋𝑑 ).
→ Fπ‘ž defined by 𝑔(X, 𝑋𝑑+1 ) = 𝑓 (X). Clearly, we have
Consider the function 𝑔 : F𝑑+1
π‘ž
be an affine map. Let
𝛿(𝑔, 0) = 𝛿. We claim that 𝑔 ∈ π’ž 𝑑↗(𝑑+1) . Let 𝐴 : Fπ‘‘π‘ž → F𝑑+1
π‘ž
𝐴′ : Fπ‘‘π‘ž → Fπ‘‘π‘ž be the affine map given by the projection of 𝐴 onto its first 𝑑 coordinates, i.e.
𝐴′ (X) = (𝐴(X)1 , . . . , 𝐴(X)𝑑 ). Since 𝑓 ∈ π’ž and π’ž is affine-invariant, by Theorem 4.2.1 we
have 𝑓 ∘ 𝐴′ ∈ π’ž. Therefore, 𝑔 ∘ 𝐴 = 𝑓 ∘ 𝐴′ ∈ π’ž. Since 𝐴 was arbitrary, this shows that
𝑔 ∈ π’ž 𝑑↗(𝑑+1) .
The next lemma proves Proposition 5.2.4 (2), and uses a simple probabilistic argument.
Lemma 5.2.6. If π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } is linear affine-invariant, and π‘š ≥ 𝑑, then 𝛿(π’ž π‘‘β†—π‘š ) ≥
𝛿(π’ž) − π‘ž −𝑑 .
𝑑
π‘š
Proof. Let 𝑓, 𝑔 ∈ π’ž π‘‘β†—π‘š be distinct. Let x ∈ Fπ‘š
π‘ž such that 𝑓 (x) ΜΈ= 𝑔(x). Let 𝐴 : Fπ‘ž → Fπ‘ž be
a random affine map such that 𝐴(0) = x, so that 𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴 ∈ π’ž (by Proposition 5.1.2) and
are distinct. Then
𝛿(π’ž) ≤ E𝐴 [𝛿 (𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴)]
(Proposition B.1.3) ≤ 𝛿(𝑓, 𝑔) + π‘ž −𝑑 .
50
(5.1)
(5.2)
The next two lemmas will help in proving Proposition 5.2.4, for the cases π‘ž = 2 and π‘ž = 3
respectively.
Lemma 5.2.7. For all π‘š ≥ 2, if 𝛿 >
1
2π‘š−1
and 𝑓 : Fπ‘š
2 → F2 such that 0 < 𝛿(𝑓, 0) < 𝛿, then
there exists a hyperplane 𝐻 ⊂ Fπ‘š
2 such that 0 < 𝛿(𝑓 |𝐻 , 0) < 𝛿.
Proof. We proceed by induction on π‘š. The base case π‘š = 2 is straightforward to verify.
Now suppose π‘š > 2 and our assertion holds for π‘š − 1. Let 𝐻0 , 𝐻1 be the affine subspaces
given by π‘‹π‘š = 0 and π‘‹π‘š = 1, respectively. For 𝑖 ∈ {0, 1}, let 𝛿𝑖 , 𝛿(𝑓 |𝐻𝑖 , 0). Note that
𝛿 > 𝛿(𝑓, 0) = (𝛿0 + 𝛿1 )/2. If both 𝛿0 , 𝛿1 > 0, then by averaging we have 0 < 𝛿𝑖 < 𝛿, for some
𝑖 ∈ {0, 1}, and so 𝐻 = 𝐻𝑖 does the job. Otherwise, suppose without loss of generality that
𝛿1 = 0. Note that 0 < 𝛿0 < 2𝛿 and 2𝛿 >
1
.
2π‘š−2
Thus, by the induction hypothesis, there
exists an (π‘š − 2)-dimensional affine subspace 𝐻0′ ⊆ 𝐻0 such that 0 < 𝛿(𝑓 |𝐻0′ , 0) < 2𝛿. Let
𝐻1′ , 𝐻0′ + eπ‘š , where eπ‘š is the π‘š-th standard basis vector. Note that 𝛿(𝑓 |𝐻1′ , 0) = 0. Let
′
𝐻 = 𝐻0′ ∪ 𝐻1′ . Then 𝐻 is an (π‘š − 1)-dimensional subspace of Fπ‘š
2 (spanned by 𝐻0 and eπ‘š )
)οΈ€
(οΈ€
such that 0 < 𝛿(𝑓 |𝐻 , 0) = 𝛿(𝑓 |𝐻0′ , 0) + 𝛿(𝑓 |𝐻1′ , 0) /2 < 𝛿.
Lemma 5.2.8. For all π‘š ≥ 2, if 𝑓 : Fπ‘š
3 → F3 and 𝛿(𝑓, 0) ≥
1
,
3π‘š−1
then there exists a
hyperplane 𝐻 ⊆ Fπ‘š
3 such that 0 < 𝛿(𝑓 |𝐻 , 0) ≤ 𝛿(𝑓, 0).
Proof. Let 𝛿 , 𝛿(𝑓, 0). We proceed by induction on π‘š.For the base case, π‘š = 2, 𝛿 ≥ 13 .
Suppose 𝑓 : F23 → F3 and for each 𝑖 ∈ F3 , let 𝑓𝑖 be the restriction of 𝑓 to the hyperplane
defined by 𝑋2 = 𝑖. If 𝑓𝑖 ΜΈ= 0 for every 𝑖 ∈ F3 , then by averaging there is some 𝑖 ∈ F3 for
which 0 < 𝛿(𝑓𝑖 , 0) ≤ 𝛿. Otherwise, without loss of generality, suppose 𝑓2 = 0. Further,
without loss of generality, suppose 𝑓0 ΜΈ= 0 and 𝑓 (0, 0) ΜΈ= 0. Now, if 𝛿 ≥ 32 , then the line
𝐻 = {(π‘₯, 𝑦) ∈ F32 | π‘₯ = 0} does the job, since 0 < 𝛿(𝑓 |𝐻 , 0) ≤
2
3
≤ 𝛿. If 𝛿 < 23 , then there
must exist some π‘Ž, 𝑏 ∈ F3 and 𝑐 ∈ {0, 1} such that 𝑓 (π‘Ž, 𝑐) ΜΈ= 0 and 𝑓 (𝑏, 1 − 𝑐) = 0. Then the
line 𝐻 = {(π‘Ž, 𝑐), (𝑏, 1 − 𝑐), (2𝑏 − π‘Ž, 2)} does the job, since 0 < 𝛿(𝑓 |𝐻 , 0) =
1
3
≤ 𝛿.
Now suppose π‘š > 2 and the assertion holds for π‘š − 1. For 𝑖 ∈ F3 , let 𝐻𝑖 be the
hyperplane cut out by π‘‹π‘š = 𝑖 and let 𝛿𝑖 , 𝛿(𝑓 |𝐻𝑖 , 0) for each 𝑖 ∈ F3 . Then 𝛿0 + 𝛿1 + 𝛿3 = 3𝛿.
51
If 𝛿𝑖 > 0 for all 𝑖 ∈ F3 , then by simple averaging, for some 𝑖 ∈ F3 , we have 0 < 𝛿𝑖 ≤ 𝛿, so
assume without loss of generality that 𝛿2 = 0 and 𝛿0 ≥ 𝛿1 .
First, suppose 𝛿0 ≥
1
.
3π‘š−2
Then, by the inductive hypothesis, there exists an (π‘š − 2)-
dimensional affine subspace 𝐻 (0) ⊂ 𝐻0 such that 0 < 𝛿(𝑓 |𝐻 , 0) ≤ 𝛿0 . Suppose 𝐻 (0) is defined
∑οΈ€
π‘š+1
. For
by the linear equations π‘š
𝑖=1 π‘Žπ‘– 𝑋𝑖 − π‘Ž0 = 0 and π‘‹π‘š = 0 for some (π‘Ž0 , . . . , π‘Žπ‘š ) ∈ Fπ‘ž
∑οΈ€
(𝑖)
each 𝑖, 𝑗 ∈ F3 , let 𝐻𝑗 ⊂ 𝐻𝑖 be the affine subspace defined by π‘š
𝑖=1 π‘Žπ‘– 𝑋𝑖 −π‘Ž0 = 𝑗 and π‘‹π‘š = 𝑖.
(︁
)︁
(1)
(2)
By averaging, for some 𝑗 ∈ F3 , we have 𝛿 𝑓 |𝐻 (1) , 0 ≤ 𝛿1 . Take 𝐻 , 𝐻 (0) ∪ 𝐻𝑗 ∪ 𝐻2𝑖 .
𝑗
Then
0 < 𝛿(𝑓 |𝐻 , 0) =
Now, suppose 𝛿0 <
(οΈ€
)οΈ€
(οΈ€
)οΈ€
(οΈ€
)οΈ€
𝛿 𝑓 |𝐻 (0) , 0 + 𝛿 𝑓 |𝐻 (1) , 0 + 𝛿 𝑓 |𝐻 (2) , 0
1
,
3π‘š−2
𝑗
2𝑗
3
so 𝛿0 , 𝛿1 ≤
2
.
3π‘š−1
subspace 𝐻 (0) ⊂ 𝐻0 such that 𝛿(𝑓 |𝐻 (0) , 0) =
≤
𝛿0 + 𝛿1
= 𝛿.
3
There exists an (π‘š − 2)-dimensional affine
1
.
3π‘š−1
To see this, let a, b ∈ 𝐻0 such that
𝑓 (a), 𝑓 (b) ΜΈ= 0, and suppose π‘Žπ‘˜ ΜΈ= π‘π‘˜ , for some π‘˜ ∈ [π‘š]. Then take 𝐻 (0) to be the subspace
(𝑖)
defined by π‘‹π‘˜ = π‘Žπ‘˜ and π‘‹π‘š = 0. For 𝑖, 𝑗 ∈ F3 , let 𝐻𝑗 be the (π‘š − 2)-dimensional subspace
defined by π‘‹π‘˜ = π‘Žπ‘˜ + 𝑗 and π‘‹π‘š = 𝑖. Since 𝛿1 ≤
Then, taking 𝐻 , 𝐻
(0)
∪
(1)
𝐻𝑗
0 < 𝛿(𝑓 |𝐻 , 0) =
∪
(2)
𝐻2𝑗 ,
2
,
3π‘š−2
there is 𝑗 ∈ F3 such that 𝑓 |𝐻 (1) = 0.
𝑗
we have
)οΈ€
(οΈ€
)οΈ€
(οΈ€
)οΈ€
(οΈ€
𝛿 𝑓 |𝐻 (0) , 0 + 𝛿 𝑓 |𝐻 (1) , 0 + 𝛿 𝑓 |𝐻 (2) , 0
𝑗
3
2𝑗
=
1
< 𝛿.
3π‘š
Now, we are ready to prove Proposition 5.2.4.
Proof of Proposition 5.2.4. Parts 1 and 2 follow immediately from Lemmas 5.2.5 and 5.2.5,
respectively, so we proceed with proving part 3. In light of part 1, it suffices to show that
𝛿(π’ž π‘‘β†—π‘š ) ≥ 𝛿(π’ž).
We start with the π‘ž = 2 case. We proceed by induction on π‘š − 𝑑. Indeed, the inductive
(οΈ€
)οΈ€π‘š−1β†—π‘š
step is straightforward since, by Proposition 5.2.2, we have π’ž π‘‘β†—π‘š = π’ž π‘‘β†—π‘š−1
, so by
induction 𝛿(π’ž π‘‘β†—π‘š ) ≥ 𝛿(π’ž π‘‘β†—π‘š−1 ) ≥ 𝛿(π’ž). The main case is therefore the base case π‘š = 𝑑 + 1.
Suppose 𝑓 ∈ π’ž 𝑑↗𝑑+1 such that 0 < 𝛿(𝑓, 0) < 𝛿(π’ž). By assumption, 𝛿(π’ž) >
52
1
,
2𝑑
so we
may apply Lemma 5.2.7 to conclude that there exists a hyperplane 𝐻 ⊂ Fπ‘š
2 such that
9 < 𝛿(𝑓 |𝐻 , 0) < 𝛿(π’ž), contradicting the fact that 𝑓 |𝐻 ∈ π’ž (since 𝑓 ∈ π’ž 𝑑↗𝑑+1 ).
Finally, we consider the π‘ž = 3 case. Again, we proceed by induction on π‘š − 𝑑, and again
the main case is the base case π‘š = 𝑑 + 1. Suppose 𝑓 ∈ π’ž 𝑑↗𝑑+1 such that 0 < 𝛿(𝑓, 0) < 𝛿(π’ž).
If 𝛿(𝑓, 0) ≥
1
,
3π‘š−1
then, by Lemma 5.2.8, there exists a hyperplane 𝐻 ⊂ Fπ‘š
π‘ž such that
0 < 𝛿(𝑓 |𝐻 , 0) ≤ 𝛿(𝑓, 0) < 𝛿(π’ž), contradicting the fact that 𝑓 |𝐻 ∈ π’ž. If 𝛿(𝑓, 0) <
1
,
3π‘š−1
then
there are at most two points a, b ∈ Fπ‘š
3 such that 𝑓 (a), 𝑓 (b) ΜΈ= 0. Let 𝑖 ∈ [π‘š] such that
π‘Žπ‘– ΜΈ= 𝑏𝑖 and let 𝐻 be the hyperplane defined by 𝑋𝑖 = π‘Žπ‘– . Then 𝑓 |𝐻 is nonzero only on a, so
0 < 𝛿(𝑓 |𝐻 ) =
5.3
1
3π‘š−1
< 𝛿(π’ž), again contradicting the fact that 𝑓 |𝐻 ∈ π’ž.
Local Decoding and Correcting
In this section, we explore some of the algorithmic decoding properties of lifted codes. In
particular, for codes of distance 𝛿 > 0, we give simple algorithms for locally correcting up to
𝛿/4 and then 𝛿/2 fraction errors. We then show that all linear affine-invariant codes — in
particular, lifted codes — have explicit interpolating sets, thereby showing that lifted codes
are locally decodable as well.
5.3.1
Local Correcting up to 1/4 Distance
The following algorithm is an abstraction of a simple algorithm for locally correcting ReedMuller codes of distance 𝛿 > 0 from 𝛿/4 fraction errors.
Theorem 5.3.1. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant with distance 𝛿 , 𝛿(π’ž), and
(οΈ€
)οΈ€
let π‘š ≥ 𝑑. Then, for every πœ– > 0, the lifted code π’ž π‘‘β†—π‘š is π‘ž 𝑑 , ( 14 − πœ–)𝛿 − π‘ž −𝑑 , 12 − 2πœ– -locally
correctable.
Proof. Let Corrπ’ž be a correction algorithm for π’ž, so that for every 𝑓 : Fπ‘ž → Fπ‘ž that is 𝛿/2close to some 𝑔 ∈ π’ž, Corrπ’ž (𝑓 ) = 𝑔. The following algorithm is a local correction algorithm
that achieves the desired parameters.
53
Local correction algorithm: Oracle access to received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž .
On input x ∈ Fπ‘š
π‘ž :
1. Choose uniform random affine map 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž such that 𝐴(0) = x.
2. Set 𝑓 , π‘Ÿ ∘ 𝐴.
3. Compute 𝑓̂︀ , Corrπ’ž (𝑓 ).
4. Output 𝑓̂︀(0).
Analysis: Let 𝜏 , ( 14 − πœ–)𝛿. Fix a received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž . Let 𝑐 ∈ π’ž be a codeword
such that 𝛿(π‘Ÿ, 𝑐) < 𝜏 − π‘ž −𝑑 . Let 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž be a random affine map such that 𝐴(0) = x.
By Proposition B.1.3, the expected distance is E𝐴 [𝛿 (π‘Ÿ ∘ 𝐴, 𝑐 ∘ 𝐴)] ≤ 𝛿(π‘Ÿ, 𝑐) + π‘ž −𝑑 < 𝜏 , so, by
Markov’s inequality, with probability at least
1
2
+ 2πœ–, we have 𝛿(π‘Ÿ ∘ 𝐴, 𝑐 ∘ 𝐴) < 2𝛿 . For such
𝐴, setting 𝑓 , π‘Ÿ ∘ 𝐴, we have 𝑓̂︀ = Corrπ’ž (𝑓 ) = 𝑐 ∘ 𝐴, hence 𝑓̂︀(0) = 𝑐(𝐴(0)) = 𝑐(x).
5.3.2
Local Correcting up to 1/2 Distance
The following algorithm can locally correct lifted codes up to half the minimum distance.
The basic idea is to decode along multiple lines and weight their opinions based on distance to
the base code. This is a direct translation of the elegant line-weight local decoding algorithm
for matching-vector codes [BET10] to the Reed-Muller code and lifted codes setting.
Theorem 5.3.2. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant with distance 𝛿 , 𝛿(π’ž), and
(οΈ€
)οΈ€
let π‘š ≥ 𝑑. Then, for every πœ–, πœ‚ > 0, the lifted code π’ž π‘‘β†—π‘š is 𝑄, ( 12 − πœ–)𝛿 − π‘ž −𝑑 , πœ‚ -locally
correctable for 𝑄 = 𝑂 (ln(1/πœ‚)/πœ–2 · π‘ž 𝑑 ).
Proof. Let Corrπ’ž be a correction algorithm for π’ž, so that for every 𝑓 : Fπ‘ž → Fπ‘ž that is 𝛿/2close to some 𝑔 ∈ π’ž, Corrπ’ž (𝑓 ) = 𝑔. The following algorithm is a local correction algorithm
that achieves the desired parameters.
54
Local correction algorithm: Oracle access to received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž .
On input x ∈ Fπ‘š
π‘ž :
⌉︁
⌈︁
2
4
1. Let 𝑐 = πœ–2 ln πœ‚ and choose affine maps 𝐴1 , . . . , 𝐴𝑐 ∈ Fπ‘š
π‘ž such that 𝐴𝑖 (0) = x for each
𝑖 ∈ [𝑐] independently and uniformly at random.
2. For each 𝑖 ∈ [𝑐]:
(a) Set π‘Ÿπ‘– , π‘Ÿ ∘ 𝐴𝑖 .
(b) Compute 𝑠𝑖 , Corrπ’ž (π‘Ÿπ‘– ) and 𝛿𝑖 , 𝛿(π‘Ÿπ‘– , 𝑠𝑖 ).
(︁
(c) Assign the value 𝑠𝑖 (0) a weight π‘Šπ‘– , max 1 −
3. Set π‘Š ,
∑︀𝑐
𝑖=1
π‘Šπ‘– . For every 𝛼 ∈ Fπ‘ž , let 𝑀(𝛼) :=
1
π‘Š
)︁
𝛿𝑖
,0
𝛿/2
.
∑οΈ€
𝑖:𝑠𝑖 (0)=𝛼
π‘Šπ‘– . If there is an 𝛼 ∈ Fπ‘ž
with 𝑀(𝛼) > 21 , output 𝛼, otherwise fail.
Analysis: Let 𝜏 , ( 12 −πœ–)𝛿. Fix a received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž . Let 𝑐 ∈ π’ž be a codeword such
that 𝛿(π‘Ÿ, 𝑐) < 𝜏 − π‘ž −𝑑 . The query complexity follows from the fact that the algorithm queries
𝑂(ln(1/πœ‚)/πœ–2 ) subspaces, each consisting of at most π‘ž 𝑑 points. Fix an input x ∈ Fπ‘š
π‘ž . We wish
to show that, with probability at least 1 − πœ‚, the algorithm outputs 𝑐(x), i.e. 𝑀(𝑐(x)) > 12 .
For each affine map 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž , define the following:
𝜏𝐴 , 𝛿(π‘Ÿ ∘ 𝐴, 𝑐 ∘ 𝐴)
𝑠𝐴 , Corrπ’ž (π‘Ÿ ∘ 𝐴)
𝛿𝐴 , 𝛿(π‘Ÿ ∘ 𝐴, 𝑠𝐴 )
(οΈ‚
)οΈ‚
𝛿𝐴
π‘Šπ΄ , max 1 −
,0
𝛿/2
⎧
βŽͺ
βŽ¨π‘Šπ΄ 𝑠𝐴 = 𝑐 ∘ 𝐴
𝑋𝐴 ,
βŽͺ
⎩0
𝑠𝐴 ΜΈ= 𝑐 ∘ 𝐴.
Let 𝑝 , Pr𝐴 [𝑠𝐴 = 𝑐 ∘ 𝐴]. Note that if 𝑠𝐴 = 𝑐 ∘ 𝐴, then 𝛿𝐴 = 𝜏𝐴 , otherwise 𝛿𝐴 ≥ 𝛿 − 𝜏𝐴 .
Hence, if 𝑠𝐴 = 𝑐 ∘ 𝐴, then π‘Šπ΄ ≥ 1 −
𝜏𝐴
,
𝛿/2
otherwise π‘Šπ΄ ≤
55
𝜏𝐴
𝛿/2
− 1.
Define
𝜏good , E𝐴 [𝜏𝐴 | 𝑠𝐴 = 𝑐 ∘ 𝐴]
𝜏bad , E𝐴 [𝜏𝐴 | 𝑠𝐴 ΜΈ= 𝑐 ∘ 𝐴]
π‘Šgood , E𝐴 [π‘Šπ΄ | 𝑠𝐴 = 𝑐 ∘ 𝐴] ≥ 1 −
π‘Šbad , E𝐴 [π‘Šπ΄ | 𝑠𝐴 ΜΈ= 𝑐 ∘ 𝐴] ≤
𝜏good
𝛿/2
𝜏bad
− 1.
𝛿/2
Observe that
E𝐴 [𝜏𝐴 ] ≤
1 + (𝜏 − 1π‘ž )(π‘ž − 1)
π‘ž
≤𝜏
E𝐴 [𝑋𝐴 ] = 𝑝 · π‘Šgood
E𝐴 [π‘Šπ΄ ] = 𝑝 · π‘Šgood + (1 − 𝑝) · π‘Šbad .
We claim that
𝑝 · π‘Šgood ≥ (1 − 𝑝) · π‘Šbad + 2πœ–.
(5.3)
To see this, we start from
(οΈ‚
)οΈ‚
1
− πœ– 𝛿 = 𝜏 ≥ E𝐴 [𝜏𝐴 ] = 𝑝 · 𝜏good + (1 − 𝑝) · 𝜏bad .
2
Dividing by 𝛿/2 yields
1 − 2πœ– ≥ 𝑝 ·
𝜏good
𝜏bad
+ (1 − 𝑝) ·
.
𝛿/2
𝛿/2
Re-writing 1 − 2πœ– on the left-hand side as 𝑝 + (1 − 𝑝) − 2πœ– and re-arranging, we get
(οΈ‚
)οΈ‚
(οΈ‚
)οΈ‚
𝜏good
𝜏bad
≥ (1 − 𝑝) ·
− 1 + 2πœ–.
𝑝· 1−
𝛿/2
𝛿/2
The left-hand side is bounded from above by 𝑝 · π‘Šgood while the right-hand side is bounded
from below by (1 − 𝑝) · π‘Šbad + 2πœ–, hence (5.3) follows.
Consider the random affine maps 𝐴1 , . . . , 𝐴𝑐 chosen by the algorithm. Note that the 𝑋𝐴
56
are defined such that 𝐴𝑖 contributes weight
probability at least 1 − πœ‚,
𝑋𝐴𝑖
π‘Š
to 𝑀(𝑐(x)), so it suffices to show that, with
∑︀𝑐
𝑋
1
∑︀𝑐𝑖=1 𝐴𝑖 > .
2
𝑖=1 π‘Šπ΄π‘–
Each 𝑋𝐴 , π‘Šπ΄ ∈ [0, 1], so by Proposition 2.2.3,
[οΈƒ βƒ’ 𝑐
βƒ’ 1 ∑︁
βƒ’
Pr βƒ’
𝑋𝐴𝑖 − E𝐴 [𝑋𝐴 ]
⃒𝑐
𝑖=1
[οΈƒ βƒ’ 𝑐
βƒ’ 1 ∑︁
βƒ’
Pr βƒ’
π‘Šπ΄π‘– − E𝐴 [π‘Šπ΄ ]
⃒𝑐
𝑖=1
βƒ’
]οΈƒ
βƒ’
βƒ’
≤ 2 exp(−πœ–2 𝑐/2) ≤ πœ‚/2
βƒ’ > πœ–/2
βƒ’
βƒ’
]οΈƒ
βƒ’
βƒ’
≤ 2 exp(−πœ–2 𝑐/2) ≤ πœ‚/2.
βƒ’ > πœ–/2
βƒ’
Therefore, by a union bound, with probability at least 1 − πœ‚ we have
∑︀𝑐
𝑋
∑︀𝑐𝑖=1 𝐴𝑖
𝑖=1 π‘Šπ΄π‘–
E𝐴 [𝑋𝐴 ] − πœ–/2
E[π‘Šπ΄ ] + πœ–/2
𝑝 · π‘Šgood − πœ–/2
=
𝑝 · π‘Šgood + (1 − 𝑝) · π‘Šbad + πœ–/2
(1 − 𝑝) · π‘Šbad + 3πœ–/2
by (5.3) ≥
2(1 − 𝑝) · π‘Šbad + 5πœ–/2
1
.
>
2
5.3.3
≥
Local Decoding
We show that all linear affine-invariant codes, in particular lifted codes, have explicit interpolating sets, which allows us to immediately translate the local correctability of lifted codes
into local decodability.
Finite field isomorphisms. Let Tr : Fπ‘žπ‘š → Fπ‘ž be the Fπ‘ž -linear trace map 𝑧 ↦→
∑οΈ€π‘ž−1
𝑖=0
𝑖
π‘§π‘ž .
Let 𝛼1 , . . . , π›Όπ‘š ∈ Fπ‘žπ‘š be linearly independent over Fπ‘ž and let πœ‘ : Fπ‘žπ‘š → Fπ‘ž be the map 𝑧 ↦→
(Tr(𝛼1 𝑧), . . . , Tr(π›Όπ‘š 𝑧)). Since Tr is Fπ‘ž -linear, πœ‘ is Fπ‘ž -linear, and in fact πœ‘ is an isomorphism.
Observe that πœ‘ induces a Fπ‘ž -linear isomorphism πœ‘* : {Fπ‘š
π‘ž → Fπ‘ž } → {Fπ‘ž π‘š → Fπ‘ž } defined by
57
πœ‘* (𝑓 ) = 𝑓 ∘ πœ‘.
It is straightforward to verify that if π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž }, and 𝑆 ⊆ Fπ‘ž π‘š is an interpolating
set for πœ‘* (π’ž), then πœ‘(𝑆) is an interpolating set for π’ž.
Theorem 5.3.3. Let π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } be a nontrivial affine-invariant code with dimFπ‘ž (π’ž) =
𝐷. Let πœ” ∈ Fπ‘žπ‘š be a generator, i.e. πœ” has order π‘ž π‘š − 1. Let 𝑆 = {πœ”, πœ” 2 , . . . , πœ” 𝐷 } ⊆ Fπ‘žπ‘š .
Then πœ‘(𝑆) ⊆ Fπ‘š
π‘ž is an interpolating set for π’ž.
*
Proof. The map πœ‘ induces a map πœ‘* : {Fπ‘š
π‘ž → Fπ‘ž } → {Fπ‘ž π‘š → Fπ‘ž } defined by πœ‘ (𝑓 ) =
𝑓 ∘ πœ‘. It suffices to show that 𝑆 is an interpolating set for π’ž ′ , πœ‘* (π’ž). Observe that π’ž ′
is affine-invariant over Fπ‘žπ‘š , and hence has a degree set Deg(π’ž ′ ), by Proposition 4.1.3. By
Proposition 4.1.4, dimFπ‘ž (π’ž ′ ) = | Deg(π’ž ′ )|, so suppose Deg(π’ž ′ ) = {𝑖1 , . . . , 𝑖𝐷 }. Every 𝑔 ∈ π’ž ′
∑οΈ€
𝑖𝑗
is of the form 𝑔(𝑍) = 𝐷
𝑗=1 π‘Žπ‘— 𝑍 , where π‘Žπ‘— ∈ Fπ‘ž π‘š . By linearity, it suffices to show that if
𝑔 ∈ π’ž ′ is nonzero, then 𝑔(𝑧) ΜΈ= 0 for some 𝑧 ∈ 𝑆. We have
⎑
πœ” 𝑖1
πœ” 𝑖2
···
πœ” 𝑖𝐷
⎀⎑
π‘Ž1
⎀
⎑
𝑔(πœ”)
⎀
⎒
βŽ₯
βŽ₯⎒ βŽ₯ ⎒
⎒ 2𝑖1
⎒ βŽ₯ ⎒
2 βŽ₯
2𝑖2
2𝑖𝐷 βŽ₯
⎒
βŽ’πœ”
βŽ₯
βŽ₯
⎒
𝑔(πœ” ) βŽ₯
πœ”
··· πœ”
π‘Ž
⎒
βŽ₯
βŽ₯⎒ 2βŽ₯ = ⎒
⎒ ..
..
.. βŽ₯ ⎒ .. βŽ₯ ⎒ .. βŽ₯
..
.
⎒ .
.
. βŽ₯⎒ . βŽ₯ ⎒ . βŽ₯
⎣
⎦
⎦⎣ ⎦ ⎣
𝑔(πœ” 𝐷 )
πœ” 𝐷𝑖1 πœ” 𝐷𝑖2 · · · πœ” 𝐷𝑖𝐷
π‘Žπ·
and the leftmost matrix is invertible since it is a generalized Vandermonde matrix. Therefore,
if 𝑔 ΜΈ= 0, then the right-hand side, which is simply the vector of evaluations of 𝑔 on 𝑆, is
nonzero.
Theorem 5.3.4. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant with distance 𝛿 , 𝛿(π’ž), and
(οΈ€
)οΈ€
let π‘š ≥ 𝑑. Then, for every πœ–, πœ‚ > 0, the lifted code π’ž π‘‘β†—π‘š is 𝑄, ( 12 − πœ–)𝛿 − π‘ž −𝑑 , πœ‚ -locally
decodable where 𝑄 = 𝑂(ln(1/πœ‚)/πœ–2 ).
Proof. Follows immediately from Theorems 5.3.2 and 5.3.3 and Proposition 3.2.5.
58
5.4
5.4.1
Local Testing and Robust Testing
Local Testing
Many testing results may be viewed as robustness statements about properties. A characterization statement would say “an object 𝑋 has a property 𝑃 globally if and only if it
has property 𝑃 locally”. For example, degree 𝑑 polynomials over Fπ‘ž of characteristic 𝑝 are
⌉︁
⌈︁
𝑑+1
,
characterized by the fact that, when restricted to 𝑑-dimensional subspaces, for 𝑑 = π‘ž−π‘ž/𝑝
they are degree 𝑑 polynomials. This was proved in [KR06]. A testing statement would go
further to say “if 𝑋 does not have property 𝑃 globally, then locally it often fails to have
property 𝑃 ”. For example, building on our previous example, a low-degree testing statement
would say that if a polynomial over Fπ‘ž of characteristic 𝑝 has degree greater than 𝑑, then on
many 𝑑-dimensional subspaces, it has degree greater than 𝑑. This was also proved in [KR06].
A robust testing statement would go even further to say “if 𝑋 is far from having property
𝑃 globally, then locally its average distance from 𝑃 is also far”.
The lift operator is natural for many reasons, but primarily because it suggests such a
natural test for lifted codes. The lifted code π’ž π‘‘β†—π‘š , by construction, is locally characterized
by the fact that any codeword, when restricted to 𝑑-dimensional subspaces, is a codeword
of π’ž. Along with the symmetry provided by affine-invariance, the work of Kaufman and
Sudan [KS08] immediately imply that the natural 𝑑-dimensional test is Ω(π‘ž −2𝑑 )-sound, i.e. if
𝑓∈
/ π’ž π‘‘β†—π‘š , then on at least Ω(π‘ž −2𝑑 )-fraction of 𝑑-dimensional subspaces 𝐴, 𝑓 |𝐴 ∈
/ π’ž.
Definition 5.4.1. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant, and let 𝑑 ≤ π‘˜ ≤ π‘š. The
natural π‘˜-dimensional test for π’ž π‘‘β†—π‘š is the π‘ž π‘˜ -local tester with the following distribution: it
π‘š
selects a uniformly random π‘˜-flat 𝐴 ⊆ Fπ‘š
π‘ž and accepts a function 𝑓 : Fπ‘ž → Fπ‘ž if and only if
𝑓 |𝐴 ∈ π’ž π‘‘β†—π‘˜ .
Theorem 5.4.2 ([KS08]). Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant, and let π‘š ≥ 𝑑.
Then the natural 𝑑-dimensional test for the lifted code π’ž π‘‘β†—π‘š is (π‘ž −2𝑑 /2)-sound.
59
5.4.2
Robust Testing
However, for our high-rate code constructions, we often take π‘š to be constant and let π‘ž → ∞
to make our code longer. As such, a soundness of Ω(π‘ž −𝑂(𝑑) ) is insufficient as this quickly
approaches zero. In fact, in Chapter 6, we prove that the natural 2𝑑-dimensional (as opposed
to 𝑑-dimensional) test is 𝛼-robust, where 𝛼 is a polynomial in the distance of the code, but
does not depend on π‘ž, 𝑑, or π‘š. In particular, by Proposition 3.3.2, this implies that the
2𝑑-dimensional test is 𝛼-sound.
Theorem 5.4.3. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant, and let π‘š ≥ 𝑑. Then the
natural (2𝑑)-dimensional test for the lifted code π’ž π‘‘β†—π‘š is
𝛿(π’ž)72
-robust.
2·1052
The proof of Theorem 5.4.3 is somewhat long and technical, so we defer it to Chapter 6.
60
Chapter 6
Robust Testing of Lifted Codes
6.1
Robustness of Lifted Codes
In this section, we prove Theorem 6.1.18, which is simply a more precise restatement of
Theorem 1.2.2. That is, we prove that the natural 2𝑑-dimensional test for the π‘š-dimensional
lift of a 𝑑-dimensional code over Fπ‘ž is 𝛼-robust, where 𝛼 depends only on the distance of the
code and not on 𝑑 or π‘š or the field size. Our approach is a standard one — we first analyze
the test for low-dimensional settings (Theorem 6.1.4), and then use a general projection
argument (“bootstrapping”) to get an analysis for all dimensions (Theorem 6.1.18).
6.1.1
Preliminaries
We begin by presenting some basic definitions and results on robust testing. We define the
robustness of a lifted code, specializing the definition to robustness with respect to subspace
testers. We include the dimension of the testing subspace as a parameter in the robustness
since this will be convenient later.
Definition 6.1.1. Let 𝑑 ≤ π‘˜ ≤ π‘š. The code π’ž π‘‘β†—π‘š is (𝛼, π‘˜)-robust if, for every π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž ,
)οΈ€
[οΈ€ (οΈ€
)οΈ€]οΈ€
(οΈ€
E𝐴 𝛿 π‘Ÿ|𝐴 , π’ž π‘‘β†—π‘˜ ≥ 𝛼 · 𝛿 π‘Ÿ, π’ž π‘‘β†—π‘š
61
where the expectation is over uniformly random π‘˜-dimensional subspaces 𝐴 ⊆ Fπ‘š
π‘ž . When π‘˜
is clear from context, we say the code is 𝛼-robust.
Observe that if 𝐴 is a random π‘˜1 -dimensional affine subspace and 𝐡 is a random π‘˜2 dimensional affine subspace, where π‘˜2 ≥ π‘˜1 , then
)οΈ€]οΈ€
)οΈ€]οΈ€]οΈ€
[οΈ€ (οΈ€
)οΈ€]οΈ€
[οΈ€
[οΈ€ (οΈ€
[οΈ€ (οΈ€
≤ E𝐡 𝛿 π‘Ÿ|𝐡 , π’ž π‘‘β†—π‘˜2
E𝐴 𝛿 π‘Ÿ|𝐴 , π’ž π‘‘β†—π‘˜1 = E𝐡 E𝐴⊆𝐡 𝛿 π‘Ÿ|𝐴 , π’ž π‘‘β†—π‘˜1
so if π’ž π‘‘β†—π‘š is (𝛼, π‘˜1 )-robust, then it is also (𝛼, π‘˜2 )-robust.
As a corollary to Theorem 5.4.2, the π‘˜-dimensional test (for π‘˜ ≥ 𝑑) for π’ž π‘‘β†—π‘š is 𝑂(π‘ž −3𝑑 )robust.
−3𝑑
Corollary 6.1.2. If π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } is linear affine-invariant, then π’ž π‘‘β†—π‘š is ( π‘ž 2 , π‘˜)-robust
for π‘˜ ≥ 𝑑.
−3𝑑
Proof. It suffices to show that π’ž π‘‘β†—π‘š is ( π‘ž 2 , 𝑑)-robust, which follows immediately from Theorem 5.4.2 and Proposition 3.3.2.
Proposition 6.1.3 (Robustness composes multiplicatively). Let 𝑑 ≤ π‘˜1 ≤ π‘˜2 ≤ π‘š
and let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant. If π’ž π‘‘β†—π‘š is (𝛼2 , π‘˜2 )-robust and π’ž π‘‘β†—π‘˜2 is
(𝛼1 , π‘˜1 )-robust, then π’ž π‘‘β†—π‘š is (𝛼1 · 𝛼2 , π‘˜1 )-robust.
6.1.2
Robustness for Small Dimension
Throughout Sections 6.1.2 and 6.1.3, fix a linear affine invariant code π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } with
relative distance 𝛿 , 𝛿(π’ž). Let 𝑛 ≥ 1 be an integer (we will use 𝑛 = 3 or 4) and let π‘š = 𝑛𝑑.
Two codes that play a prominent role are the lift π’ž π‘‘β†—π‘š of π’ž to π‘š dimensions, and the
𝑛-fold tensor product π’ž ⊗𝑛 of π’ž, which is also an π‘š-dimensional code.
We begin by giving a tester with robust analysis for π’ž π‘‘β†—π‘š for this restricted choice of
(οΈ€ 𝑛 )︀𝑂(1)
π‘š. We will show that the (π‘š − 𝑑)-dimensional test is 𝛿𝑛
-robust. (Note the robustness
degrades poorly with 𝑛 = π‘š/𝑑 and so can only be applied for small 𝑛). It is important, for
Section 6.1.4, that there is no dependence on 𝑑.
62
Theorem 6.1.4. Let 𝑛 ≥ 3 and 𝑑 ≥ 1 and set π‘š = 𝑛𝑑. Then π’ž π‘‘β†—π‘š is (𝛼0 , π‘š − 𝑑)-robust for
𝛼0 =
𝛿 3𝑛
.
16(𝑛2 +3𝑛+2)3
Overview. For simplicity, we describe the proof idea for 𝑑 = 1. Suppose the average
local distance to π’ž 1β†—(π‘š−1) on random hyperplanes is small. For a ∈ Fπ‘š
π‘ž , let π’ža be the
code consisting of tensor codewords in π’ž ⊗𝑛 whose restrictions to lines in direction a are also
β‹‚οΈ€
codewords of π’ž. Note that a π’ža = π’ž 1β†—π‘š . Our main technical result (Theorem 6.1.11) of this
(οΈ€ 𝑛 )︀𝑂(1)
-robust. Now, observe that choosing a random hyperplane
section shows that π’ža is 𝛿𝑛
can be done by choosing π‘š random linearly independent directions, choosing an additional
random direction a that is not spanned by any π‘š − 1 of the former, and choosing a random
hyperplane spanned by π‘š − 1 of these π‘š + 1 random directions (call such a hyperplane
“special”). Viewing the first π‘š chosen directions as the standard basis directions, we see
that the average local distance to π’ž 1β†—(π‘š−1) , and hence to π’ž ⊗(π‘š−1) , when restricting to special
hyperplanes, is still small. Therefore, for most a, the average local distance to π’ž ⊗(π‘š−1) on
special hyperplanes is small. By the robustness of π’ža , this implies that our received word
is close to some codeword 𝑐a ∈ π’ža for most a. But these codewords 𝑐a are all codewords
of π’ž ⊗π‘š and close to each other, so they must be the same codeword 𝑐 ∈ π’ž ⊗π‘š . So we have
shown that we are close to 𝑐 ∈ π’ž ⊗π‘š . We proceed by showing that in fact 𝑐 ∈ π’ž 1β†—π‘š . Note
that 𝑐 ∈ π’ža for most a. Another technical result, Corollary 6.2.7, implies that in fact 𝑐 ∈ π’ža
for all a and we are done.
To generalize to 𝑑 > 1, we replace dimension π‘˜ subspaces with dimension π‘˜π‘‘ subspaces
throughout. Some work needs to be done to ensure that Theorem 6.1.11 still works, and also
Corollary 6.2.7 must be generalized appropriately to remove the dependence on 𝑑. These
issues will be discussed in the corresponding sections.
(οΈ€ π‘š )οΈ€
Definition 6.1.5. Let β„“ ≤ π‘š be an integer. A collection 𝐷 ⊆ Fπ‘‘π‘ž is β„“-proper if for every
⋃︀
π‘˜ and every distinct 𝐴1 , . . . , π΄π‘˜ ∈ 𝐷, the union π‘˜π‘–=1 𝐴𝑖 contains at least min {π‘˜π‘‘, π‘š − β„“}
linearly independent vectors.
Definition 6.1.6. For a set 𝐷 ⊆
(οΈ€Fπ‘š )οΈ€
π‘˜
π‘ž
, for every x ∈ Fπ‘š
π‘ž define 𝒱𝐷 (x) to be the collection
𝑑
63
of subspaces through x in directions from π‘˜ different sets from 𝐷. More precisely,
βƒ’
(οΈ‚ )οΈ‚}οΈƒ
π‘˜
βƒ’
⋃︁
𝐷
βƒ’
(x, 𝐴) βƒ’ 𝐴 =
𝐷𝑖 , {𝐷1 , . . . , π·π‘˜ } ∈
βƒ’
π‘˜
{οΈƒ
π’±π·π‘˜ (x) ,
𝑖=1
Define
π’±π·π‘˜ ,
⋃︁
π’±π·π‘˜ (x).
x∈F𝑛
π‘ž
The testing subspaces through x are 𝒯𝐷 (x) , π’±π·π‘š−1 (x) and the decoding subspaces through
x are π’Ÿπ· (x) , 𝒱𝐷1 (x). Similarly, the testing subspaces are 𝒯𝐷 , π’±π·π‘š−1 and the decoding
subspaces are π’Ÿπ· , 𝒱𝐷1 . If 𝑆 = (x, ∪π‘˜π‘–=1 𝐷𝑖 ) ∈ π’±π·π‘˜ , then the blocks of 𝑆 are the sets
𝐷1 , . . . , π·π‘˜ . Two testing subspaces are adjacent if they differ in at most one block.
Remark 6.1.7. If 𝐷 is β„“-proper for β„“ ≤ 𝑑 then for any π‘˜ < 𝑛 we have that π’±π·π‘˜ consists of
π‘˜π‘‘-dimensional subspaces.
𝑛
Definition 6.1.8. Define π’žπ·
to be the code of all words 𝑓 : Fπ‘š
π‘ž → Fπ‘ž such that 𝑓 |𝑒 ∈ π’ž for
every decoding subspace 𝑒 ∈ π’Ÿπ· .
𝑛
for any 𝐷. If
Remark 6.1.9. Observe that π’ž π‘‘β†—π‘š is a subcode of π’žπ·
⋃︀
𝐷 contains the
𝑛
standard basis vectors, then π’žπ·
is a subcode of π’ž ⊗𝑛 .
⌊οΈ‚
⌋οΈ‚
𝑛 log( 1𝛿 )+log(𝑛2 +3𝑛+2)+1
Proof of Theorem 6.1.4. Define β„“ ,
. We note that for the most
log(π‘ž)
interesting cases, where 𝛿 > 0 and 𝑛 are fixed and π‘ž → ∞, β„“ = 0. Our first step handles
the less interesting cases (by appealing to a known result). Specifically, if β„“ ≥ 𝑑 then by
Corollary 6.1.2 we are done since
(οΈƒ
π‘ž −3𝑑
π‘ž −3β„“
π‘ž
≥
≥
2
2
−3
( )
𝑛 log 1 +log(𝑛2 +3𝑛+2)+1
𝛿
log(π‘ž)
)οΈƒ
=
2
𝛿 3𝑛
= 𝛼0 .
16(𝑛2 + 3𝑛 + 2)3
[οΈ€ (οΈ€
)οΈ€]οΈ€
Now assume β„“ < 𝑑 and let 𝜌 , E𝑣 𝛿 π‘Ÿ|𝑣 , π’ž 𝑑↗(π‘š−𝑑) , where 𝑣 ⊆ Fπ‘š
π‘ž is a uniformly random
(π‘š − 𝑑)-dimensional affine subspace. We will assume without lost of generality that 𝜌 ≤ 𝛼0
and in particular 𝜌 ≤
𝛿 3𝑛
𝑛+1 2 2
(𝑛 +3𝑛+2)
𝑛−1
16(
)
≤
𝛿 2𝑛 π‘ž −β„“
𝑛+1
8(𝑛−1
)
2
64
.
Observe that
[οΈ€ (οΈ€
)οΈ€]οΈ€
𝑑↗(π‘š−𝑑)
𝑛
𝜌 = E𝐴1 ,...,𝐴𝑛 ∈(Fπ‘š
π‘ž E𝐴∈ Fπ‘ž E𝑣∈𝒯{𝐴 ,...,𝐴𝑛 ,𝐴} 𝛿 π‘Ÿ|𝑣 , π’ž
1
)
( )
𝑑
𝑑
where 𝐴1 , . . . , 𝐴𝑛 are random sets such that their union is linearly independent, and 𝐴 is a
random set such that {𝐴1 , . . . , 𝐴𝑛 , 𝐴} is β„“-proper. Fix 𝐴1 , . . . , 𝐴𝑛 such that
[οΈ€ (οΈ€
)οΈ€]οΈ€
𝑑↗(π‘š−𝑑)
E𝐴∈(Fπ‘š
≤ 𝜌.
π‘ž E𝑣∈𝒯{𝐴 ,...,𝐴𝑛 ,𝐴} 𝛿 π‘Ÿ|𝑣 , π’ž
1
)
𝑑
Since π’ž 𝑑↗(π‘š−𝑑) ⊆ π’ž ⊗(𝑛−1) ,
[οΈ€ (οΈ€
)οΈ€]οΈ€
⊗(𝑛−1)
E𝐴∈(Fπ‘š
≤ 𝜌.
π‘ž E𝑣∈𝒯{𝐴 ,...,𝐴𝑛 ,𝐴} 𝛿 π‘Ÿ|𝑣 , π’ž
1
)
𝑑
By affine-invariance, we may assume without loss of generality that 𝐴1 , . . . , 𝐴𝑛 form the
(οΈ€Fπ‘š )οΈ€
π‘ž
standard basis vectors for Fπ‘š
, let 𝐷𝐴 , {𝐴1 , . . . , 𝐴𝑛 , 𝐴}. By Markov’s
π‘ž . For any 𝐴 ∈
𝑑
inequality,
[︁
[οΈ€ (οΈ€
)οΈ€
]οΈ€]︁ 1 𝑛
⊗(𝑛−1)
−𝑛
Pr E𝑣∈𝒯𝐷𝐴 𝛿 π‘Ÿ|𝑣 , π’ž
≥ 2𝛿 𝜌 < 𝛿 .
𝐴
2
So, for more than 1 − 12 𝛿 𝑛 fraction of blocks 𝐴 such that 𝐷𝐴 is β„“-proper, we have a codeword
(︀𝑛+1)οΈ€
⊗𝑛
−𝑛
𝑛
⊆
π’ž
such
that
(by
Theorem
6.1.11)
𝛿(π‘Ÿ,
𝑐
)
<
2𝛿
𝜌
< 12 𝛿 𝑛 . For every
𝑐𝐴 ∈ π’žπ·
𝐴
𝑛−1
𝐴
two such blocks 𝐴, 𝐴′ , we have 𝛿(𝑐𝐴 , 𝑐𝐴′ ) ≤ 𝛿(𝑐𝐴 , π‘Ÿ) + 𝛿(π‘Ÿ, 𝑐𝐴′ ) < 𝛿 𝑛 = 𝛿(π’ž ⊗𝑛 ), so there is
some codeword 𝑐 ∈ π’ž ⊗𝑛 such that 𝑐𝐴 = 𝑐 for every such 𝐴. For such 𝐴, it follows that for
every b ∈ Fπ‘š
π‘ž , the restriction of 𝑐 to the subspace (b, 𝐴) is a codeword of π’ž, i.e. 𝑐|(b,𝐴) ∈ π’ž.
(οΈ€ )οΈ€ π‘ž−𝑑
π‘ž −𝑙
By Claim 6.1.17, for more than 1 − 21 𝛿 𝑛 − 𝑛2 π‘ž−1
− 𝑛 π‘ž−1
fraction of blocks 𝐴 (without the
⊗𝑛
requirement that 𝐷𝐴 be proper), 𝑐|(b,𝐴) ∈ π’ž for every b ∈ Fπ‘š
and
π‘ž . In particular, 𝑐 ∈ π’ž
(οΈ€
)οΈ€
−𝑑
−β„“
π‘ž
π‘ž
for every b ∈ Fπ‘š
/ π’ž for less than 12 𝛿 𝑛 + 𝑛2 π‘ž−1
+ 𝑛 π‘ž−1
fraction of 𝐴. It sufficient to
π‘ž , 𝑐|(b,𝐴) ∈
(οΈ€
)οΈ€
π‘ž −𝑑
π‘ž −𝑙
show that 21 𝛿 𝑛 + 𝑛2 π‘ž−1
+ 𝑛 π‘ž−1
≤ 𝛿 𝑛 − (𝑛 + 1)π‘ž −𝑑 . Then it will follow from Corollary 6.2.7
65
that 𝑐 ∈ π’ž π‘‘β†—π‘š and since 𝛿(π‘Ÿ, 𝑐) ≤ 2𝛿 −𝑛 𝜌
(︀𝑛+1)οΈ€
𝑛−1
we are done. Calculating:
(οΈ‚ )οΈ‚ −𝑑
(οΈ‚ )οΈ‚ −𝑙
𝑛 π‘ž
π‘ž −𝑙
𝑛 π‘ž
π‘ž −𝑙
π‘ž −𝑙
−𝑑
+𝑛
+ (𝑛 + 1)π‘ž
≤
+𝑛
+ (𝑛 + 1)
2 π‘ž−1
π‘ž−1
2 π‘ž−1
π‘ž−1
π‘ž−1
(οΈ‚ 2
)οΈ‚
−𝑙
𝑛 + 3𝑛 + 2
π‘ž
=
π‘ž−1
2
𝑛
π‘žπ›Ώ
≤
4(π‘ž − 1)
1 𝑛
𝛿 .
≤
2
The composability of robust tests immediately yields the following corollary where the
test is now 2𝑑 dimensional.
21
𝛿
Corollary 6.1.10. π’ž 𝑑↗4𝑑 is (𝛼1 , 2𝑑)-robust, where 𝛼1 ≥ 6·10
10 .
(︁ 12
)︁
(︁ 9
)︁
𝛿
𝛿
Proof. By Theorem 6.1.4, π’ž 𝑑↗4𝑑 is 432,000
, 3𝑑 -robust and π’ž 𝑑↗3𝑑 is 128,000
, 2𝑑 -robust. There-
fore, by composing, the 2𝑑-dimensional robustness of π’ž 𝑑↗4𝑑 is at least
𝛿 12
432,000
·
𝛿9
128,000
=
𝛿 21
55,296,000,000
6.1.3
Robustness of Special Tensor Codes
In this section, we prove the main technical result (Theorem 6.1.11) used in Section 6.1.2.
(οΈ€ 𝑛 )οΈ€
Theorem 6.1.11. Let 𝑛 ≥ 3 and β„“ ≤ 𝑑. Set π‘š = 𝑛𝑑. Let 𝐷 ⊆ Fπ‘‘π‘ž be β„“-proper with |𝐷| ≥ 𝑛
[οΈ€ (οΈ€
)οΈ€]οΈ€
𝑛 π‘ž −β„“
⊗(𝑛−1)
blocks. Let π‘Ÿ : Fπ‘š
. If 𝜌 < 𝛿 |𝐷|
2 , then
π‘ž → Fπ‘ž be a word with 𝜌 , E𝑣∈𝒯𝐷 𝛿 π‘Ÿ|𝑣 , π’ž
4(𝑛−1)
(οΈ€
)οΈ€
|𝐷|
𝑛
𝛿(π‘Ÿ, π’žπ·
) ≤ 𝜌 𝑛−1
.
Overview. Our analysis is an adaptation of Viderman’s [Vid12]. For simplicity, assume
𝑑 = 1. We define a function 𝑐 : Fπ‘š
π‘ž → Fπ‘ž , which we show is both close to π‘Ÿ and a codeword
𝑛
of π’žπ·
. Following Viderman’s analysis, we partition Fπ‘š
π‘ž into “good”, “fixable”, and “bad”
points. Each hyperplane 𝑣 ∈ 𝒯𝐷 has an associated codeword 𝑐𝑣 ∈ π’ž ⊗(π‘š−1) , the nearest
codeword to π‘Ÿ|𝑣 , and an opinion 𝑐𝑣 (x) about x. “Good” points are points for which any
66
hyperplane agrees with π‘Ÿ. “Fixable” points are points for which hyperplanes agree with each
other, but not with π‘Ÿ. “Bad” points are points for which at least two hyperplanes disagree
with each other. For good or fixable x, we naturally define 𝑐(x) to be the common opinion
𝑐𝑣 (x) of any hyperplane 𝑣 through x. Claim 6.1.13 implies that there are not many bad
points, which immediately shows that 𝑐 is close to π‘Ÿ.
So far, our proof has been a straightforward adaptation of Viderman’s. However, at
this point, we are forced to depart from Viderman’s proof. A hyperplane is “bad” if it
has more than 12 𝛿 π‘š−1 fraction bad points. Claim 6.1.12 shows that every bad point is in
a bad hyperplane, and Claim 6.1.14 shows that there are less than 21 π›Ώπ‘ž bad hyperplanes.
𝑛
and 𝒯𝐷 , this is
In [Vid12], which analyses π’ž ⊗π‘š and axis-parallel hyperplanes instead of π’žπ·
already enough, since this implies that in each axis-parallel direction, there are less than
π›Ώπ‘ž bad hyperplanes, so the remaining points are all good or fixable and with a little bit
more work, one can show that 𝑐 can be extended uniquely to a tensor codeword using the
erasure-decoding properties of tensor codes. Unfortunately, we do not have this structure.
We say a line is “good” if it is contained in some good hyperplane, otherwise it is bad.
We must further partition the bad points into merely bad and “super-bad” points, which
are points such that either every hyperplane is bad, or there are two disagreeing good hyperplanes. For merely bad x, we define 𝑐(x) to be the common opinion 𝑐𝑣 (x) of any good
hyperplane 𝑣 through x. For super-bad x, we pick any line 𝑒 through x, take the restriction
of 𝑐 to the non-super-bad points on 𝑒, and extend it to a codeword 𝑐𝑒 ∈ π’ž, and define
𝑐(x) , 𝑐𝑒 (x). Two non-trivial steps remain: showing that 𝑐(x) is well-defined for super-bad
𝑛
x, and showing that 𝑐 ∈ π’žπ·
.
Claim 6.1.15 shows that, for any special plane, there are less than
1
π›Ώπ‘ž
2
lines in each
direction that are bad (not contained in any good hyperplane) or contain a super-bad point.
This is proved by exhibiting, for each such undesirable line, a bad hyperplane in a fixed
direction containing the line. If there were too many undesirable lines, this would result in
too many parallel bad hyperplanes, contradicting Claim 6.1.14. Finally, Claim 6.1.16 shows
that if 𝑒 is a line with no super-bad points, then 𝑐|𝑒 ∈ π’ž is a codeword.
67
Now, we show that 𝑐 is well-defined on super-bad x. Let 𝑒1 , 𝑒2 be two lines through x. Let
𝑃 be the plane through x containing 𝑒1 , 𝑒2 . On this plane, by Claim 6.1.15, in each direction
we have enough lines 𝑒 with no super-bad points, for which 𝑐|𝑒 ∈ π’ž (by Claim 6.1.16), so
that we can uniquely extend 𝑐 onto the entire plane (by Proposition A.2.5). This gives a
well-defined value for 𝑐(x).
𝑛
Finally, we show that 𝑐 ∈ π’žπ·
. Let 𝑒 be any line. If 𝑒 has no super-bad points, then
𝑐|𝑒 ∈ π’ž follows from Claim 6.1.16. If 𝑐 does have a super-bad point x, then 𝑐|𝑒 ∈ π’ž by the
way we defined 𝑐(x).
This completes our analysis for the case 𝑑 = 1. To generalize to 𝑑 > 1, we replace lines
with “decoding subspaces” (subspaces of dimension 𝑑), planes with subspaces of dimension
2𝑑, and hyperplanes with “testing subspaces” (subspaces of codimension 𝑑). Some care must
be taken when proving Claim 6.1.15, because the intersection of two decoding subspaces
may have non-trivial dimension. We therefore require the notion of “β„“-properness” of 𝐷,
and must modify Claim 6.1.14 and also prove Claim 6.1.17 to accommodate this notion.
Details follow.
Proof of Theorem 6.1.11. For each testing subspace 𝑣 ∈ 𝒯𝐷 , define 𝑐𝑣 ∈ π’ž ⊗(𝑛−1) to be the
closest codeword to π‘Ÿ|𝑣 (break ties arbitrarily). We will partition Fπ‘š
π‘ž into three disjoint sets
𝐺, 𝐹, 𝐡 (good, fixable, and bad points, respectively) as follows:
𝐺 ,
{οΈ€
x ∈ Fπ‘š
π‘ž | 𝑐𝑣 (x) = π‘Ÿ(x) for every 𝑣 ∈ 𝒯𝐷 (x)
}οΈ€
}οΈ€
′
x ∈ Fπ‘š
π‘ž | 𝑐𝑣 (x) = 𝑐𝑣 ′ (x) ΜΈ= π‘Ÿ(x) for every 𝑣, 𝑣 ∈ 𝒯𝐷 (x)
{οΈ€
}οΈ€
′
𝐡 , x ∈ Fπ‘š
π‘ž | 𝑐𝑣 (x) ΜΈ= 𝑐𝑣 ′ (x) for some 𝑣, 𝑣 ∈ 𝒯𝐷 (x) .
𝐹 ,
{οΈ€
Call a testing subspace bad if at least 21 𝛿 𝑛−1 fraction of its points are in 𝐡, and good otherwise.
A decoding subspace is good if it is contained in some good testing subspace, and bad
otherwise. Further, define the set 𝐡 ′ of super-bad points
𝐡 ′ , {x ∈ 𝐡 | every 𝑣 ∈ 𝒯𝐷 (x) is bad or ∃ good 𝑣, 𝑣 ′ ∈ 𝒯𝐷 (x) such that 𝑐𝑣 (x) ΜΈ= 𝑐𝑣′ (x)}.
68
Claim 6.1.12. If 𝑣, 𝑣 ′ ∈ 𝒯𝐷 are adjacent good testing subspaces, then 𝑐𝑣 |𝑣∩𝑣′ = 𝑐𝑣′ |𝑣∩𝑣′ . In
particular, every bad point is in a bad testing subspace.
Proof. Suppose b ∈ 𝑣 ∩ 𝑣 ′ and 𝑐𝑣 (b) ΜΈ= 𝑐𝑣′ (b). Since 𝑣, 𝑣 ′ are adjacent, they have 𝑛 − 2
blocks 𝐴1 , . . . , 𝐴𝑛−2 in common. Let 𝑣 have blocks 𝐴1 , . . . , 𝐴𝑛−2 , 𝐴 and let 𝑣 ′ have blocks
𝐴1 , . . . , 𝐴𝑛−2 , 𝐴′ .
Let 𝑒 ∈ π’Ÿπ· be the decoding subspace (b, 𝐴1 ). Since 𝑐𝑣 |𝑒 , 𝑐𝑣′ |𝑒 ∈ π’ž disagree on b, they
are distinct codewords and hence disagree on at least π›Ώπ‘ž 𝑑 points of 𝑒, say x1 , . . . , xπ›Ώπ‘žπ‘‘ . For
each 𝑖 ∈ [π›Ώπ‘ž 𝑑 ], let 𝑣𝑖 ∈ 𝒯𝐷 be the testing subspace (x𝑖 , 𝐴2 ∪ · · · ∪ 𝐴𝑛−2 ∪ 𝐴 ∪ 𝐴′ ).
Since 𝑐𝑣 (x𝑖 ) ΜΈ= 𝑐𝑣′ (x𝑖 ), that means 𝑐𝑣𝑖 disagrees with one of 𝑐𝑣 , 𝑐𝑣′ at x𝑖 . Without loss
of generality, suppose 𝑐𝑣 disagrees with 𝑐𝑣1 , . . . , π‘π‘£π›Ώπ‘žπ‘‘ /2 . We will show that 𝑣 is bad, which
proves the first part of the claim.
For each 𝑖 ∈ [π›Ώπ‘ž 𝑑 ], let 𝑀𝑖 = (x𝑖 , 𝐴2 ∪ · · · ∪ 𝐴𝑛−2 ∪ 𝐴). Note that 𝑒 ∩ 𝑀𝑖 = {x𝑖 } (since
𝐷 is β„“-proper, for β„“ ≤ 𝑑), so all 𝑀𝑖 are different parallel subspaces and hence disjoint. Since
𝑀𝑖 ∈ 𝒱𝐷𝑛−2 , the restrictions 𝑐𝑣 |𝑀𝑖 , 𝑐𝑣𝑖 |𝑀𝑖 ∈ π’ž ⊗𝑛−2 are codewords and are distinct because they
disagree on x𝑖 , therefore, by Proposition A.2.4, they disagree on at least 𝛿 𝑛−2 π‘ž π‘š−2𝑑 points in
𝑀𝑖 , which are therefore bad. Thus, each 𝑣𝑖 contributes 𝛿 𝑛−2 π‘ž π‘š−2𝑑 bad points to 𝑣, for a total
of 21 𝛿 𝑛−1 π‘ž π‘š−𝑑 bad points since the 𝑀𝑖 are disjoint.
For the second part, suppose b ∈ 𝐡 is a bad point. We will show that b lies in a bad
testing subspace. By definition, there are two testing subspaces 𝑣, 𝑣 ′ ∈ 𝒯𝐷 (b) such that
𝑐𝑣 (b) ΜΈ= 𝑐𝑣′ (b). Suppose 𝑣 has blocks 𝐴1 , . . . , 𝐴𝑛−1 and 𝑣 ′ has directions 𝐴′1 , . . . , 𝐴′𝑛−1 . Assume, without lost of generality, that if 𝐴𝑖 = 𝐴′𝑗 then 𝑖 = 𝑗. Define 𝑣0 , 𝑣, and for 𝑖 ∈ [𝑛−1],
define 𝑣𝑖 ∈ 𝒯𝐷 to be the testing subspace through b in directions 𝐴′1 , . . . , 𝐴′𝑖 , 𝐴𝑖+1 , . . . , 𝐴𝑛−1 .
Consider the sequence 𝑣0 , 𝑣1 , . . . , 𝑣𝑛−1 of testing subspaces. For each 𝑖, the testing subspaces
𝑣𝑖 , 𝑣𝑖+1 are adjacent. Since 𝑐𝑣0 (b) ΜΈ= 𝑐𝑣𝑛−1 (b), there exists some 𝑖 such that 𝑐𝑣𝑖 (b) ΜΈ= 𝑐𝑣𝑖+1 (b),
and by the first part of the claim it follows that one of 𝑣𝑖 , 𝑣𝑖+1 is bad.
Claim 6.1.13. 𝜌 ≥
|𝐹 |
π‘žπ‘š
+
|𝐡|
|𝐷|
π‘ž π‘š (𝑛−1
)
69
Proof. Observe that |𝒯𝐷 | = π‘ž 𝑑
(οΈ€ |𝐷| )οΈ€
. Therefore,
𝑛−1
)οΈ€]οΈ€
[οΈ€ (οΈ€
𝜌 = E𝑣∈𝒯𝐷 𝛿 π‘Ÿ|𝑣 , π’ž ⊗(𝑛−1)
= E𝑣∈𝒯𝐷 [𝛿 (π‘Ÿ|𝑣 , 𝑐𝑣 )]
∑︁ 1 ∑︁
1
=
1𝑐𝑣 (x)ΜΈ=π‘Ÿ(x)
(οΈ€ |𝐷| )οΈ€
π‘ž 𝑑 𝑛−1 𝑣∈𝒯𝐷 π‘ž π‘š−𝑑 x∈𝑣
∑︁
1
=
#{𝑣 ∈ 𝒯𝐷 (x) | 𝑐𝑣 (x) ΜΈ= π‘Ÿ(x)}
(οΈ€ |𝐷| )οΈ€
π‘ž π‘š 𝑛−1 x∈Fπ‘š
π‘ž
(οΈƒ
)οΈƒ
∑︁
∑︁ (οΈ‚ |𝐷| )οΈ‚ ∑︁
1
0+
+
1
≥
(οΈ€ |𝐷| )οΈ€
π‘š−1
π‘ž π‘š 𝑛−1
x∈𝐺
x∈𝐹
x∈𝐡
=
|𝐡|
|𝐹 |
+ (οΈ€ |𝐷| )οΈ€ .
π‘š
π‘š
π‘ž
π‘ž 𝑛−1
Claim 6.1.14. There are less than 21 π›Ώπ‘ž 𝑑−β„“ bad testing subspaces.
(οΈ€ |𝐷| )οΈ€ π‘š
Proof. By Claim 6.1.13, there are at most |𝐡| ≤ 𝜌 𝑛−1
π‘ž bad points. Each bad testing
(οΈ€ |𝐷| )οΈ€
subspace has at least 𝛿 𝑛−1 π‘ž π‘š−𝑑 /2 bad points by definition. Each bad point has at most 𝑛−1
bad testing subspaces through it. Therefore, the number of testing subspaces is bounded by
(οΈ‚
)οΈ‚
(οΈ‚
)οΈ‚2
|𝐷|
2𝜌
1
|𝐷|
|𝐡|
≤ 𝑛−1
π‘ž 𝑑 < π›Ώπ‘ž 𝑑−β„“ .
1 𝑛−1 π‘š−𝑑 ·
𝑛−1
𝛿
2
𝑛−1
𝛿 π‘ž
2
𝑛
Now we proceed to prove the lemma. We construct a codeword 𝑐 ∈ π’žπ·
with 𝛿(π‘Ÿ, 𝑐) ≤
(οΈ€ |𝐷| )οΈ€
𝜌 𝑛−1 in stages, as follows. First, for x ∈ 𝐺 ∪ 𝐹 , we define 𝑐(x) , 𝑐𝑣 (x) for any testing
subspace 𝑣 ∈ 𝒯𝐷 (x). This is well-defined by the definition of 𝐺 and 𝐹 . Furthermore, since
(οΈ€ |𝐷| )οΈ€
𝑐(x) = 𝑐𝑣 (x) = π‘Ÿ(x) for x ∈ 𝐺, we already guarantee that 𝛿(π‘Ÿ, 𝑐) ≤ |𝐹 |+|𝐡|
≤
𝜌
.
π‘š
𝑛−1
π‘ž
For x ∈ 𝐡 βˆ– 𝐡 ′ , define 𝑐(x) , 𝑐𝑣 (x) for any good testing subspace 𝑣 ∈ 𝒯𝐷 (x), whose
existence is guaranteed by the fact that x ∈
/ 𝐡 ′ . This is well-defined because if 𝑣, 𝑣 ′ ∈ 𝒯𝐷 (x)
are both good, then it follows from the fact that x ∈
/ 𝐡 ′ that 𝑐𝑣 (x) = 𝑐𝑣′ (x).
70
Claim 6.1.15. Let 𝑀 ∈ 𝒱𝐷2 be a subspace in directions 𝐴1 , 𝐴2 ∈ 𝐷. For each 𝑖 ∈ {1, 2}, 𝑀
contains less than 12 π›Ώπ‘ž 𝑑 decoding subspaces in direction 𝐴𝑖 which intersect 𝐡 ′ or are bad (not
contained in any good testing subspace).
Proof. By symmetry, it suffices to consider 𝑖 = 2. Let 𝐴3 , . . . , 𝐴𝑛 ∈ 𝐷 be blocks in some
other direction. Let 𝑒1 , . . . , π‘’π‘˜ ⊆ 𝑀 be decoding subspaces in direction 𝐴2 such that, for
each 𝑗 ∈ [π‘˜], 𝑒𝑗 intersects 𝐡 ′ or is bad. It suffices to exhibit, for each 𝑗 ∈ [π‘˜], a bad testing
subspace 𝑣 ∈ 𝒯𝐷 containing 𝑒𝑗 which has block 𝐴2 but not 𝐴1 . In this case we will show
that |𝑣 ∩ 𝑀| ≤ π‘ž 𝑑+β„“ so each such bad testing subspace contain at most π‘ž β„“ of the 𝑒𝑖 -s. Since,
by Claim 6.1.14, there are at most 12 π›Ώπ‘ž 𝑑−β„“ such subspaces, we get that π‘˜ ≤ 12 π›Ώπ‘ž 𝑑 . Indeed,
since 𝐷 is β„“-proper, the subspace 𝑒 + 𝑣 ∈ 𝒱𝐷𝑛 has dimension at least π‘š − β„“. Therefore,
dim(𝑣 ∩ 𝑀) = dim(𝑣) + dim(𝑀) − dim(𝑣 + 𝑀) ≤ π‘š − 𝑑 + 2𝑑 − (π‘š − 𝑙) = 𝑑 + 𝑙
and |𝑣 ∩ 𝑀| ≤ π‘ž 𝑑+β„“ .
Fix 𝑗 ∈ [𝑙] and 𝑒 , 𝑒𝑗 and we will show that 𝑒 is contained in a bad testing subspace.
If 𝑒 is bad, then we are done, since any testing subspace containing 𝑒, in particular the
testing subspace in directions 𝐴2 , . . . , 𝐴𝑛 , is bad. Now suppose 𝑒 has a point x ∈ 𝑒 ∩ 𝐡 ′ .
Let 𝑣 ∈ 𝒯𝐷 be the testing subspace (x, 𝐴2 ∪ · · · ∪ 𝐴𝑛 ). If 𝑣 is bad, we are done. Otherwise,
since x ∈ 𝐡 ′ , there exists another good hyperplane 𝑣 ′ ∈ 𝐷, in directions 𝐴′1 , . . . , 𝐴′𝑛−1 , such
that 𝑐𝑣 (x) ΜΈ= 𝑐𝑣′ (x). Without loss of generality, assume that if 𝐴𝑖 = 𝐴′𝑗 then 𝑖 = 𝑗 (in
particular 𝐴1 ∈
/ {𝐴′2 , . . . , 𝐴′𝑛−1 }). For each 𝑖 ∈ [𝑛 − 1], if 𝐴2 = 𝐴′2 define 𝑣𝑖 ∈ 𝒯𝐷 (x) to be
the testing subspace (x, 𝐴′2 , . . . , 𝐴′𝑖 , 𝐴𝑖+1 , . . . , 𝐴𝑛 ), and if 𝐴2 ΜΈ= 𝐴′2 define 𝑣1 to be 𝑣 and 𝑣𝑖
to be (x, 𝐴2 , 𝐴′2 , . . . , 𝐴′𝑖 , 𝐴𝑖+1 , . . . , 𝐴𝑛−1 ). In any case define 𝑣𝑛 , 𝑣 ′ . For every 𝑖 ∈ [𝑛 − 1],
𝑣𝑖 and 𝑣𝑖+1 are adjacent. Note that for every 𝑖 ∈ [𝑛 − 1], 𝑣𝑖 contains the direction 𝐴2 and
does not contain the direction 𝐴1 . We will show that 𝑣𝑖 is bad for some 𝑖 ∈ [𝑛 − 1]. Since
𝑐𝑣1 (x) ΜΈ= 𝑐𝑣𝑛 (x), there exists some 𝑖 ∈ [𝑛 − 1] such that 𝑐𝑣𝑖 (x) ΜΈ= 𝑐𝑣𝑖+1 (x), and therefore, by
Claim 6.1.12, one of 𝑣𝑖 , 𝑣𝑖+1 is bad. If 𝑖 < 𝑛 − 1, then 𝑖, 𝑖 + 1 ≤ 𝑛 − 1, and so we are done.
If 𝑖 = 𝑛 − 1, then by assumption 𝑣𝑛 = 𝑣 ′ is good, so it must be that 𝑣𝑛−1 is bad.
71
Claim 6.1.16. If 𝑒 ∈ π’Ÿπ· is a decoding subspace and 𝑒 ∩ 𝐡 ′ = ∅, then for every x ∈ 𝑒 there
is a codeword 𝑐x ∈ π’ž defined on 𝑒 such that 𝑐x (x) = 𝑐(x) and 𝛿(𝑐x , 𝑐|𝑒 ) < 2𝛿 . In particular,
𝑐|𝑒 ∈ π’ž.
Proof. Fix x ∈ 𝑒. Let 𝐴 = {a1 , . . . , a𝑑 } ∈ 𝐷 be the directions of 𝑒. Since x ∈
/ 𝐡 ′ , there
is a good testing subspace 𝑣 ∈ 𝒯𝐷 (x). Let 𝐴′ = {a′1 , . . . , a′𝑑 } be some block in 𝑣 not equal
to 𝐴 and consider the subspace 𝑀 = (x, 𝐴 ∪ 𝐴′ ) ∈ 𝒱𝐷2 . For s, s′ ∈ Fπ‘‘π‘ž , let 𝑀(s, s′ ) ,
∑οΈ€
∑οΈ€
x + 𝑑𝑖=1 𝑠𝑖 a𝑖 + 𝑑𝑖=1 𝑠′𝑖 a′𝑖 . Let
𝑀(s, *) , {𝑀(s, s′ ) | s′ ∈ Fπ‘‘π‘ž } ∈ π’Ÿπ· ,
𝑀(*, s′ ) , {𝑀(s, s′ ) | s ∈ Fπ‘‘π‘ž } ∈ π’Ÿπ· .
Let 𝐼 ⊆ Fπ‘‘π‘ž βˆ– {0} be the set of points s ΜΈ= 0 such that 𝑀(s, *) intersects 𝐡 ′ or is bad.
Similarly, let 𝐼 ′ ⊆ Fπ‘‘π‘ž βˆ– {0} be the set of points s′ ΜΈ= 0 such that 𝑀(*, s′ ) intersects 𝐡 ′ or is
bad. By Claim 6.1.15, |𝐼|, |𝐼 ′ | < 12 π›Ώπ‘ž 𝑑 . Note that for each (s, s′ ) ∈ (Fπ‘‘π‘ž βˆ– 𝐼) × (Fπ‘‘π‘ž βˆ– 𝐼 ′ ), we have
𝑀(s, s′ ) ∈
/ 𝐡 ′ : if s ΜΈ= 0 or s′ ΜΈ= 0, this follows from the definition of 𝐼 and 𝐼 ′ ; if s = s′ = 0, then
𝑀(s, s′ ) = x ∈
/ 𝐡 ′ . Thus 𝑐 is defined on 𝑀((Fπ‘‘π‘ž βˆ–πΌ)×(Fπ‘‘π‘ž βˆ–πΌ ′ )). Note that for each s ∈ Fπ‘‘π‘ž βˆ–πΌ, the
decoding subspace 𝑀(s, *) is good and hence contained in a good testing subspace 𝑣s ∈ 𝒯𝐷 ,
therefore 𝑐𝑣s |𝑀(s,*) ∈ π’ž. Similarly, for each s′ ∈ Fπ‘‘π‘ž βˆ– 𝐼 ′ , the decoding subspace 𝑀(*, s′ ) is
contained in a good testing subspace 𝑣s′ ∈ 𝒯𝐷 , hence 𝑐𝑣s′ |𝑀(*,s′ ) ∈ π’ž. Since |𝐼|, |𝐼 ′ | < 12 π›Ώπ‘ž 𝑑 , by
Proposition A.2.5, 𝑐 can be extended uniquely into 𝑐𝑀 ∈ π’ž ⊗2 defined on 𝑀. Define 𝑐x , 𝑐𝑀 |𝑒 .
Note that 𝑐x ∈ π’ž since it is the restriction of 𝑐𝑀 ∈ π’ž ⊗2 to 𝑒 = 𝑀(*, 0). Also, if s ∈ Fπ‘‘π‘ž βˆ– 𝐼, then
𝑐|𝑀(s,*) = 𝑐𝑣s |𝑀(s,*) = 𝑐𝑀 |𝑀(s,*) and in particular, 𝑐(𝑀(s, 0)) = 𝑐𝑀 (𝑀(s, 0)) = 𝑐x (𝑀(s, 0)). So
𝛿(𝑐, 𝑐x ) ≤
|𝐼|
π‘žπ‘‘
< 2𝛿 . Finally, since 0 ∈
/ 𝐼, 𝐼 ′ , we have 𝑐(x) = 𝑐(𝑀(0, 0)) = 𝑐x (𝑀(0, 0)) = 𝑐x (x).
This proves the first part of the claim.
For the second part (showing 𝑐|𝑒 ∈ π’ž), fix some x0 ∈ 𝑒. For each x ∈ 𝑒, let 𝑐x be the
codeword guaranteed by the previous part. Then, for every x ∈ 𝑒, 𝛿(𝑐x0 , 𝑐x ) ≤ 𝛿(𝑐x0 , 𝑐|𝑒 ) +
𝛿(𝑐|𝑒 , 𝑐x ) < 𝛿, therefore 𝑐x0 = 𝑐x . Moreover, for all x ∈ 𝑒, 𝑐x0 (x) = 𝑐x (x) = 𝑐(x), so
𝑐|𝑒 = 𝑐x0 ∈ π’ž.
72
We proceed to define 𝑐(x) for π‘₯ ∈ 𝐡 ′ . For such an x, pick any decoding subspace
𝑒 ∈ π’Ÿπ· (x), extend 𝑐|π‘’βˆ–π΅ ′ to a codeword 𝑐𝑒 ∈ π’ž, and define 𝑐(x) , 𝑐𝑒 (x). We now argue that
this is well-defined. Suppose 𝑒1 , 𝑒2 ∈ π’Ÿπ· (x) in directions 𝐴1 , 𝐴2 ∈ 𝐷, respectively. We need
to show that 𝑐𝑒1 , 𝑐𝑒2 are well-defined and that 𝑐𝑒1 (x) = 𝑐𝑒2 (x). Let 𝑀 ∈ 𝒱𝐷2 be the unique
subspace containing 𝑒1 , 𝑒2 , so 𝑀 = (x, 𝐴1 ∪ 𝐴2 ). By Claim 6.1.15, in each direction 𝐴1 , 𝐴2 ,
there are less than 21 π›Ώπ‘ž 𝑑 decoding subspaces in that direction in 𝑀 which intersect 𝐡 ′ . In
particular, this implies that 𝑒1 , 𝑒2 each contain less than π›Ώπ‘ž 𝑑 points from 𝐡 ′ . By what we just
showed, there are sets 𝐽1 , 𝐽2 ⊆ Fπ‘‘π‘ž of size |𝐽1 |, |𝐽2 | > (1 − 𝛿)π‘ž 𝑑 such that the “sub-rectangle”
𝑅 , 𝑀(𝐽1 × π½2 ) contains no points from 𝐡 ′ , and therefore 𝑐 has already been defined on 𝑅.
By Claim 6.1.16, on each decoding subspace 𝑒 in 𝑅 in either direction 𝐴1 or 𝐴2 , 𝑐|𝑒 ∈ π’ž.
Applying Proposition A.2.5, we see that 𝑐|𝑅 can be uniquely extended to a tensor codeword
𝑐𝑀 ∈ π’ž ⊗2 on 𝑀, and this gives a way to extend 𝑐|𝑒𝑖 βˆ–π΅ ′ to the codeword 𝑐𝑒𝑖 , 𝑐𝑀 |𝑒𝑖 ∈ π’ž
for 𝑖 ∈ {1, 2}. Therefore, the extensions 𝑐𝑒1 , 𝑐𝑒2 agree on x since 𝑐𝑒1 (x) = 𝑐𝑀 (x) = 𝑐𝑒2 (x),
and moreover for each decoding subspace 𝑒𝑖 this extension is unique since each decoding
subspace has less than π›Ώπ‘ž 𝑑 points from 𝐡 ′ .
Now that we have defined 𝑐 : Fπ‘š
π‘ž → Fπ‘ž and have shown that 𝛿(π‘Ÿ, 𝑐) ≤ 𝜌
(οΈ€ |𝐷| )οΈ€
𝑛−1
, it only
𝑛
. Let 𝑒 ∈ π’Ÿπ· be a decoding subspace. If 𝑒 ∩ 𝐡 ′ = ∅, then
remains to show that 𝑐 ∈ π’žπ·
𝑐|𝑒 ∈ π’ž follows from Claim 6.1.16. If 𝑒 intersects 𝐡 ′ , by the way we defined 𝑐(x) for x ∈ 𝐡 ′ ,
we showed that for any decoding subspace 𝑒 through x ∈ 𝐡 ′ , 𝑐|𝑒 ∈ π’ž by extending 𝑐|π‘’βˆ–π΅ ′ to
a codeword.
(οΈ€ π‘š )οΈ€
Claim 6.1.17. Let β„“ ≤ 𝑑 be a natural number and 𝐴1 , . . . , 𝐴𝑛 ∈ Fπ‘‘π‘ž be such that their
(οΈ€ )οΈ€ π‘ž−𝑑
(οΈ€ π‘š )οΈ€
π‘ž −β„“
union is linearly independent. Then at least 1 − 𝑛2 π‘ž−1
− 𝑛 π‘ž−1
fraction of 𝐴 ∈ Fπ‘‘π‘ž satisfy
that 𝐴, 𝐴1 , . . . , 𝐴𝑛 is β„“-proper.
Proof. Let a1 , . . . , a𝑑 be the random elements comprising 𝐴. By a union bound, it suffices to
(οΈ€ [𝑛] )οΈ€
(︀⋃︀
)οΈ€
show for any 𝑆 ∈ 𝑛−2
that
𝐴
∪ 𝐴 is linearly independent with probability at least
𝑖
𝑖∈𝑆
)οΈ€
(︀⋃︀
(οΈ€
)οΈ€
−𝑑
[𝑛]
π‘ž
and for any 𝑇 ∈ 𝑛−1
the probability that
1 − π‘ž−1
𝑖∈𝑇 𝐴𝑖 ∪ 𝐴 contains at least π‘š − β„“
−β„“
π‘ž
linearly independent elements is at least 1 − π‘ž−1
.
(οΈ€ [𝑛] )οΈ€
(︀⋃︀
)οΈ€
Fix 𝑆 ∈ 𝑛−2 .For any 𝑗 ∈ [𝑑], the probability that a𝑗 ∈ Fπ‘š
π‘ž is in the span of
𝑖∈𝑆 𝐴𝑖 ∪
73
{a1 , . . . , a𝑗−1 }, conditioned on the event that the latter is linearly independent, is
(︀⋃︀
)οΈ€
π‘ž −2𝑑+𝑗−1 . So the probability that
𝑖∈𝑆 𝐴𝑖 ∪ 𝐴 is linearly independent is
𝑑
∏︁
(οΈ€
1−π‘ž
−2𝑑+𝑗−1
)οΈ€
≥ 1−
𝑗=1
𝑑
∑︁
π‘ž π‘š−2𝑑+𝑗−1
π‘žπ‘š
=
π‘ž −2𝑑+𝑗−1
𝑗=1
≥ 1−π‘ž
−𝑑
∞
∑︁
π‘ž −𝑗
𝑗=1
−𝑑
= 1−
Now fix 𝑇 ∈
(οΈ€ [𝑛] )οΈ€
𝑛−1
π‘ž
.
π‘ž−1
. Similarly, The probability that π‘Žπ‘— ∈
(︀⋃︀
𝑖∈𝑇
)οΈ€
𝐴𝑖 ∪ {a1 , . . . , a𝑗−1 },
condition on the event that the later linearly independent is π‘ž −𝑑+𝑗−1 .
(︀⋃︀
)οΈ€
𝑖∈𝑆 𝐴𝑖 ∪ {a1 , . . . , a𝑑−β„“ } are linearly independent is
So we get that
𝑑−β„“
𝑑−β„“
∏︁
∑︁
(οΈ€
)οΈ€
1 − π‘ž −𝑑+𝑗−1 ≥ 1 −
π‘ž −𝑑+𝑗−1
𝑗=1
𝑗=1
≥ 1 − π‘ž −β„“
∞
∑︁
π‘ž −𝑗
𝑗=1
−β„“
= 1−
6.1.4
π‘ž
.
π‘ž−1
Robustness for Large Dimension
In this section, we prove our main result of the chapter:
Theorem 6.1.18. Let 𝜌 , E𝑣 [𝛿(π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 )], where 𝑣 is a random affine subspace of dimension 2𝑑. Let 𝛼1 be the 2𝑑-dimensional robustness of π’ž 𝑑↗4𝑑 given by Corollary 6.1.10. If
(οΈ€
)οΈ€
3
1𝛿
𝜌 < 𝛼400
− 3π‘ž −𝑑 , then 𝜌 ≥ 1 − 4𝛿 · 𝛿(π‘Ÿ, π’ž π‘‘β†—π‘š ). In particular, π’ž π‘‘β†—π‘š is (𝛼2 , 2𝑑)-robust, where
𝛼2 ≥
𝛿 72
.
2·1052
74
Notation. Throughout Section 6.1.4, fix the received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž and define 𝜌 ,
E𝑣 [𝛿(π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 )], and we will assume that 0 < 𝜌 <
𝛼1 𝛿 3
400
− 3π‘ž −𝑑 . The case where
𝛼1 𝛿 3
400
> 3π‘ž −𝑑
is easily dealt with at the end of the proof by using Corollary 6.1.2. Note that, since
𝛼1 , 𝛿 ≤ 1, this implies π‘ž −𝑑 ≤
𝛿
.
1200
Throughout this section we will assume π‘š ≥ 4𝑑. If
^
π‘š < 4𝑑 we can pad the function 𝑓 to get a function 𝑓^ : F4𝑑
π‘ž → Fπ‘ž (by setting 𝑓 (x, y) = 𝑓 (x)
4𝑑−π‘š
) and applying our tester to 𝑓^. We will typically use
for every x ∈ Fπ‘š
π‘ž and y ∈ Fπ‘ž
𝑒, 𝑣, 𝑀 to denote affine subspaces of dimension 𝑑, 2𝑑, and 4𝑑 respectively. For any affine
𝑑↗dim(𝐴)
subspace 𝐴 ⊆ Fπ‘š
be the codeword nearest to π‘Ÿ|𝐴 , breaking ties arbitrarily.
π‘ž , let 𝑐𝐴 ∈ π’ž
Let 𝜌𝐴 , E𝑣⊆𝐴 [𝛿(π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 )], where the expectation is taken over uniformly random 2𝑑dimensional subspaces 𝑣 ⊆ 𝐴. Fix the following constants:
𝛼1 𝛿 2
− 𝛼1 π‘ž −𝑑
40
𝜌 + 2π‘ž −𝑑
πœ– ,
.
𝛾
𝛾 ,
In particular, these constants are chosen so that the following bounds hold:
𝛿
2
𝛿
πœ– ≤
.
10
20𝛿 −1 (𝛼1−1 𝛾 + π‘ž −𝑑 ) ≤
Overview. This proof is a straightforward generalization of “bootstrapping” proofs originating in the work of Rubinfeld and Sudan [RS96] and which also appears in [ALM+ 98,
AS03, Aro94]. Our writeup in particular follows [Aro94]. For simplicity, assume 𝑑 = 1.
Our approach is to define a function 𝑐 : Fπ‘š
π‘ž → Fπ‘ž and then show that it is both close to
π‘Ÿ and a codeword of π’ž 1β†—π‘š . The definition of 𝑐 is simple: for every x ∈ Fπ‘š
π‘ž , consider the
opinion 𝑐𝑒 (x) for every line 𝑒 through x, and define 𝑐(x) as the majority opinion. We need
to show that 𝑐 is well-defined (the majority is actually a majority). Our main technical
lemma (Lemma 6.1.21) of this section shows that most lines agree with each other, so 𝑐 is
well-defined. Lemma 6.1.21 uses Claim 6.1.19, which shows that for a 4-dimensional affine
75
subspace 𝑀, if πœŒπ‘€ is small, then for every x ∈ 𝑀, most lines 𝑒 ⊆ 𝑀 satisfy 𝑐𝑒 (x) = 𝑐𝑀 (x).
To prove Claim 6.1.19 we use the results of Section 6.1.2, in particular the robustness of
the plane test in π‘š = 4 dimensions (Corollary 6.1.10). Since the average 𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 ) over 𝑒
through x is about 𝛿(π‘Ÿ|𝑀 , 𝑐𝑀 ), by robustness this is less than 𝛼1−1 πœŒπ‘€ , which is small since πœŒπ‘€
is small. Therefore, for most 𝑒, 𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 ) is small and so it must be that 𝑐𝑒 = 𝑐𝑀 |𝑒 .
Once we have shown that 𝑐 is well-defined, showing that 𝑐 is close to π‘Ÿ requires just a bit
of calculation. Showing that 𝑐 ∈ π’ž 1β†—π‘š involves more work. For each line 𝑒, define 𝑐′𝑒 ∈ π’ž
to be the nearest codeword to 𝑐|𝑒 . Fix a line 𝑒 and a point x ∈ 𝑒. We want to show that
𝑐|𝑒 (x) = 𝑐′𝑒 (x). The idea is to show the existence of a “good” 4-dimensional 𝑀 ⊇ 𝑒 such that
πœŒπ‘€ is small and for more than 1 − 2𝛿 fraction of points y ∈ 𝑒 (including x) are “good” in the
sense that 𝑐(y) = 𝑐𝑒′ (y) for a non-negligible fraction of lines 𝑒′ through y. Once we have such
a 𝑀, we show that for every good y ∈ 𝑒, 𝑐(y) = 𝑐𝑀 (y). Since 𝑒 has more than 1 − 2𝛿 fraction
good points, this implies that 𝛿(𝑐|𝑒 , 𝑐𝑀 |𝑒 ) < 2𝛿 , hence 𝑐′𝑒 = 𝑐|𝑀 , so 𝑐′𝑒 (x) = 𝑐|𝑀 (x) = 𝑐(x), as
desired.
Claim 6.1.19. If 𝑀 ⊆ Fπ‘š
π‘ž be a 4𝑑-dimensional affine subspace with πœŒπ‘€ ≤ 𝛾, then for every
x ∈ 𝑀, at least 1 −
𝛿
20
fraction of 𝑑-dimensional subspaces 𝑒 ⊆ 𝑀 satisfy 𝑐𝑒 (x) = 𝑐𝑀 (x).
Proof. Fix x ∈ 𝑀. Let π‘ˆ be the set of 𝑑-dimensional subspaces 𝑒 containing x such that
𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 ) < 20𝛿 −1 (𝛼1−1 πœŒπ‘€ +π‘ž −𝑑 ). By Corollary 6.1.10, E𝑒⊆𝑀 [𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 )] ≤ 𝛿(π‘Ÿ|𝑀 , 𝑐𝑀 )+π‘ž −𝑑 ≤
𝑒∋x
𝛼1−1 πœŒπ‘€ +π‘ž −𝑑 , so by Markov’s inequality, the probability that 𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 ) ≥ 20𝛿 −1 (𝛼1−1 πœŒπ‘€ +π‘ž −𝑑 )
is at most
−𝑑
𝛼−1
1 πœŒπ‘€ +π‘ž
−1
−1
20𝛿 (𝛼1 πœŒπ‘€ +π‘ž −𝑑 )
=
𝛿
.
20
For 𝑒 ∈ π‘ˆ , since 𝛿(π‘Ÿ|𝑒 , 𝑐𝑀 |𝑒 ) < 20𝛿 −1 (𝛼1−1 πœŒπ‘€ + π‘ž −𝑑 ) ≤
𝛿
2
and
𝑐𝑀 |𝑒 ∈ π’ž, we have 𝑐𝑒 = 𝑐𝑀 |𝑒 and therefore 𝑐𝑒 (x) = 𝑐𝑀 (x).
The following claim says that E𝑀 [πœŒπ‘€ ] ≈ 𝜌, even if we insist that 𝑀 contains a fixed
𝑑-dimensional subspace.
−𝑑
Claim 6.1.20. For any 𝑑-dimensional affine subspace 𝑒 ⊆ Fπ‘š
π‘ž , E𝑀⊇𝑒 [πœŒπ‘€ ] ≤ 𝜌 + 2π‘ž , where
𝑀 is a random 4𝑑-dimensional affine subspace containing 𝑒. In particular, for any point
−𝑑
x ∈ Fπ‘š
π‘ž , E𝑀∋x [πœŒπ‘€ ] ≤ 𝜌 + 2π‘ž .
76
Proof. Observe that
)οΈ€]οΈ€
[οΈ€ (οΈ€
𝜌 = E𝑣 𝛿 π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑
)οΈ€]οΈ€
[οΈ€ (οΈ€
≥ E𝑣:𝑒∩𝑣=∅ 𝛿 π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 · Pr[𝑒 ∩ 𝑣 = ∅]
𝑣
)οΈ€
)οΈ€]οΈ€ (οΈ€
[οΈ€ (οΈ€
𝑑↗2𝑑
· 1 − π‘ž −(π‘š−3𝑑) .
(Lemma B.2.1) ≥ E𝑣:𝑒∩𝑣=∅ 𝛿 π‘Ÿ|𝑣 , π’ž
Therefore,
[οΈ€
[οΈ€
]οΈ€]οΈ€
E𝑀⊇𝑒 [πœŒπ‘€ ] = E𝑀⊇𝑒 E𝑣⊆𝑀 𝛿(π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 )
[οΈ‚
]οΈ‚
βƒ’
[οΈ€
]οΈ€
𝑑↗2𝑑 βƒ’
≤ E𝑀⊇𝑒 E𝑣⊆𝑀 𝛿(π‘Ÿ|𝑣 , π’ž
) 𝑒 ∩ 𝑣 = ∅ + Pr [𝑒 ∩ 𝑣 ΜΈ= ∅]
𝑣⊆𝑀
βƒ’
[οΈ€
[οΈ€
]οΈ€]οΈ€
𝑑↗2𝑑 βƒ’
(Lemma B.2.1) ≤ E𝑀⊇𝑒 E𝑣⊆𝑀 𝛿(π‘Ÿ|𝑣 , π’ž
) 𝑒 ∩ 𝑣 = ∅ + π‘ž −𝑑
= E𝑣:𝑒∩𝑣=∅ [𝛿(π‘Ÿ|𝑣 , π’ž 𝑑↗2𝑑 )] + π‘ž −𝑑
𝜌
≤
+ π‘ž −𝑑
−(π‘š−3𝑑)
1−π‘ž
≤ 𝜌 + 2π‘ž −𝑑
Lemma 6.1.21 (Main). For every x ∈ Fπ‘š
π‘ž , there is a collection π‘ˆ1 of at least 1 −
𝛿
5
−
𝛿
600
fraction of the 𝑑-dimensional affine subspaces through x, such that 𝑐𝑒 (x) = 𝑐𝑒′ (x) for every
𝑒, 𝑒′ ∈ π‘ˆ1 .
Proof. Let π‘ˆ be the set of all 𝑑-dimensional affine subspaces 𝑒 through x. Partition π‘ˆ into
disjoint collections π‘ˆ1 , . . . , π‘ˆπ‘˜ with |π‘ˆ1 | ≥ · · · ≥ |π‘ˆπ‘˜ | according to the value of 𝑐𝑒 (x). We will
show that Pr𝑒∋π‘₯ [𝑒 ∈ π‘ˆ1 ] ≥ 1 −
𝛿
5
−
𝛿
.
600
For every 4𝑑-dimensional subspace 𝑀, let π‘ˆπ‘€ be the
collection of 𝑑-dimensional subspaces 𝑒 through x, guaranteed by Claim 6.1.19, satisfying
77
𝑐𝑒 (x) = 𝑐𝑀 (x). Then
Pr [𝑒 ∈ π‘ˆ1 ] ≥
𝑒∋x
=
(Lemma B.2.2) ≥
Pr [∃𝑖 𝑒, 𝑒′ ∈ π‘ˆπ‘– ]
𝑒,𝑒′ ∋x
Pr [𝑐𝑒 (x) = 𝑐𝑒′ (x)]
𝑒,𝑒′ ∋x
[𝑐𝑒 (x) = 𝑐𝑒′ (x)] − π‘ž −(π‘š−2𝑑)
𝑒∩𝑒 ={x}
⎑
⎀
Pr
′
= E𝑀∋x ⎣
Pr
′
[𝑐𝑒 (x) = 𝑐𝑒′ (x)]⎦ − π‘ž −(π‘š−2𝑑)
𝑒,𝑒 ⊆𝑀
𝑒∩𝑒′ ={x}
⎑
⎀
(Lemma B.2.2) ≥ E𝑀∋x ⎣ Pr
[𝑐𝑒 (x) = 𝑐𝑒′ (x)⎦ − π‘ž −2𝑑 − π‘ž −(π‘š−2𝑑)
′
𝑒,𝑒 ⊆𝑀
𝑒,𝑒′ ∋x
⎀
⎑
𝛿
[𝑐𝑒 (x) = 𝑐𝑒′ (x)⎦ −
≥ E𝑀∋x ⎣ Pr
′
𝑒,𝑒 ⊆𝑀
600
𝑒,𝑒′ ∋x
βƒ’
⎀
⎑
βƒ’
βƒ’
𝛿
≥ E𝑀∋x ⎣ Pr
[𝑐𝑒 (x) = 𝑐𝑒′ (x)] βƒ’βƒ’ πœŒπ‘€ ≤ 𝛾 ⎦ · Pr [πœŒπ‘€ ≤ 𝛾] −
′
𝑀∋x
𝑒,𝑒 ⊆𝑀
600
βƒ’
𝑒,𝑒′ ∋x
βƒ’
⎑
⎀
βƒ’
βƒ’
𝛿
≥ E𝑀∋x ⎣ Pr
[𝑒, 𝑒′ ∈ π‘ˆπ‘€ ] βƒ’βƒ’ πœŒπ‘€ ≤ 𝛾 ⎦ · Pr [πœŒπ‘€ ≤ 𝛾] −
′
𝑀∋x
𝑒,𝑒 ⊆𝑀
600
βƒ’
𝑒,𝑒′ ∋x
(οΈ‚
)οΈ‚2
𝛿
𝛿
(Claim 6.1.19) ≥
1−
· Pr [πœŒπ‘€ ≤ 𝛾] −
𝑀∋x
20
600
(οΈ‚
)οΈ‚2 (οΈ‚
)οΈ‚
E𝑀∋x [πœŒπ‘€ ]
𝛿
𝛿
· 1−
(Markov) ≥
1−
−
20
𝛾
600
(οΈ‚
)οΈ‚2 (οΈ‚
)οΈ‚
𝛿
𝜌 + 2π‘ž −𝑑
𝛿
(Claim 6.1.20) ≥
1−
· 1−
−
20
𝛾
600
−𝑑
𝛿
𝜌 + 2π‘ž
𝛿
≥ 1−
−
−
10
𝛾
600
𝛿
𝛿
= 1−
−πœ–−
10
600
𝛿
𝛿
≥ 1− −
5 600
We are now ready to prove the main theorem.
78
Proof of Theorem 6.1.18. We will define a function 𝑐 : Fπ‘š
π‘ž → Fπ‘ž and then show that it is
close to π‘Ÿ and is a codeword of π’ž π‘‘β†—π‘š . For x ∈ Fπ‘š
π‘ž , define 𝑐(x) , Majority𝑒∋x {𝑐𝑒 (x)}, where
𝛿
the majority is over 𝑑-dimensional affine subspaces 𝑒 through x. Since 5𝛿 + 600
< 12 , it follows
from Lemma 6.1.21 that 𝑐 is well-defined.
Next, we show that 𝑐 is close to π‘Ÿ. Indeed,
𝜌 = E𝑣 [𝛿(π‘Ÿ|𝑣 , 𝑐𝑣 )]
≥ E𝑒 [𝛿(π‘Ÿ|𝑒 , 𝑐𝑒 )]
[οΈ€
[οΈ€
]οΈ€]οΈ€
= E𝑒 Ex∈𝑒 1𝑐𝑒 (x)ΜΈ=π‘Ÿ(x)
[οΈ€
[οΈ€
]οΈ€]οΈ€
= Ex E𝑒∋x 1𝑐𝑒 (x)ΜΈ=π‘Ÿ(x)
[οΈ€
[οΈ€
]οΈ€
]οΈ€ βƒ’
≥ Ex E𝑒∋x 1𝑐𝑒 (x)ΜΈ=π‘Ÿ(x) βƒ’ 𝑐(x) ΜΈ= π‘Ÿ(x) · Pr[𝑐(x) ΜΈ= π‘Ÿ(x)]
βƒ’
[︁
]︁ x
βƒ’
≥ Ex Pr [𝑐𝑒 (x) = 𝑐(x)] βƒ’ 𝑐(x) ΜΈ= π‘Ÿ(x)] · 𝛿(π‘Ÿ, 𝑐)
𝑒∋x
)οΈ‚
(οΈ‚
𝛿
· 𝛿(π‘Ÿ, 𝑐).
(Lemma 6.1.21) ≥
1−
4
Finally, we show that 𝑐 ∈ π’ž π‘‘β†—π‘š . Let 𝑒 ⊆ Fπ‘š
π‘ž a 𝑑-dimensional affine subspace. We wish
to show that 𝑐|𝑒 ∈ π’ž. Let 𝑐′𝑒 ∈ π’ž be the codeword of π’ž nearest to 𝑐|𝑒 (not to be confused
with 𝑐𝑒 , the nearest codeword to π‘Ÿ|𝑒 ). Let x ∈ 𝑒. We will show that 𝑐′𝑒 (x) = 𝑐|𝑒 (x). For a
4𝑑-dimensional affine subspace 𝑀 ⊆ Fπ‘š
π‘ž , we say a point y ∈ 𝑀 is good for 𝑀 if Pr𝑒′ ⊆𝑀 [𝑐𝑒′ (y) =
𝑒′ ∋y
𝑐(y)] >
𝛿
.
20
We will show, by a union bound, that there exists a 4𝑑-dimensional affine
subspace 𝑀 ⊇ 𝑒 such that
1. πœŒπ‘€ ≤ 𝛾;
2. x is good for 𝑀;
3. more than 1 −
𝛿
2
fraction of points y ∈ 𝑒 are good for 𝑀.
Observe that for any y ∈ 𝑒, picking a random 4𝑑-dimensional 𝑀 containing 𝑒 and then
picking a random 𝑑-dimensional 𝑒′ ⊆ 𝑀 through y that intersect 𝑒 only on 𝑦 is equivalent to
picking a random 𝑑-dimensional 𝑒′ through y that intersect 𝑒 only on 𝑦 and then picking a
79
random 4𝑑-dimensional 𝑀 containing both 𝑒, 𝑒′ . Therefore, for any fixed y ∈ 𝑒
⎑
⎀
E𝑀⊇𝑒 ⎣ Pr
[𝑐𝑒′ (y) ΜΈ= 𝑐(y)]⎦ = E
′
[οΈ€
1𝑐𝑒′ (y)ΜΈ=𝑐(y)
≤ E
[οΈ€
1𝑐𝑒′ (y)ΜΈ=𝑐(y) | 𝑒 ∩ 𝑒′ = {y}
𝑀⊇𝑒
𝑒′ ⊆𝑀,𝑒′ ∋y
𝑒 ⊆𝑀
𝑒′ ∋y
𝑀⊇𝑒
𝑒′ ⊆𝑀,𝑒′ ∋y
+
Pr
𝑀⊇𝑒
𝑒′ ⊆𝑀,𝑒′ ∋y
]οΈ€
]οΈ€
[𝑒 ∩ 𝑒′ ΜΈ= {y}]
[οΈ€
]οΈ€
(Lemma B.2.2) ≤ E𝑒′ ∋y 1𝑐𝑒′ (y)ΜΈ=𝑐(y) | 𝑒 ∩ 𝑒′ = {𝑦} + π‘ž −2𝑑
[οΈ€
]οΈ€
(Lemma B.2.2) ≤ E𝑒′ ∋y 1𝑐𝑒′ (y)ΜΈ=𝑐(y) + π‘ž −(π‘š−2𝑑) + π‘ž −2𝑑
𝛿
𝛿
𝛿
𝛿
𝛿
+
+ 2π‘ž −2𝑑 ≤ +
≤ .
(Lemma 6.1.21 and definition of 𝑐) ≤
5 600
5 300
4
Therefore, by Markov’s inequality, for any fixed y ∈ 𝑒,
[οΈ‚
𝛿
Pr [y is not good for 𝑀] = Pr ′ Pr ′ [𝑐𝑒′ (y) ΜΈ= 𝑐(y)] ≥ 1 −
𝑀⊇𝑒
𝑀⊇𝑒 𝑒 ⊇𝑀,𝑒 ∋y
20
]οΈ‚
𝛿
4
≤
𝛿
1 − 20
5
≤
· 𝛿.
19
In particular, this applies for y = x. Further applying Markov’s inequality, we find that
[οΈ‚
]οΈ‚
10
𝛿
5𝛿/19
Pr fraction of not good y in 𝑒 ≥
= .
≤
𝑀⊇𝑒
2
𝛿/2
19
Finally, since E𝑀⊇𝑒 [πœŒπ‘€ ] ≤ 𝜌 + 2π‘ž −𝑑 (by Claim 6.1.20), we have
𝜌 + 2π‘ž −𝑑
𝛿
Pr [πœŒπ‘€ > 𝛾] ≤
=πœ–≤ .
𝑀⊇𝑒
𝛾
10
Since 𝛿 ≤ 1 and
5
19
+
10
19
+
1
10
< 1, by the union bound such a desired 𝑀 exists.
Now that we have such a subspace 𝑀, consider 𝑐𝑀 . We claim that it suffices to prove
that if y ∈ 𝑒 is good, then 𝑐𝑀 (y) = 𝑐(y). Indeed, since more than 1 −
in 𝑒 are good, we have 𝛿(𝑐𝑀 |𝑒 , 𝑐|𝑒 ) <
𝛿
.
2
𝛿
2
fraction of points
Therefore 𝑐𝑀 |𝑒 = 𝑐′𝑒 , and since x is good, we
80
have 𝑐(x) = 𝑐𝑀 (x) = 𝑐′𝑒 (x) as desired. It remains to prove that 𝑐𝑀 (y) = 𝑐(y) for good
y ∈ 𝑒. By Claim 6.1.19, at least 1 −
𝛿
20
fraction of 𝑑-dimensional 𝑒′ ⊆ 𝑀 through y satisfy
𝑐𝑒′ (y) = 𝑐𝑀 (y). Since y is good, more than
𝛿
20
fraction of 𝑑-dimensional 𝑒′ ⊆ 𝑀 through y
satisfy 𝑐𝑒′ (y) = 𝑐(y). Therefore, there must be some 𝑑-dimensional 𝑒′ ⊆ 𝑀 through y which
satisfies 𝑐𝑀 (y) = 𝑐𝑒′ (y) = 𝑐(y).
Finally, for the robustness statement: if π‘ž −𝑑 ≥
is at least
6.2
π‘ž −3𝑑
2
≥
𝛿 72
.
2·1052
𝛿 24
,
1014
then by Corollary 6.1.2, the robustness
Otherwise, the robustness is at least
𝛼1 𝛿 3
57,600·400
− 3π‘ž −𝑑 ≥
𝛿 24
.
2·1014
Technical Algebraic Results
The purpose of this section is to prove Theorem 6.2.6 and its Corollaries 6.2.7 and 6.2.8. If
we allow our robustness in Theorem 6.1.18 to depend on 𝑑, the dimension of the base code,
then proving what we need for Theorem 6.2.6 is easy. However, removing the dependence
on 𝑑 requires some new ideas, including the definition of a new operation (“degree lifting”)
on codes, and the analysis of the distance of degree lifted codes. In Section 6.2.1, we define
degree lifting and analyze the degree lifted codes (Proposition 6.2.4). In Section 6.2.2, we
prove Theorem 6.2.6 and its corollaries.
6.2.1
Degree Lift
In this section, we define the degree lift operation on codes with degree sets. The operation
can be thought of as “Reed-Mullerization”, in the sense that the degree lift of the ReedSolomon code of degree 𝑑 is the Reed-Muller code of degree 𝑑. This resembles the degree lift
operation of Ben-Sasson et al. [BGK+ 13] who defined a “Reed-Mullerization” for algebraicgeometry codes (in contrast, we want to define it for codes over Fπ‘š
π‘ž spanned by monomials).
Definition 6.2.1 (Degree lift). Let π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } have degree set Deg(π’ž). For positive
integer 𝑠 ≥ 1, define the 𝑠-wise degree lift π’ž(𝑠) ⊆ {Fπ‘šπ‘ 
→ Fπ‘ž } of π’ž to be the code with
π‘ž
81
degree set
{οΈƒ
Deg(π’ž(𝑠)) ,
(d1 , . . . , d𝑠 ) ∈ {0, 1, . . . , π‘ž − 1}π‘š×𝑠
βƒ’
}οΈƒ
𝑠
βƒ’ ∑︁
βƒ’
d𝑗 ∈ Deg(π’ž) .
βƒ’
βƒ’
𝑗=1
Our goal with this definition is to prove Proposition 6.2.4, which says that the distance
of π’ž(𝑠) is nearly the same as the distance of π’ž. One can show that 𝛿(π’ž(𝑠)) ≥ 𝛿(π’ž) − π‘šπ‘ž −1
To do so, we will use the following fact.
Proposition 6.2.2. Let 𝑑, 𝑛 ≥ 1 and let π‘š = 𝑛𝑑. Let π’ž0 ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affineinvariant and let π’ž , (π’ž0 )⊗𝑛 . For each 𝑖 ∈ [𝑛], let X𝑖 = (𝑋𝑖1 , . . . , 𝑋𝑖𝑑 ). If 𝑓 (X1 , . . . , X𝑛 ) ∈ π’ž,
and 𝐴1 , . . . , 𝐴𝑛 : Fπ‘‘π‘ž → Fπ‘‘π‘ž are affine transformations, then 𝑓 (𝐴1 (X1 ), . . . , 𝐴𝑛 (X𝑛 )) ∈ π’ž.
Proof. By linearity, it suffices to consider the case where 𝑓 (X1 , . . . , X𝑛 ) =
∏︀𝑛
𝑖=1
X𝑖 d𝑖 is a
monomial, where d𝑖 = (𝑑𝑖1 , . . . , 𝑑𝑖𝑑 ) ∈ {0, 1, . . . , π‘ž − 1}𝑑 . Each X𝑖 d𝑖 ∈ π’ž0 , so by affineinvariance 𝐴𝑖 (X𝑖 )d𝑖 ∈ π’ž0 . Therefore, by Proposition A.2.2, 𝑓 (𝐴1 (X1 ), . . . , 𝐴)𝑛 (X𝑛 )) =
∏︀𝑛
d𝑖
∈ (π’ž0 )⊗𝑛 = π’ž.
𝑖=1 𝐴𝑖 (X𝑖 )
Overview. To prove Proposition 6.2.4, we show, through Lemma 6.2.3, that there is a
special subset of π‘š-dimensional subspaces 𝐴, such that for any 𝑓 ∈ π’ž(𝑠), 𝑓 |𝐴 ∈ π’ž. Then,
we analyze the distance of from 𝑓 to the zero function by looking at the distance on a
random special π‘š-dimensional 𝐴. This will yield a distance of 𝛿(π’ž(𝑠)) ≥ 𝛿(π’ž) − π‘œ(1) as
long as the special subspaces sample Fπ‘š
π‘ž well. However, we require the π‘œ(1) term to be
(π‘›π‘ž −𝑑 )𝑂(1) , otherwise we would not be able to remove the dependence on 𝑑 in the robustness
of Theorem 6.1.18. In order to do so, we need to further assume that π’ž is the tensor product
(π’ž0 )𝑛 of some 𝑑-dimensional code π’ž0 (which is satisfied by our use case).
We now describe the special subspaces we consider in Lemma 6.2.3. Label the variables of
𝑛𝑑𝑠
Fπ‘šπ‘ 
by 𝑋𝑐𝑖𝑗 , where 𝑐 ∈ [𝑛], 𝑖 ∈ [𝑑], 𝑗 ∈ [𝑠]. Let π‘Œπ‘π‘– , for 𝑐 ∈ [𝑛], 𝑖 ∈ [𝑑], be the variables
π‘ž = Fπ‘ž
parameterizing 𝐴. Note that an arbitrary subspace restriction corresponds to substituting,
for each 𝑋𝑐𝑖𝑗 , an affine function of all of the variables π‘Œ11 , . . . , π‘Œπ‘›π‘‘ . This is too much to hope
for. However, if we substitute for 𝑋𝑐𝑖𝑗 an affine function of just π‘Œπ‘1 , . . . , π‘Œπ‘π‘‘ , this works.
82
Lemma 6.2.3. Let 𝑑, 𝑛 ≥ 1 and π‘š = 𝑛𝑑. Let π’ž0 ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant
and let π’ž , (π’ž0 )⊗𝑛 . Let 𝑠 ≥ 1, and let 𝑓 (X) ∈ π’ž(𝑠), with variables X = (𝑋𝑐𝑖𝑗 )𝑐∈[𝑛],𝑖∈[𝑑],𝑗∈[𝑠] .
Let 𝑔(π‘Œ11 , . . . , π‘Œπ‘›π‘‘ ) be the π‘š-variate polynomial obtained from 𝑓 (X) by setting, for each 𝑐 ∈
∑οΈ€
[𝑛], 𝑖 ∈ [𝑑], and 𝑗 ∈ [𝑠], 𝑋𝑐𝑖𝑗 = π‘‘π‘˜=1 π‘Žπ‘π‘–π‘—π‘˜ π‘Œπ‘π‘˜ + 𝑏𝑐𝑖𝑗 , for some π‘Žπ‘π‘–π‘—π‘˜ , 𝑏𝑐𝑖𝑗 ∈ Fπ‘ž . That is, for all
(𝑐, 𝑖, 𝑗) ∈ [𝑛] × [𝑑] × [𝑠] 𝑋𝑐𝑖𝑗 is an affine function of π‘Œπ‘1 , . . . , π‘Œπ‘π‘‘ . Then 𝑔 ∈ π’ž.
Proof. By linearity, it suffices to consider the case where 𝑓 (X) =
∏︀𝑛 ∏︀𝑑
𝑐=1
𝑖=1
∏︀𝑠
𝑗=1
𝑑
𝑋𝑐𝑖𝑗𝑐𝑖𝑗 is a
monomial, for some 0 ≤ 𝑑𝑐𝑖𝑗 ≤ π‘ž − 1. For each 𝑗 ∈ [𝑠], define d𝑗 , (𝑑11𝑗 , . . . , 𝑑𝑛𝑑𝑗 ), so that
∑οΈ€
(d1 , . . . , d𝑠 ) ∈ Deg(π’ž(𝑠)), i.e. 𝑠𝑖=1 d𝑖 ∈ Deg(π’ž). Then
𝑔(π‘Œ11 , . . . , π‘Œπ‘›π‘‘ ) =
(οΈƒ 𝑑
𝑛 ∏︁
𝑑 ∏︁
𝑠
∏︁
∑︁
𝑐=1 𝑖=1 𝑗=1
)︃𝑑𝑐𝑖𝑗
π‘Žπ‘π‘–π‘—π‘˜ π‘Œπ‘π‘˜ + 𝑏𝑐𝑖𝑗
π‘˜=1
𝑛 ∏︁
𝑑 ∏︁
𝑠
𝑑
∏︁
∑︁ (︂𝑑𝑐𝑖𝑗 )οΈ‚ 𝑒 ∏︁
π‘’π‘π‘–π‘—π‘˜ π‘’π‘π‘–π‘—π‘˜
𝑐𝑖𝑗0
=
𝑏𝑐𝑖𝑗
π‘Žπ‘π‘–π‘—π‘˜
π‘Œπ‘π‘˜
e𝑐𝑖𝑗
𝑐=1 𝑖=1 𝑗=1 e ≤𝑑
π‘˜=1
𝑐𝑖𝑗
=
∑︁
(· · · )
e𝑐𝑖𝑗 ≤𝑝 𝑑𝑐𝑖𝑗
∀ 𝑖,𝑗
𝑐𝑖𝑗
𝑛 ∏︁
𝑑
∏︁
∑︀𝑑
π‘Œπ‘π‘˜
𝑖=1
∑︀𝑠
𝑗=1 π‘’π‘π‘–π‘—π‘˜
𝑐=1 π‘˜=1
where the (· · · ) denotes constants in Fπ‘ž . So, it suffices to show that each monomial of the
∑︀𝑑 ∑︀𝑛
∏οΈ€ ∏οΈ€
π‘’π‘π‘–π‘—π‘˜
form 𝑛𝑐=1 π‘‘π‘˜=1 π‘Œπ‘π‘˜ 𝑖=1 𝑗=1
∈ π’ž, which we show in the remainder of the proof.
Let β„Ž(π‘Œ11 , . . . , π‘Œπ‘›π‘‘ ) be the π‘š-variate polynomial obtained from 𝑓 by substituting 𝑋𝑐𝑖𝑗 =
∏οΈ€ ∏οΈ€ ∏οΈ€
𝑑
π‘Œπ‘π‘– for each 𝑐 ∈ [𝑛], 𝑖 ∈ [𝑑], 𝑗 ∈ [𝑠]. Then β„Ž(π‘Œ11 , . . . , π‘Œπ‘›π‘‘ ) = 𝑛𝑐=1 𝑑𝑖=1 𝑠𝑗=1 π‘Œπ‘π‘– 𝑐𝑖𝑗 =
∑︀𝑠
(︁∑οΈ€
)︁
∏︀𝑛 ∏︀𝑑
∑︀𝑠
∑︀𝑠
𝑠
𝑗=1 𝑑𝑐𝑖𝑗
π‘Œ
is
a
monomial
with
degree
𝑑
,
.
.
.
,
𝑑
=
𝑐=1
𝑖=1 𝑐𝑖
𝑗=1 11𝑗
𝑗=1 𝑛𝑑𝑗
𝑗=1 d𝑗 ∈
Deg(π’ž), hence β„Ž ∈ π’ž. Now, consider applying an affine transformation as follows: for each
∑︀𝑑
1 ≤ 𝑐 ≤ 𝑛 and each 1 ≤ 𝑖 ≤ 𝑑, substitute π‘Œπ‘π‘– ←
π‘˜=1 π›Όπ‘π‘–π‘˜ π‘Œπ‘π‘˜ + 𝛽𝑐𝑖 , and call the new
83
polynomial β„Ž′ . By Proposition 6.2.2, β„Ž′ ∈ π’ž. On the other hand,
β„Ž′ (π‘Œ1 , . . . , π‘Œπ‘š ) =
(οΈƒ 𝑑
𝑛 ∏︁
𝑑
∏︁
∑︁
𝑐=1 𝑖=1
=
)οΈƒ∑︀𝑠𝑗=1 𝑑𝑐𝑖𝑗
π›Όπ‘π‘–π‘˜ π‘Œπ‘π‘˜ + 𝛽𝑐𝑖
π‘˜=1
(οΈƒ 𝑑
𝑛 ∏︁
𝑑 ∏︁
𝑠
∏︁
∑︁
𝑐=1 𝑖=1 𝑗=1
π‘˜=1
𝑛 ∏︁
𝑑 ∏︁
𝑠
∏︁
∑︁
)︃𝑑𝑐𝑖𝑗
π›Όπ‘π‘–π‘˜ π‘Œπ‘π‘˜ + 𝛽𝑐𝑖
)οΈ‚
𝑑
∏︁
𝑑𝑐𝑖𝑗
π‘’π‘π‘–π‘—π‘˜ π‘’π‘π‘–π‘—π‘˜
𝑒𝑐𝑖𝑗0
=
𝛽𝑐𝑖
π‘Œπ‘π‘˜
π›Όπ‘π‘–π‘˜
eπ‘π‘–π‘—π‘˜
𝑐=1 𝑖=1 𝑗=1 e𝑐𝑖𝑗 ≤𝑝 𝑑𝑐𝑖𝑗
π‘˜=1
)οΈƒ
(οΈƒ (οΈ‚
)οΈƒ 𝑛 𝑑 (οΈƒ 𝑑
∑︀𝑑 ∑︀𝑠
∑︁
∏︁ 𝑑𝑐𝑖𝑗 )οΈ‚ 𝑒
∏︁ ∏︁ ∏︁ ∑︀𝑠 π‘’π‘π‘–π‘—π‘˜
π‘’π‘π‘–π‘—π‘˜
𝑐𝑖𝑗0
𝑗=1
=
π‘Œπ‘π‘˜ 𝑖=1 𝑗=1
𝛽𝑐𝑖
π›Όπ‘π‘–π‘˜
eπ‘π‘–π‘—π‘˜
𝑐=1 π‘˜=1
𝑐,𝑖,𝑗
𝑖=1
e ≤ 𝑑
𝑐𝑖𝑗
(οΈ‚
𝑝 𝑐𝑖𝑗
∀ 𝑐,𝑖,𝑗
and since the π›Όπ‘π‘–π‘˜ and the 𝛽𝑐𝑖 are arbitrary and π’ž has a degree set Deg(π’ž) = Deg(π’ž0 )𝑛 , each
∑︀𝑑 ∑︀𝑛
∏οΈ€ ∏︀𝑑
𝑖=1
𝑗=1 π‘’π‘π‘–π‘—π‘˜
π‘Œ
∈ π’ž, as desired.
monomial 𝑛−
π‘˜=1 π‘π‘˜
𝑐=1
Proposition 6.2.4. Let 𝑑, 𝑛 ≥ 1 and π‘š = 𝑛𝑑. Let π’ž0 ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be linear affine-invariant
and let π’ž , (π’ž0 )⊗𝑛 . For any positive integer 𝑠 ≥ 1, 𝛿(π’ž(𝑠)) ≥ 𝛿(π’ž) − π‘›π‘ž −𝑑 = 𝛿(π’ž0 )𝑛 − π‘›π‘ž −𝑑 .
Proof. Let 𝑓 (X) ∈ π’ž(𝑠) be a nonzero codeword with variables X = (𝑋𝑐𝑖𝑗 )𝑐∈[𝑛],𝑖∈𝑑],𝑗∈[𝑠] . For
𝑐 ∈ [𝑛], 𝑖 ∈ [π‘š], 𝑗 ∈ [𝑠], π‘˜ ∈ [𝑑], let π‘Žπ‘π‘–π‘—π‘˜ , 𝑏𝑐𝑖𝑗 ∈ Fπ‘ž , and let a , (π‘Žπ‘π‘–π‘—π‘˜ )𝑐∈[𝑛],𝑖∈[𝑑],𝑗∈[𝑠],π‘˜∈[𝑑] and
b , (𝑏𝑐𝑖𝑗 )𝑐∈[𝑛],𝑖∈[𝑑],𝑗∈[𝑠] . Let 𝑔a,b (π‘Œ11 , . . . , π‘Œπ‘›π‘‘ ) be the π‘š-variate polynomial obtained by setting
∑οΈ€
𝑋𝑐𝑖𝑗 = π‘‘π‘˜=1 π‘Žπ‘π‘–π‘—π‘˜ π‘Œπ‘π‘˜ + 𝑏𝑐𝑖𝑗 for each 1 ≤ 𝑐 ≤ 𝑛 and 1 ≤ 𝑖 ≤ 𝑑.
By linearity of π’ž and thus of π’ž(𝑠), it suffices to show that 𝛿(𝑓, 0) ≥ 𝛿(π’ž) − π‘›π‘ž −𝑑 . Let
b ∈ F𝑛𝑑𝑠
be a point such that 𝑓 (b) ΜΈ= 0. Consider choosing a uniformly at random. Then
π‘ž
𝑔a,b ΜΈ= 0 since 𝑔a,b (0) = 𝑓 (b) ΜΈ= 0. For fixed 𝑦11 , . . . , 𝑦𝑛𝑑 , as long as for each 𝑐 ∈ [𝑛] there
∑οΈ€
is some π‘˜ ∈ [𝑑] such that π‘¦π‘π‘˜ ΜΈ= 0, then the points π‘‘π‘˜=1 π‘Žπ‘π‘–π‘—π‘˜ π‘¦π‘π‘˜ + 𝑏𝑐𝑖𝑗 are independent and
84
uniform over Fπ‘ž . This occurs with probability at least 1 − π‘›π‘ž −𝑑 . Therefore,
𝛿(π’ž) ≤ Ea [𝛿 (𝑔a,b , 0)]
[οΈ€ [οΈ€
]οΈ€]οΈ€
= Ea Ey 1𝑔a,b (y)ΜΈ=0
[οΈ€ [οΈ€
]οΈ€]οΈ€
= Ey Ea 1𝑔a,b (y)ΜΈ=0
[οΈ€
]οΈ€
≤ π‘›π‘ž −𝑑 + EyΜΈ=0 1𝑔a,b (y)ΜΈ=0
]︁
[︁
= π‘›π‘ž −𝑑 + EyΜΈ=0 1𝑓 ((∑οΈ€π‘‘π‘˜=1 π‘Žπ‘π‘–π‘—π‘˜ π‘¦π‘π‘˜ +𝑏𝑐𝑖𝑗 ))ΜΈ=0
= π‘›π‘ž −𝑑 + 𝛿(𝑓, 0).
6.2.2
Analysis of Subspace Restrictions
In this section we prove Theorem 6.2.6 and its corollaries.
Overview. Corollary 6.2.7 says that if a codeword 𝑓 of the tensor product π’ž ⊗𝑛 of a 𝑑dimensional code π’ž is not a codeword of π’ž 𝑑↗𝑛𝑑 , then on there is a point b such that on many
𝑑-dimensional subspaces 𝑒 through b, the restriction 𝑓 |𝑒 ∈
/ π’ž. We use this in the proof of
Theorem 6.1.4 when arguing that if a tensor codeword 𝑐 ∈ π’ž ⊗π‘š satisfies 𝑐 ∈ π’ža (see overview)
β‹‚οΈ€
for many a, then 𝑐 ∈ a π’ža = π’ž 1β†—π‘š . A special case of Corollary 6.2.8 says that if 𝑓 is a
lifted Reed-Solomon codeword but not a Reed-Muller codeword, then on many planes 𝑓 is
not a bivariate Reed-Muller code. The actual corollary merely generalizes this to arbitrary
𝑑 and codes π’ž0 , π’ž1 .
Both Corollaries 6.2.7 and 6.2.8 are proved in a similar manner. Note that both are statements of the form “if 𝑓 is in some big code but not in a lifted code, then on many subspaces
it is not a codeword of the base code”. A natural approach is to write 𝑓 out as a linear
combination of monomials, restrict to an arbitrary subspace of the appropriate dimension,
re-write the restriction as a linear combination of monomials in the parameterizing variables,
and note that the coefficients of the monomials are functions in the parameterization coeffi85
cients. Since 𝑓 is not in the lift, there is a monomial outside the base code whose coefficient
(the “offending coefficient”) is a nonzero function. Then, one shows that these functions
belong to a code with good distance, so for many choices of parameterizing coefficients, the
offending coefficient is nonzero.
Theorem 6.2.6 abstracts the above approach and shows that, in the case of Corollary 6.2.7,
the offending coefficient is a codeword of the degree lift (π’ž ⊗𝑛 )(𝑑) of π’ž ⊗𝑛 , and in the case of
Corollary 6.2.8, the offending coefficient is a codeword of a lifted code. This necessitates the
analysis of the distance of degree lifted codes, hence the need for Section 6.2.1.
Lemma 6.2.5. Let π’ž ⊆ {Fπ‘š
π‘ž → Fπ‘ž } be a linear code with a 𝑝-shadow-closed degree set. If
𝑓 ∈ π’ž, and
(οΈƒ
𝑓
π‘Ž10 +
𝑑
∑︁
π‘Ž1𝑗 π‘Œπ‘— , . . . , π‘Žπ‘š0 +
𝑗=1
𝑑
∑︁
)οΈƒ
π‘Žπ‘šπ‘— π‘Œπ‘—
π‘š(𝑑+1)
𝑓e (a) =
𝑓e (a) · Ye
e∈{0,1,...,π‘ž−1}𝑑
𝑗=1
where a = (π‘Žπ‘–π‘— )1≤𝑖≤π‘š;0≤𝑗≤𝑑 ∈ Fπ‘ž
∑︁
=
, then, for every e ∈ {0, 1, . . . , π‘ž − 1}𝑑 ,
∑︁
d∈𝐷
E≤𝑝 d
β€–E*𝑗 β€–=𝑒𝑗 ∀𝑗
𝑓d ·
)οΈ‚
π‘š (οΈ‚
∏︁
𝑑𝑖
𝑖=1
E𝑖*
π‘Žπ‘’π‘–0𝑖0
𝑑
∏︁
𝑒
π‘Žπ‘–π‘—π‘–π‘— .
𝑗=1
In particular,
1. 𝑓e ∈ π’ž(𝑑 + 1), the (𝑑 + 1)-wise degree lift of π’ž (see Definition 6.2.1);
2. if π’ž = (π’ž0 )1β†—π‘š for some linear affine-invariant code π’ž0 ⊆ {Fπ‘ž → Fπ‘ž }, then 𝑓e ∈
(π’ž0 )1β†—π‘š(𝑑+1)
Proof. Let 𝐷 be the degree set of π’ž. Write 𝑓 (X) =
86
∑οΈ€
d∈𝐷
𝑓d · Xd . Let 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž be the
(︁
)︁
∑οΈ€
∑οΈ€
affine map Y ↦→ π‘Ž10 + 𝑑𝑗=1 π‘Ž1𝑗 π‘Œπ‘— , . . . , π‘Žπ‘š0 + 𝑑𝑗=1 π‘Žπ‘šπ‘— π‘Œπ‘— . Expanding, we get
(𝑓 ∘ 𝐴)(Y) =
∑︁
d∈𝐷
𝑓d ·
π‘š
∏︁
(οΈƒ
π‘Žπ‘–0 +
𝑑
∑︁
𝑖=1
)︃𝑑𝑖
π‘Žπ‘–π‘— π‘Œπ‘—
𝑗=1
βŽ›
⎞
𝑑
∑︁ (︂𝑑𝑖 )οΈ‚
∏︁
𝑒
𝑒
⎝
=
𝑓d ·
π‘Žπ‘’π‘–0𝑖0 ·
π‘Žπ‘–π‘—π‘–π‘— π‘Œπ‘— 𝑖𝑗 ⎠
e𝑖
𝑖=1
𝑗=1
d∈𝐷
e𝑖 ≤𝑝 𝑑𝑖
(οΈƒ π‘š (οΈ‚ )οΈ‚
)οΈƒ 𝑑
𝑑
∑︁
∑︁ ∏︁ 𝑑𝑖
∏︁
∏︁ β€–E β€–
𝑒𝑖𝑗
𝑒𝑖0
=
𝑓d ·
π‘Žπ‘–0
π‘Žπ‘–π‘— ·
π‘Œπ‘— *𝑗
E𝑖*
𝑖=1
𝑗=1
𝑗=1
d∈𝐷
E≤𝑝 d
)οΈ‚
π‘š (οΈ‚
𝑑
∑︁
∑︁
∏︁
∏︁
𝑑𝑖
𝑒
e
=
Y ·
𝑓d ·
π‘Žπ‘’π‘–0𝑖0
π‘Žπ‘–π‘—π‘–π‘—
E
𝑖*
𝑑
𝑖=1
𝑗=1
d∈𝐷
∑︁
π‘š
∏︁
e∈{0,1,...,π‘ž−1}
E≤𝑝 d
β€–E*𝑗 β€– mod* π‘ž=𝑒𝑗 ∀𝑗
and therefore, for each e ∈ {0, 1, . . . , π‘ž − 1}𝑑 ,
𝑓e (a) =
∑︁
d∈𝐷
E≤𝑝 d
β€–E*𝑗 β€– mod* π‘ž=𝑒𝑗 ∀𝑗
)οΈ‚
π‘š (οΈ‚
𝑑
∏︁
∏︁
𝑑𝑖
𝑒
𝑒𝑖0
𝑓d ·
π‘Žπ‘–0
π‘Žπ‘–π‘—π‘–π‘— .
E𝑖*
𝑖=1
𝑗=1
View the variables a = (π‘Žπ‘–π‘— ) in the order (π‘Ž10 , . . . , π‘Žπ‘š0 , . . . , π‘Ž1𝑑 , . . . , π‘Žπ‘šπ‘‘ ), and interpret E
as (E*0 , . . . , E*𝑑 ). If E ≤𝑝 d and d ∈ 𝐷 = Deg(π’ž), then E ∈ Deg(π’ž(𝑑 + 1)). Therefore,
𝑓e ∈ π’ž(𝑑 + 1).
Now suppose π’ž = (π’ž0 )1β†—π‘š for some linear affine-invariant π’ž0 ⊆ {Fπ‘ž → Fπ‘ž }. It suffices to
show that if d = (𝑑1 , . . . , π‘‘π‘š ) ∈ Deg(π’ž) and E ≤𝑝 d with entries 𝑒𝑖𝑗 , 𝑖 ∈ [π‘š], 0 ≤ 𝑗 ≤ 𝑑, then
the length π‘š(𝑑 + 1) vector (E*0 , E*1 , . . . , E*𝑑 ) ∈ Deg((π’ž0 )1β†—π‘š(𝑑+1) ). By Proposition 5.2.3, it
∑οΈ€
suffices to show that, if 𝑒𝑖𝑗 ≤𝑝 𝑒𝑖𝑗 for every 𝑖 ∈ [π‘š] and 0 ≤ 𝑗 ≤ 𝑑, then 𝑖𝑗 𝑒𝑖𝑗 mod* π‘ž ∈
Deg(π’ž0 ). Since d ∈ Deg(π’ž) = Deg((π’ž0 )1β†—π‘š ), this implies that if 𝑒′𝑖 ≤𝑝 𝑑𝑖 for 𝑖 ∈ [π‘š], then
∑οΈ€ ′
∑︀𝑑
*
′
𝑒
mod
π‘ž
∈
Deg(π’ž
).
Set
𝑒
,
0
𝑖
𝑖
𝑖
𝑗=0 𝑒𝑖𝑗 . Observe that, since (𝑒𝑖0 , 𝑒𝑖1 , . . . , 𝑒𝑖𝑑 ) ≤𝑝 𝑑𝑖 , this
∑οΈ€
∑οΈ€
implies that 𝑒′𝑖 ≤𝑝 𝑑𝑖 . Therefore, 𝑖𝑗 𝑒𝑖𝑗 mod* π‘ž = 𝑖 𝑒′𝑖 mod* π‘ž ∈ Deg(π’ž0 ), as desired.
Theorem 6.2.6. Let 1 ≤ 𝑑 < π‘š. Let π’ž1 ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be a linear affine-invariant code, and
π‘‘β†—π‘š
let π’ž2 ⊆ {Fπ‘š
. Then the
π‘ž → Fπ‘ž } have a 𝑝-shadow-closed degree set. Suppose 𝑓 ∈ π’ž2 βˆ– π’ž1
87
following hold:
1. if π’ž2 = (π’ž0 )⊗𝑛 for some linear affine-invariant code π’ž0 ⊆ {Fπ‘‘π‘ž → Fπ‘ž }, where π‘š = 𝑛𝑑,
𝑛
−𝑑
then there exists a point b ∈ Fπ‘š
fraction of
π‘ž such that for at least 𝛿(π’ž0 ) − (𝑛 + 1)π‘ž
/ π’ž1 ;
𝑑-dimensional affine subspaces 𝐴 ⊆ Fπ‘š
π‘ž passing through b, the restriction 𝑓 |𝐴 ∈
2. if π’ž2 = (π’ž0 )1β†—π‘š for some linear affine-invariant code π’ž0 ⊆ {Fπ‘ž → Fπ‘ž }, then for at least
/ π’ž1 .
𝛿(π’ž0 ) − π‘ž −1 fraction of 𝑑-dimensional affine subspaces 𝐴 ⊆ Fπ‘š
π‘ž , the restriction 𝑓 |𝐴 ∈
Proof. Let 𝑝 be the characteristic of Fπ‘ž . Let 𝐴 be parameterized by 𝑋𝑖 = π‘Žπ‘–0 +
∑︀𝑑
𝑗=1
π‘Žπ‘–π‘— π‘Œπ‘— ,
π‘š×𝑑
has full rank. Write
where the matrix {π‘Žπ‘–π‘— }π‘š,𝑑
𝑖=1,𝑗=1 ∈ Fπ‘ž
𝑓 |𝐴 (Y) =
∑︁
𝑓e (a) · Ye .
e∈{0,1,...,π‘ž−1}𝑑
/ Deg(π’ž1 ) such that 𝑓e ΜΈ= 0.
Since 𝑓 ∈
/ π’ž1π‘š , there exists e ∈
1. By Corollary A.2.3, π’ž2 has a 𝑝-shadow-closed degree set. By Lemma 6.2.5 (1), 𝑓e ∈
π’ž2 (𝑑 + 1). For each b = (𝑏1 , . . . , π‘π‘š ) ∈ Fπ‘š
π‘ž , let 𝑓e,b denote the polynomial 𝑓e with
the variable π‘Žπ‘–0 fixed to value 𝑏𝑖 for each 𝑖 ∈ [π‘š] (i.e. insisting that 𝐴 passes through
b). Observe that each 𝑓e,b ∈ π’ž2 (𝑑). Since 𝑓e ΜΈ= 0, there exists b ∈ Fπ‘š
π‘ž such that
𝑓e,b ΜΈ= 0. By Proposition 6.2.4, for at least 𝛿(π’ž2 (𝑑)) ≥ 𝛿(π’ž0 )𝑛 − π‘›π‘ž −𝑑 fraction of matrices
{π‘Žπ‘–π‘— }𝑖∈[π‘š];𝑗∈[𝑑] , we have 𝑓e,b ({π‘Žπ‘–π‘— }𝑖∈[π‘š];𝑗∈[𝑑] ) ΜΈ= 0. Since, by Lemma B.2.2, at least 1−π‘ž 𝑑−π‘š
fraction of such matrices have full rank, we get that for at least 𝛿(π’ž0 )𝑛 − π‘›π‘ž −𝑑 − π‘ž 𝑑−π‘š ≥
(︁
)︁
𝛿(π’ž0 )𝑛 −(𝑛+1)π‘ž −𝑑 of the full rank matrices satisfy 𝑓e,b {π‘Žπ‘–π‘— }π‘š,𝑑
𝑖=1,𝑗=1 ΜΈ= 0 , and therefore
𝑓 |𝐴 (Y) ∈
/ π’ž1 .
2. By Proposition 4.1.3 it has a 𝑝-shadow-closed degree set. By Lemma 6.2.5 (2), 𝑓e ∈
(π’ž0 )1β†—π‘š(𝑑+1) , so 𝑓e (a) ΜΈ= 0 for at least 𝛿((π’ž0 )1β†—π‘š(𝑑+1) ) ≥ 𝛿(π’ž0 )−π‘ž −1 fraction of choices a
(including such that the corresponding matrix does not have full rank), and therefore,
by Lemma B.2.2, 𝑓 |𝐴 (Y) ∈
/ π’ž1 for at least 𝛿(π’ž0 ) − π‘ž −1 − π‘ž 𝑑−π‘š ≥ 𝛿(π’ž0 ) − 2π‘ž −1 .
88
Corollary 6.2.7. Let 𝑑, 𝑛 ≥ 1 and let π‘š = 𝑛𝑑. Let π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be a linear affine𝑛
−𝑑
invariant code. If 𝑓 ∈ π’ž ⊗𝑛 βˆ–π’ž π‘‘β†—π‘š , then there is a point b ∈ Fπ‘š
π‘ž such that for 𝛿(π’ž) −(𝑛+1)π‘ž
fraction of 𝑑-dimensional subspaces 𝑒 through b, the restriction 𝑓 |𝑒 ∈
/ π’ž.
Proof. Follows immediately from Theorem 6.2.6 (1) with π’ž0 = π’ž1 = π’ž, and π’ž2 = π’ž ⊗𝑛 .
Corollary 6.2.8. Let 1 ≤ 𝑑 ≤ π‘š. Let π’ž0 ⊆ {Fπ‘ž → Fπ‘ž } be a linear affine-invariant code.
Let π’ž1 ( (π’ž0 )1↗𝑑 be a linear affine-invariant code that is a strict subcode of (π’ž0 )1↗𝑑 . If
𝑓 ∈ (π’ž0 )1β†—π‘š βˆ– (π’ž1 )π‘‘β†—π‘š , then for at least 𝛿(π’ž0 ) − 2π‘ž −1 fraction of 𝑑-dimensional subspaces
/ π’ž1 .
𝐴 ⊆ Fπ‘š
π‘ž , the restriction 𝑓 |𝐴 ∈
Proof. Follows immediately from Theorem 6.2.6 (2) with π’ž2 = (π’ž0 )1β†—π‘š .
89
90
Chapter 7
Applications
7.1
Lifted Reed-Solomon Code
We begin by describing the central construction of the thesis — the lifted Reed-Solomon. As
its name suggests, the lifted Reed-Solomon is simply obtained by lifting the Reed-Solomon
code. With judicious choice of parameters, the lifted Reed-Solomon code is the first code
to achieve a combination of parameters never achieved by one code before. For now, we
describe our construction for an arbitrary choice of parameters.
Let π‘ž be a prime power, let π‘š ≥ 2 be an integer, and let 𝑑 < π‘ž. Let RS , RS(π‘ž, 𝑑) and let
π’ž , RS1β†—π‘š . The code π’ž is the lifted Reed-Solomon code. It automatically inherits distance, a
well-structured degree set, local correctability, decodability, and robust testability by virtue
of being a lifted code.
Proposition 7.1.1. The distance of π’ž is 𝛿(π’ž) ≥ 1 −
𝑑+1
.
π‘ž
Proof. By Proposition 5.2.4, 𝛿(π’ž) ≥ 𝛿(RS) − π‘ž −1 , and by Proposition 3.4.2, 𝛿(RS) = 1 − π‘‘π‘ž ,
so 𝛿(π’ž) ≥ 𝛿(RS) − π‘ž −1 ≥ 1 −
𝑑+1
.
π‘ž
Proposition 7.1.2. The code π’ž is linear affine-invariant and d ∈ Deg(π’ž) if and only if, for
∑οΈ€
*
every e ≤𝑝 d, the sum π‘š
𝑖=1 𝑒𝑖 mod π‘ž ≤ 𝑑.
Proof. Follows immediately from Proposition 5.2.3.
91
(οΈ€
)οΈ€
Proposition 7.1.3. For every πœ–, πœ‚ > 0, the code π’ž is 𝑄, ( 12 − πœ–)𝛿 − π‘ž −1 , πœ‚ -locally correctable
and decodable, where 𝛿 , 1 −
𝑑
π‘ž
and 𝑄 = 𝑂(ln(1/πœ‚)/πœ–2 ).
Proof. The local correctability follows immediately from Theorem 5.3.2 while local decodability follows from Theorem 5.3.4.
Proposition 7.1.4. Let 𝛿 , 1 − π‘‘π‘ž . The code π’ž has a π‘ž 2 -local tester that is
𝛿 72
-robust.
2·1052
Proof. Follows immediately from Theorem 5.4.3.
7.1.1
Relationship to Reed-Muller
We proceed by examining the relationship between the lifted Reed-Solomon code and the
Reed-Muller code. It follows immediately from definitions that the Reed-Muller code is
contained in the lifted Reed-Solomon code, i.e. RM(π‘ž, 𝑑, π‘š) ⊆ RS(π‘ž, 𝑑)1β†—π‘š . Kaufman and
Ron [KR06] proved the following characterization: a polynomial over Fπ‘š
π‘ž has degree 𝑑 if and
⌈︁
⌉︁
𝑑+1
only if on every 𝑑-dimensional affine subspace its restriction has degree 𝑑, where 𝑑 = π‘ž−π‘ž/𝑝
and 𝑝 is the characteristic of Fπ‘ž . This generalizes a result of Friedl and Sudan [FS95], which
says that if 𝑑 ≤ π‘ž − π‘ž/𝑝 and a polynomial has degree 𝑑 on every line, then its global degree
is 𝑑, and that this is not necessarily true if 𝑑 ≥ π‘ž − π‘ž/𝑝. Observe that, in the language of
lifting, this statement simply says that RS(π‘ž, 𝑑)1β†—π‘š = RM(π‘ž, 𝑑, π‘š) if and only if 𝑑 < π‘ž − π‘ž/𝑝.
In Theorem 7.1.6, we give a more efficient proof of the above result of [FS95] using the
technology of affine-invariance and lifting. In Theorem 7.1.8, we prove a specialization of
the characterization of [KR06]: if π‘ž − π‘ž/𝑝 ≤ 𝑑 < π‘ž, then polynomials of degree 𝑑 are
characterized by having degree 𝑑 on planes, i.e. RM(π‘ž, 𝑑, 2)2β†—π‘š = RM(π‘ž, 𝑑, π‘š), again using
affine-invariance and lifting.
First, we prove a handy lemma.
Lemma 7.1.5. Let π‘š ≥ 𝑑, let π‘ž be a prime power, and let 𝑑 < π‘ž. If RM(π‘ž, 𝑑, π‘š) =
RM(π‘ž, 𝑑, 𝑑)π‘‘β†—π‘š , then RM(π‘ž, 𝑑 − 1, π‘š) = RM(π‘ž, 𝑑 − 1, 𝑑)π‘‘β†—π‘š .
Proof. Let 𝑓 ∈ RM(π‘ž, 𝑑 − 1, 𝑑)π‘‘β†—π‘š . Define 𝑔 : Fπ‘š
π‘ž → Fπ‘ž by 𝑔(X) , 𝑋1 · 𝑓 (X). Since
𝑓 ∈ RM(π‘ž, 𝑑 − 1, 𝑑)π‘‘β†—π‘š , by Proposition 5.1.2 it follows that deg(𝑓 ∘ 𝐴) ≤ 𝑑 − 1 < π‘ž − 1
92
for every affine 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž Note that (𝑔 ∘ 𝐴)(Y) = 𝑔(𝐴(Y)) = (𝐴(Y))1 · (𝑓 ∘ 𝐴)(Y), so
π‘‘β†—π‘š
=
deg(𝑔 ∘ 𝐴) ≤ deg(𝑓 ∘ 𝐴) + 1 ≤ 𝑑 for every affine 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž , hence 𝑔 ∈ RM(π‘ž, 𝑑, 𝑑)
RM(π‘ž, 𝑑, π‘š), so deg(𝑔) ≤ 𝑑. Therefore, deg(𝑓 ) = deg(𝑔) − 1 ≤ 𝑑 − 1.
Theorem 7.1.6. Let 𝑝 be a prime and let π‘ž be a power of 𝑝. Let 𝑑 < π‘ž. Let π‘š ≥ 2. Then
RM(π‘ž, 𝑑, π‘š) = RS(π‘ž, 𝑑)1β†—π‘š if and only if 𝑑 < π‘ž − π‘ž/𝑝.
Proof. By Lemma 7.1.5, it suffices to show that equality holds for 𝑑 = π‘ž − π‘ž/𝑝 − 1 and does
not hold for 𝑑 = π‘ž − π‘ž/𝑝. Let 𝑑 = π‘ž − π‘ž/𝑝 − 1 = (1 − 𝑝−1 )π‘ž − 1. This corresponds to the
construction of π’ž , RS(π‘ž, 𝑑)1β†—π‘š in Section 7.1 with 𝑐 = 1. Let 𝑠 ≥ 1 be such that π‘ž = 𝑝𝑠 .
∑οΈ€
Let d = (𝑑1 , . . . , π‘‘π‘š ) ∈ Deg(π’ž). We will show that π‘š
𝑖=1 𝑑𝑖 ≤ 𝑑 = π‘ž − π‘ž/𝑝 − 1. Suppose,
∑οΈ€π‘š
for the sake of contradiction, that 𝑖=1 𝑑𝑖 ≥ π‘ž − π‘ž/𝑝. We will exhibit e ≤𝑝 d such that
∑οΈ€
π‘ž − π‘ž/𝑝 ≤ π‘š
𝑖=1 𝑒𝑖 < π‘ž, which contradicts the fact that d ∈ Deg(π’ž) by Proposition 5.2.3. If
∑οΈ€π‘š (𝑠)
∑οΈ€π‘š
∑οΈ€π‘š
𝑖=1 𝑑𝑖 .
𝑖=1 𝑑𝑖 ≥ π‘ž. Let π‘Ž ,
𝑖=1 𝑑𝑖 < π‘ž, then we are done by taking e = d, so assume
(𝑠)
By Claim 7.1.9, π‘Ž ≤ 𝑝 − 2. For each 𝑖 ∈ [π‘š], let 𝑑𝑖 = 𝑑𝑖 − 𝑑𝑖 𝑝𝑠−1 . Then
π‘š
∑︁
𝑑𝑖 =
𝑖=1
π‘š
∑︁
𝑑𝑖 − 𝑝
𝑠−1
𝑖=1
π‘š
∑︁
(𝑠)
𝑑𝑖 ≥ (𝑝 − π‘Ž)𝑝𝑠−1 .
𝑖=1
∑οΈ€
Let π‘˜ ∈ [π‘š] be the minimal integer such that π‘˜π‘–=1 𝑑𝑖 ≥ (𝑝−π‘Ž−1)𝑝𝑠−1 . Since each 𝑑𝑖 < 𝑝𝑠−1 ,
∑οΈ€
this implies that π‘˜π‘–=1 𝑑𝑖 < (𝑝 − π‘Ž)𝑝𝑠−1 . Now, for 𝑖 ∈ [π‘˜], define 𝑒𝑖 , 𝑑𝑖 , and for 𝑖 > π‘˜, define
(𝑠)
𝑒𝑖 , 𝑑𝑖 𝑝𝑠−1 . By construction, e ≤𝑝 d. On the other hand,
π‘š
∑︁
𝑒𝑖 =
𝑖=1
π‘˜
∑︁
𝑑𝑖 + 𝑝
𝑠−1
π‘š
∑︁
(𝑠)
𝑑𝑖
𝑖=1
𝑖=1
=
π‘˜
∑︁
𝑑𝑖 + π‘Žπ‘π‘ −1
𝑖=1
which is at least (𝑝 − π‘Ž − 1)𝑝𝑠−1 + π‘Žπ‘π‘ −1 = (𝑝 − 1)𝑝𝑠−1 = π‘ž − π‘ž/𝑝 and strictly less than
(𝑝 − π‘Ž)𝑝𝑠−1 + π‘Žπ‘π‘ −1 = 𝑝𝑠 = π‘ž.
π‘ž−π‘ž/𝑝
Now, let 𝑑 = π‘ž − π‘ž/𝑝. Let 𝑓 (X) = 𝑋1
π‘ž−π‘ž/𝑝
𝑋2
93
. Clearly deg(𝑓 ) = 2(π‘ž − π‘ž/𝑝) > π‘ž − π‘ž/𝑝.
We claim that 𝑓 ∈ RS(π‘ž, 𝑑)1β†—π‘š . For any a, b ∈ Fπ‘š
π‘ž , we have
𝑓 (a𝑇 + b) = (π‘Ž1 𝑇 + 𝑏1 )π‘ž−π‘ž/𝑝 (π‘Ž2 𝑇 + 𝑏2 )π‘ž−π‘ž/𝑝
π‘ž/𝑝
π‘ž/𝑝
π‘ž/𝑝
(7.1)
π‘ž/𝑝
= (π‘Ž1 𝑇 π‘ž/𝑝 + 𝑏1 )𝑝−1 (π‘Ž2 𝑇 π‘ž/𝑝 + 𝑏2 )𝑝−1
(7.2)
(7.3)
so, since 𝑇 π‘ž = 𝑇 , every monomial in 𝑓 (a𝑇 + b) is of the form 𝑇 𝑐·π‘ž/𝑝 for some 0 ≤ 𝑐 ≤ 𝑝 − 1,
thus deg(𝑓 (a𝑇 + b)) ≤ π‘ž − π‘ž/𝑝 = 𝑑.
As a corollary, we see that, over prime fields, lifting the Reed-Solomon code simply yields
the Reed-Muller code.
Corollary 7.1.7. If 𝑝 is a prime and 𝑑 < π‘ž, then RS(𝑝, 𝑑)1β†—π‘š = RM(𝑝, 𝑑, π‘š) for every
π‘š ≥ 1.
Proof. If π‘š = 1, then both codes are equal to RS(𝑝, 𝑑), so there is nothing to prove. Assume
π‘š ≥ 2. If 𝑑 = 𝑝 − 1, then both codes are equal to everything, so again there is nothing to
prove. If 𝑑 < 𝑝 − 1, then the result follows from Theorem 7.1.6.
Theorem 7.1.8. Let π‘ž be a prime power, let 𝑑 < π‘ž, and let π‘š ≥ 2. Then RM(π‘ž, 𝑑, 2)2β†—π‘š =
RM(π‘ž, 𝑑, π‘š).
Proof. By Lemma 7.1.5, it suffices to consider the case where 𝑑 = π‘ž−1. It follows immediately
from definitions that RM(π‘ž, 𝑑, π‘š) ⊆ RM(π‘ž, 𝑑, 2)2β†—π‘š , so it only remains to show the reverse
inclusion. We do so by showing that if d ∈ Deg(RM(π‘ž, 𝑑, 2)2β†—π‘š ), then β€–dβ€– ≤ 𝑑. Actually,
we show the converse. By Proposition 5.2.3, it suffices to show that if β€–dβ€– ≥ π‘ž, then there
exists E ≤𝑝 d such that (β€–E*1 β€– mod* π‘ž) + (β€–E*2 β€– mod* π‘ž) ≥ π‘ž where E has rows [π‘š] and
columns {0, 1, 2}.
Without loss of generality, assume 𝑑1 ≥ 𝑑2 ≥ · · · ≥ π‘‘π‘š . Let π‘˜ ≥ 2 be the smallest integer
such that 𝑑2 + · · · + π‘‘π‘˜ ≥ π‘ž − 𝑑1 . Then 𝑑2 + · · · + π‘‘π‘˜ ≤ π‘ž − 1, for otherwise π‘‘π‘˜ ≥ 𝑑1 + 1,
which is impossible. Construct E as follows. Define E1* , (0, 𝑑1 , 0) and for 𝑖 ∈ {2, . . . , π‘˜},
define E𝑖* , (0, 0, 𝑑𝑖 ). Finally, for 𝑖 > π‘˜, define E𝑖* = (0, 0, 0). By construction, E ≤𝑝 d,
94
and moreover β€–E*1 β€– = 𝑑1 and β€–E*2 β€– = 𝑑2 + · · · + π‘‘π‘˜ . Since β€–E*1 β€–, β€–E*2 β€– ≤ π‘ž − 1, applying
mod* π‘ž does not reduce them, hence (β€–E*1 β€– mod* π‘ž) + (β€–E*2 β€– mod* π‘ž) = β€–E*1 β€– + β€–E*2 β€– ≥
𝑑1 + (π‘ž − 𝑑1 ) = π‘ž, as desired.
7.1.2
Rate
Rate is the one key parameter that is not guaranteed by the lifting operator. However, the
potential in lifted codes comes from the fact that it is possible to get extremely dense codes
by lifting. From Section 7.1.1, we see that if π‘ž − π‘ž/𝑝 ≤ 𝑑 < π‘ž, then there are lifted ReedSolomon codewords that are not Reed-Muller codewords. We show that, in fact, there are
many such codewords. Our strategy is to lower bound the rate of the code. Observe that
the lifted Reed-Solomon code, being a linear affine-invariant code, is spanned by monomials,
i.e. has a degree set. Therefore, its dimension is equal to the size of its degree set. We use
our knowledge of the structure of its degree set to lower bound the number of such degrees.
In this section, we analyze the rate of π’ž for special values of 𝑑. Let 𝑝 be the characteristic
of Fπ‘ž , let 𝑠 ≥ 1 be such that π‘ž = 𝑝𝑠 , let 𝑐 ≤ 𝑠 and let 𝑑 = (1 − 𝑝−𝑐 )π‘ž − 1.
We begin by re-interpreting what it means for a number 𝑒 to be less than 𝑑, in terms of
∑οΈ€
the 𝑝-ary expansion of 𝑒. Through this section, for any π‘Ž ∈ N, let π‘Ž = 𝑖≥0 π‘Ž(𝑖) 𝑝𝑖 be the
𝑝-ary expansion of π‘Ž.
Claim 7.1.9. If 𝑒 ∈ Jπ‘žK, then 𝑒 ≤ 𝑑 if and only if 𝑒(𝑖) < 𝑝 − 1 for some 𝑠 − 𝑐 ≤ 𝑖 < 𝑠.
Proof. Since 𝑒 < π‘ž, 𝑒(𝑖) = 0 for 𝑖 ≥ 𝑠. Note that if 𝑒(𝑖) = 𝑝 − 1 for 𝑠 − 𝑐 ≤ 𝑖 < 𝑠, then
∑οΈ€
∑︀𝑐−1 𝑖
𝑖
𝑠−𝑐
𝑒 ≥ (𝑝 − 1) 𝑠−1
(𝑝 − 1) 𝑖=0
𝑝 = (1 − 𝑝−𝑐 )π‘ž, while if 𝑒(𝑖) < 𝑝 − 1 for some
𝑖=𝑠−𝑐 𝑝 = 𝑝
𝑠 − 𝑐 ≤ 𝑖 < 𝑠, this only decreases the value of 𝑒.
Next, we use our knowledge of the degree set of the lifted Reed-Solomon code to provide
a sufficient condition for a degree to be in the degree set.
⌈οΈ€
⌉οΈ€
Claim 7.1.10. Let 𝑏 = log𝑝 π‘š + 1 and let d ∈ Jπ‘žKπ‘š . If there is some 𝑠 − 𝑐 ≤ 𝑗 ≤ 𝑠 − 𝑏
(π‘˜)
such that 𝑑𝑖
= 0 for every 𝑖 ∈ [π‘š] and every 𝑗 ≤ π‘˜ < 𝑗 + 𝑏, then d ∈ Deg(π’ž).
95
Proof. Let e ≤𝑝 d, and let 𝑒 =
∑οΈ€π‘š
𝑖=1 𝑒𝑖
mod* π‘ž. We claim that 𝑒(𝑗+𝑏−1) = 0, which implies
𝑒 ≤ 𝑑 Claim 7.1.9 since 𝑠 − 𝑐 ≤ 𝑗 + 𝑏 − 1 < 𝑠. Note that π‘Ž ↦→ 𝑝 · π‘Ž mod* π‘ž results in a cyclic
permutation of the digits π‘Ž(𝑖) . So we may multiply d and e by an appropriate power of 𝑝,
namely 𝑝𝑠−𝑏−𝑗 , so that 𝑗 = 𝑠 − 𝑏. Therefore, we may assume without loss of generality that
𝑗 = 𝑠 − 𝑏, and we wish to show that 𝑒(𝑠−1) = 0, i.e. 𝑒 < 𝑝𝑠−1 . Note that since e ≤𝑝 d, for each
(π‘˜)
𝑖 ∈ [π‘š] and 𝑠 − 𝑏 ≤ π‘˜ ≤ 𝑠 − 1 we have 𝑒𝑖
∑οΈ€π‘š
𝑠−𝑏
< 𝑝𝑏−1 𝑝𝑠−𝑏 = 𝑝𝑠−1 .
𝑖=1 𝑒𝑖 < π‘šπ‘
= 0, i.e. 𝑒𝑖 < 𝑝𝑠−𝑏 for every 𝑖 ∈ [π‘š]. Therefore
Finally, we lower bound the rate.
⌈οΈ€
⌉οΈ€
π‘šπ‘
Theorem 7.1.11. Let 𝑏 = log𝑝 π‘š + 1 The rate of the code π’ž is at least 1 − 𝑒−(𝑐+𝑏)/(𝑏𝑝 ) .
Proof. Consider choosing d ∈ Jπ‘žKπ‘š uniformly at random. Let π‘Ž = ⌊𝑐/𝑏⌋. For 𝑗 ∈ [π‘Ž], let 𝐸𝑗
(π‘˜)
be the event that 𝑑𝑖 = 0 for every 𝑖 ∈ [π‘š] and 𝑠 − 𝑗𝑏 ≤ π‘˜ < 𝑠 − (𝑗 − 1)𝑏. By Claim 7.1.10,
]︁
[︁⋁︀
β‹οΈ€π‘Ž
π‘Ž
𝑗=1 𝐸𝑗 . We have
𝑗=1 𝐸𝑗 is sufficient for d ∈ Deg(π’ž), so we wish to lower bound Pr
Pr[𝐸𝑗 ] = 𝑝−π‘šπ‘ for every 𝑗 ∈ [π‘Ž], therefore
[οΈƒ
Pr
π‘Ž
⋁︁
]οΈƒ
𝐸𝑗
[οΈƒ
= 1 − Pr
𝑗=1
π‘Ž
⋀︁
]οΈƒ
𝐸𝑗
(7.4)
𝑗=1
= 1−
= 1−
π‘Ž
∏︁
𝑗=1
π‘Ž
∏︁
[οΈ€ ]οΈ€
Pr 𝐸𝑗
(7.5)
(1 − Pr[𝐸𝑗 ])
(7.6)
𝑗=1
= 1 − (1 − 𝑝−π‘šπ‘ )π‘Ž
(7.7)
≥ 1 − (1 − 𝑝−π‘šπ‘ )(𝑐+𝑏)/𝑏
(7.8)
≥ 1 − 𝑒−(𝑐+𝑏)/(𝑏𝑝
96
π‘šπ‘ )
.
(7.9)
7.1.3
Global List-Decoding
In this section, we present an efficient global list decoding algorithm for RS(π‘ž, 𝑑)1β†—π‘š . Define
𝛼1 , . . . , π›Όπ‘š ∈ Fπ‘žπ‘š , πœ‘, and πœ‘* as in Section 5.3.3. Our main result states that RS(π‘ž, 𝑑)1β†—π‘š
is isomorphic to a subcode of RS(π‘ž π‘š , (𝑑 + π‘š)π‘ž π‘š−1 ) ⊆ {Fπ‘žπ‘š → Fπ‘ž }. In particular, one can
simply list decode RS(π‘ž, 𝑑)1β†—π‘š by list-decoding RS(π‘ž π‘š , (𝑑 + π‘š)π‘ž π‘š−1 ) up to the Johnson
radius. We will use this algorithm for π‘š = 2 as a subroutine in our local list decoding
algorithm in Section 7.1.4.
Theorem 7.1.12. If 𝑓 ∈ RS(π‘ž, 𝑑)1β†—π‘š , then deg(πœ‘* (𝑓 )) ≤ (𝑑 + π‘š)π‘ž π‘š−1 .
Proof. By linearity, it suffices to prove this for a monomial 𝑓 (𝑋1 , . . . , π‘‹π‘š ) =
∏οΈ€π‘š
𝑖=1
𝑋𝑖𝑑𝑖 . We
have
∑︁
πœ‘* (𝑓 )(𝑍) =
(· · · )𝑍
∑οΈ€
𝑖𝑗
𝑒𝑖𝑗 π‘ž π‘š−𝑖
,
(𝑒11 ,...,𝑒1π‘š )≤𝑝 𝑑1
..
.
(π‘’π‘š1 ,...,π‘’π‘šπ‘š )≤𝑝 π‘‘π‘š
so it suffices to show that
∑οΈ€π‘š ∑οΈ€π‘š
𝑗=1
𝑖=1 𝑒𝑖𝑗 π‘ž
π‘š−𝑗
mod* π‘ž π‘š ≤ (𝑑 + π‘š)π‘ž π‘š−1 . By Proposition 7.1.2,
∑οΈ€
*
for every 𝑒𝑖 ≤𝑝 𝑑𝑖 , 𝑖 ∈ [π‘š], we have π‘š
𝑖=1 𝑒𝑖 mod π‘ž ≤ 𝑑. Therefore, there is some integer
∑οΈ€
0 ≤ π‘˜ < π‘š such that π‘š
𝑖=1 𝑒𝑖1 ∈ [π‘˜π‘ž, π‘˜(π‘ž − 1) + 𝑑]. Thus,
π‘˜π‘ž
π‘š
≤ π‘ž
π‘š−1
≤ π‘ž π‘š−1
π‘š
∑︁
𝑖=1
π‘š
∑︁
𝑒𝑖1 +
π‘š ∑︁
π‘š
∑︁
𝑗=2 𝑖=1
π‘š ∑︁
π‘š
∑︁
π‘š−2
𝑒𝑖1 + π‘ž
𝑖=1
≤ (π‘˜(π‘ž − 1) + 𝑑)π‘ž
and hence
∑οΈ€π‘š ∑οΈ€π‘š
𝑗=1
𝑖=1 𝑒𝑖𝑗 π‘ž
π‘š−𝑗
𝑒𝑖𝑗 π‘ž π‘š−𝑗
𝑒𝑖𝑗
𝑗=2 𝑖=1
π‘š−1
π‘š−1
+ π‘šπ‘ž
(7.10)
(7.11)
(7.12)
= π‘˜(π‘ž π‘š − 1) + (𝑑 + π‘š − π‘˜)π‘ž π‘š−1 + π‘˜
(7.13)
≤ π‘˜(π‘ž π‘š − 1) + (𝑑 + π‘š)π‘ž π‘š−1
(7.14)
mod* π‘ž π‘š ≤ (𝑑 + π‘š)π‘ž π‘š−1 .
Corollary 7.1.13. For every π‘š ≥ 2 and ever πœ– > 0, there is a polynomial time algorithm
2
that takes as input a function π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž and outputs a list β„’ of size |β„’| = 𝑂(1/πœ– ) which
97
contains all 𝑐 ∈ RS(π‘ž, 𝑑)1β†—π‘š such that 𝛿(π‘Ÿ, 𝑐) < 1 −
√︁
𝑑+π‘š
π‘ž
− πœ–.
′
*
Proof. Given π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž , convert it to π‘Ÿ = πœ‘ (π‘Ÿ), and then run the Guruswami-Sudan
list decoder (Theorem 3.4.3) for RS , RS(π‘ž π‘š , (𝑑 + π‘š)π‘ž π‘š−1 ) on π‘Ÿ′ to obtain a list β„’ of size
√︁
2
′
− πœ– lies in β„’.
|β„’| = 𝑂(1/πœ– ) with the guarantee that any 𝑐 ∈ RS with 𝛿(π‘Ÿ , 𝑐) < 1 − 𝑑+π‘š
π‘ž
√︁
We require that any 𝑐 ∈ RS(π‘ž, 𝑑)1β†—π‘š satisfying 𝛿(π‘Ÿ′ , 𝑐) < 1 − 𝑑+π‘š
− πœ– lies in β„’, and this
π‘ž
follows immediately from Theorem 7.1.12.
7.1.4
Local List-Decoding
In this section, we present a local list decoding algorithm for LiftedRS(π‘ž, 𝑑, π‘š), where 𝑑 =
√
(1 − 𝛿)π‘ž which decodes up to radius 1 − 1 − 𝛿 − πœ– for any constant πœ– > 0, with list size
𝑂(1/πœ–2 ) and query complexity π‘ž 3 .
Theorem 7.1.14. For every π‘š ≥ 2 and every 𝛿, πœ– > 0, setting 𝑑 = (1 − 𝛿)π‘ž, RS(π‘ž, 𝑑)1β†—π‘š is
(︁
(︁
)︁)︁
√
π‘ž, π‘ž 3 , 1 − 1 − 𝛿 − πœ–, 𝑂(1/πœ–2 ), 0.2 + π›Ώπ‘ž2 , 𝑂 πœ–61π‘žπ›Ώ -locally list-decodable.
Proof. The following algorithm is the outer local list-decoding algorithm.
Local list decoder: Oracle access to received word π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž .
1. Pick an affine transformation β„“ : Fπ‘ž → Fπ‘š
π‘ž uniformly at random.
2. Run Reed-Solomon list decoder (e.g. Guruswami-Sudan) on π‘Ÿ ∘ β„“ from 1 −
√
1−𝛿−
πœ–
2
fraction errors to get list 𝑔1 , . . . , 𝑔𝐿 : Fπ‘ž → Fπ‘ž of Reed-Solomon codewords.
3. For each 𝑖 ∈ [𝐿], output Correct(𝐴ℓ,𝑔𝑖 )
where Correct is a local correction algorithm for the lifted codes for 0.1𝛿 fraction errors
given by Theorem 5.3.1, which has a failure probability of 0.2 + π›Ώπ‘ž2 , and 𝐴 is an oracle which
takes as advice a univariate affine map and a univariate polynomial and simulates oracle
access to a function which is supposed to be β‰ͺ 0.1𝛿 close to a lifted RS codeword.
98
Oracle 𝐴ℓ,𝑔 (x):
1. Let 𝑃 : F2π‘ž → Fπ‘š
π‘ž be the unique affine map such that 𝑃 (𝑑, 0) = β„“(𝑑) for 𝑑 ∈ Fπ‘ž and
𝑃 (0, 1) = x.
2. Use the global list decoder for RS(π‘ž, 𝑑)1β†—2 given by Corollary 7.1.13 to decode π‘Ÿ ∘ 𝑃
√
from 1 − 1 − 𝛿 − 2πœ– fraction errors and obtain a list β„’.
3. If there exists a unique β„Ž ∈ β„’ such that β„Ž ∘ β„“ = 𝑔, output β„Ž(0, 1), otherwise fail.
Analysis: For the query complexity, note that the local list-decoder makes π‘ž queries to π‘Ÿ
to output its oracles. For each oracle, Correct makes π‘ž queries to 𝐴ℓ,𝑔𝑖 , which itself makes π‘ž 2
queries to π‘Ÿ, so the entire oracle makes π‘ž 3 queries to π‘Ÿ.
To show correctness, we just have to show that, with high probability over the choice of
√
β„“, for every lifted RS codeword 𝑓 such that 𝛿(π‘Ÿ, 𝑓 ) < 1 − 1 − 𝛿 − πœ–, there is 𝑖 ∈ [𝐿] such
that Correct(𝐴ℓ,𝑔𝑖 ) = 𝑓 , i.e. 𝛿(𝐴ℓ,𝑔𝑖 , 𝑓 ) ≤ 0.1𝛿.
Fix such a function 𝑓 . We will proceed in two steps:
1. First, we show that with high probability over β„“, there is some 𝑖 ∈ [𝐿] such that 𝑓 |β„“ = 𝑔𝑖 .
2. Next, we show that 𝛿(𝐴ℓ,𝑓 βˆ˜β„“ , 𝑓 ) ≤ 0.1𝛿 with high probability.
√
For the first step, note that 𝑓 |β„“ ∈ {𝑔1 , . . . , 𝑔𝐿 } if 𝛿(𝑓 ∘ β„“, π‘Ÿ ∘ β„“) ≤ 1 − 1 − 𝛿 − 2πœ– . By
√
Proposition B.1.4, 𝛿(𝑓 ∘ β„“, π‘Ÿ ∘ β„“) has mean less than 1 − 1 − 𝛿 − πœ– and variance less than
√
1/(4π‘ž). By Chebyshev’s inequality, the probability that 𝛿(𝑓 |β„“ , π‘Ÿ|β„“ ) ≤ 1 − 1 − 𝛿 − 2πœ– is at
least 1 −
1
.
πœ–2 π‘ž
For the second step, we want to show that Prx∈Fπ‘š
[𝐴ℓ,𝑓 βˆ˜β„“ (x) ΜΈ= 𝑓 (x)] ≤ 0.1𝛿. First consider
π‘ž
the probability when we randomize β„“ as well. Then 𝑃 : F2π‘ž → Fπ‘š
π‘ž is a uniformly random
affine map. We get 𝐴ℓ,𝑓 βˆ˜β„“ (x) = 𝑓 (x) as long as 𝑓 ∘ 𝑃 ∈ β„’ and no other element β„Ž ∈ β„’ has
√
β„Ž ∘ β„“ = 𝑓 ∘ β„“. By Proposition B.1.4, 𝛿(𝑓 ∘ 𝑃, π‘Ÿ ∘ 𝑃 ) has mean less than 1 − 1 − 𝛿 − πœ– and
variance less than 1/(4π‘ž 2 ). By Chebychev’s inequality, with probability at least 1 − πœ–21π‘ž2 over
√
β„“ and x , we have 𝛿(𝑓 |𝑃 , π‘Ÿ|𝑃 ) ≤ 1 − 1 − 𝛿 − 2πœ– and hence 𝑓 |𝑃 ∈ β„’. For the probability
99
that no two codewords in β„’ agree on β„“, view this as first choosing a random affine map
′
′
′
2
𝑃 : F2π‘ž → Fπ‘š
π‘ž and then choosing random affine β„“ : Fπ‘ž → Fπ‘ž and defining β„“ , 𝑃 ∘ β„“ . For
any distinct πœ‘, πœ“ : F2π‘ž → Fπ‘ž , the probability over β„“′ that πœ‘ ∘ β„“′ = πœ“ ∘ β„“′ is at most 2/π‘ž, by
Proposition B.1.5. Since |β„’| = 𝑂(1/πœ–2 ), there are at most 𝑂(1/πœ–4 ) pairs of functions from β„’,
so by the union bound, with probability at least 1 − 𝑂(1/(πœ–4 π‘ž)), all the functions in β„’ are
distinct on β„“ and hence 𝑓 ∘ 𝑃 is the unique codeword in β„’ which is consistent with 𝑓 ∘ β„“.
(︁ )︁
Therefore, the probability, over β„“ and x, that 𝐴ℓ.𝑓 βˆ˜β„“ (x) ΜΈ= 𝑓 (x) is 𝑂 πœ–41π‘ž , and thus
[︁
]︁
Pr [𝛿(𝐴ℓ,𝑓 βˆ˜β„“ , 𝑓 ∘ β„“) > 0.1𝛿] = Pr Pr[𝐴ℓ,𝑓 βˆ˜β„“ (x) ΜΈ= 𝑓 (x)] > 0.1𝛿
β„“
β„“
x
Prβ„“,x [𝐴ℓ,𝑓 βˆ˜β„“ (x) ΜΈ= 𝑓 (x)]
≤
(οΈ‚
)οΈ‚0.1𝛿
1
= 𝑂 4
.
πœ– π‘žπ›Ώ
√
So, for a fixed codeword 𝑓 ∈ RS(π‘ž, 𝑑)1β†—π‘š such that 𝛿(π‘Ÿ, 𝑓 ) < 1− 1 − 𝛿−πœ–, the probability
(︁
)︁
1
of success is 𝑂 πœ–4 π‘žπ›Ώ . Since there are 𝑂(1/πœ–2 ) such codewords, the overall probability of
(︁
)︁
success is 𝑂 πœ–61π‘žπ›Ώ .
As a corollary, we get the following testing result.
Theorem 7.1.15. For every 𝛿 > 0 and for any 𝛼 < 𝛽 < 1 −
√
1 − 𝛿, there is an algo-
rithm which, given oracle access to a function π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž , distinguishes between the cases
where π‘Ÿ is 𝛼-close to RS(π‘ž, 𝑑)1β†—π‘š (where 𝑑 , (1 − 𝛿)π‘ž), and where π‘Ÿ is 𝛽-far, while making
(︁
)︁
4
𝑂 ln(1/(𝛽−𝛼))π‘ž
queries to π‘Ÿ.
4
72
(𝛽−𝛼) 𝛿
Proof. Let 𝜌 , (𝛼 + 𝛽)/2, and let πœ– , (𝛽 − 𝛼)/8, so that 𝛼 = 𝜌 − 4πœ– and 𝛽 = 𝜌 + 4πœ–.
Set πœ‚ , 𝑂(πœ–2 ), where the constant is sufficiently large. Let 𝑇 ′ be the π‘ž 2 -query Ω(𝛿 72 )-robust
tester algorithm for 𝑅𝑆(π‘ž, 𝑑)1β†—π‘š , which rejects words that are πœ–-far with probability Ω(πœ–π›Ώ 72 ),
by Proposition 3.3.2. Let 𝑇 be the 𝑂(ln(1/πœ‚)π‘ž 2 /(πœ–π›Ώ 72 ))-query tester which runs 𝑇 ′ repeatedly
𝑂(ln(1/πœ‚)π‘ž 2 /(πœ–π›Ώ 72 )) times and accepts if and only if every iteration accepts, to increase the
rejection probability for πœ–-far words to 1 − πœ‚. Our algorithm is to run the local list-decoding
100
algorithm on π‘Ÿ with error radius 𝜌, to obtain a list of oracles 𝑀1 , . . . , 𝑀𝐿 . For each 𝑀𝑖 ,
ΜƒοΈ€ 𝑓𝑖 ) of the distance between π‘Ÿ and the
we use random sampling to compute an estimate 𝛿(π‘Ÿ,
function 𝑓𝑖 computed by 𝑀𝑖 to within πœ– additive error with failure probability πœ‚, and keep
only the ones with estimated distance less than 𝜌 + πœ–. Then, for each remaining 𝑀𝑖 , we run
𝑇 on 𝑀𝑖 . We accept if 𝑇 accepts some 𝑀𝑖 , otherwise we reject.
The number of queries required to run the local list-decoding algorithm is π‘ž. For each
ΜƒοΈ€ 𝑓𝑖 ) is 𝑂(ln(1/πœ‚)/πœ–2 ),
𝑖 ∈ [𝐿], the number of queries to 𝑀𝑖 we need to make for computing 𝛿(π‘Ÿ,
by Proposition 2.2.4. Each query to 𝑀𝑖 makes π‘ž 2 queries to π‘Ÿ, for a total of 𝑂(ln(1/πœ‚)π‘ž 2 /πœ–2 )
queries for each 𝑀𝑖 and therefore 𝑂(ln 1/πœ‚)π‘ž 2 /πœ–4 ) queries for the distance estimation step.
The tester 𝑇 makes 𝑂(ln(1/πœ‚)π‘ž 2 /(πœ–π›Ώ 72 )) queries to each 𝑀𝑖 , for a total of 𝑂(ln(1/πœ‚)π‘ž 4 /(πœ–π›Ώ 72 ))
queries to π‘Ÿ for each 𝑖 ∈ [𝐿] and therefore a grand total of 𝑂(ln(1/πœ‚)π‘ž 4 /(πœ–3 𝛿 72 )) queries to
π‘Ÿ made by 𝑇 . Therefore, the total number of queries to π‘Ÿ made by our testing algorithm is
(︁
)︁
(︁
)︁
4
ln(1/πœ–)π‘ž 4
𝑂 ln(1/πœ‚)π‘ž
=
𝑂
.
πœ–4 𝛿 72
πœ–4 𝛿 72
If π‘Ÿ is 𝛼-close to RS(π‘ž, 𝑑)1β†—π‘š , then it is 𝛼-close to some codeword 𝑓 , and by the guarantee
of the local list-decoding algorithm, there is some 𝑗 ∈ [𝐿] such that 𝑀𝑗 computes 𝑓 . Since,
ΜƒοΈ€ 𝑓 ) ≤ 𝛼 + πœ– < 𝜌, in which case 𝑀𝑗
𝛿(π‘Ÿ, 𝑓 ) ≤ 𝛼, with probability at least 1 − πœ‚ we have 𝛿(π‘Ÿ,
will not be pruned by our distance estimation. Since 𝑓 is a codeword, this 𝑀𝑗 will pass
the testing algorithm 𝑇 and so our algorithm will accept. The total failure probability is
(︁
)︁
1
therefore πœ‚ + 𝑂 πœ–6 π‘žπ›Ώ .
)︁
(︁
1β†—π‘š
1
Now suppose π‘Ÿ is 𝛽-far from RS(π‘ž, 𝑑)
. With probability 1 − 𝑂 πœ–6 π‘žπ›Ώ , the local listdecoding algorithm succeeds, so let us condition on its success. With probability 1−𝑂(πœ‚/πœ–2 ),
all the distance estimations for all the 𝑀𝑖 simultaneously succeed, and let us condition on this
as well. Consider any oracle 𝑀𝑖 output by the local list-decoding algorithm and not pruned
by our distance estimation. Let 𝑓𝑖 be the function computed by 𝑀𝑖 . Then the estimated
ΜƒοΈ€ 𝑓𝑖 ) < 𝜌 + πœ–, so the true distance is 𝛿(π‘Ÿ, 𝑓𝑖 ) < 𝜌 + 2πœ–. Since π‘Ÿ is 𝛽-far from any
distance is 𝛿(π‘Ÿ,
codeword, that means the distance from 𝑓𝑖 to any codeword is at least 𝛽 − (𝜌 + 2πœ–) > πœ–, and
hence 𝑇 will reject 𝑀𝑖 with probability 1 − πœ‚. By the union bound, all the 𝑀𝑖 not pruned
will be rejected with probability 1 − 𝑂(πœ‚/πœ–2 ). So the total failure probability in this case is
101
at most 𝑂(πœ‚/πœ–2 ) + 𝑂
(︁
1
πœ–6 π‘žπ›Ώ
)︁
.
In either case, the failure probability is at most some constant.
7.1.5
Main Result: The Code That Does It All
The following is our main theorem, Theorem 7.1.16, a more precise restatement of Theorem 1.2.1, which is the culmination of the lifting technology results of Chapters 5 and 6. It
states that there is a high-rate code that is simultaneously locally correctable and decodable,
locally list-decodable, and robustly testable.
Theorem 7.1.16. For every 𝛼, 𝛽 > 0 there exists 𝛿 > 0 such that the following holds: For
infinitely many 𝑛 ∈ N, there is π‘ž = π‘ž(𝑛) = 𝑂(𝑛𝛽 ) and a linear code π’ž ⊆ Fπ‘›π‘ž with distance
𝛿(π’ž) ≥ 𝛿, rate at least 1 − 𝛼, that is locally correctable and decodable up to
𝛿
2
fraction errors
with 𝑂(π‘ž) queries, locally list-decodable up to the Johnson bound with 𝑂(π‘ž 3 ) queries, and
Ω(𝛿 72 )-robustly testable with π‘ž 2 queries.
Proof. Let π‘š = ⌈1/𝛽⌉ and let 𝑏 = ⌈log2 π‘š⌉ + 1. Let 𝑐 = 𝑏2π‘šπ‘ · ⌈ln(1/𝛼)⌉ + 𝑏, and set 𝛿 , 2−𝑐 .
For 𝑠 ≥ 𝑐, set π‘ž = 2𝑠 and 𝑛 , π‘ž π‘š . Set 𝑑 , (1 − 𝛿)π‘ž − 1. Let π’ž , RS(π‘ž, 𝑑)1β†—π‘š .
By Proposition 7.1.1, 𝛿(π’ž) ≥ 𝛿. By Theorem 7.1.11, the rate is at least 1 − 𝛼. By
Proposition 7.1.3, π’ž is locally correctable and decodable up to
𝛿
2
fraction errors with π‘ž
queries. By Theorem 7.1.14, π’ž is locally list-decodable up to the Johnson bound with π‘ž 3
queries. By Proposition 7.1.4, π’ž is Ω(𝛿 72 )-robustly testable with π‘ž 2 queries.
7.2
Robust Low-Degree Testing
In this section, we prove Theorem 7.2.2, which is simply a more precise restatement of Theorem 1.2.3. We do so by proving Theorem 7.2.1, a generalization from which Theorem 7.2.2
follows immediately. Theorem 7.2.1 replaces the Reed-Solomon code with an arbitrary univariate linear affine-invariant code π’ž0 and replaces the bivariate Reed-Muller code with an
arbitrary 𝑑-variate linear affine-invariant code π’ž1 which is a strict subcode of (π’ž0 )1↗𝑑 .
102
Theorem 7.2.1. Let 𝑑 > 1 and let π‘š ≥ 3. Let π’ž0 ⊆ {Fπ‘ž → Fπ‘ž } and π’ž1 ⊆ {Fπ‘‘π‘ž → Fπ‘ž } be
linear affine-invariant codes such that π’ž1 ( (π’ž0 )1↗𝑑 . Let 𝛿 , 𝛿(π’ž0 ). Fix π‘Ÿ : Fπ‘š
π‘ž → Fπ‘ž . Let 𝜌 ,
. Let 𝛼2 be
E𝑒 [𝛿 (π‘Ÿ|𝑒 , π’ž1 )], where the expectation is taken over random 𝑑-dimensional 𝑒 ⊆ Fπ‘š
{︁ 2 (οΈ€
}︁ π‘ž
)οΈ€
the 𝑑-dimensional robustness of (π’ž0 )1β†—π‘š . Then 𝜌 ≥ min 𝛼24𝛿 , 2𝛿 − 2π‘ž −1 · 2𝛿 ·π›Ώ(π‘Ÿ, (π’ž1 )π‘‘β†—π‘š ).
Theorem 7.2.2 (Robust plane testing for Reed-Muller). Let π‘š ≥ 3. Fix a positive
constant 𝛿 > 0 and a degree 𝑑 = (1 − 𝛿) · π‘ž. Let RM(π‘š) be the π‘š-variate Reed-Muller codes
(︁ 74
)︁
𝛿
of degree 𝑑 over Fπ‘ž . Then RM(π‘š) is 8·10
,
2
-robust.
52
Proof. Let RS be the Reed-Solomon code over Fπ‘ž of degree 𝑑. Let 𝑝 be the characteristic of
π‘ž. Let 𝛼1 be the 2-dimensional robustness of RS1β†—π‘š . Then 𝛼2 ≥
𝛿 72
2·1052
by Theorem 6.1.4 if
π‘š = 3, and by Theorem 6.1.18 if π‘š ≥ 4.
If 𝑑 < π‘ž − π‘ž/𝑝, then RM(π‘š) = RS1β†—π‘š by Theorem 7.1.6, and so in this case the theorem
follows immediately from Theorem 6.1.18. If 𝑑 ≥ π‘ž − π‘ž/𝑝 and π‘ž ≥ 8𝛿 , then RM(2) ( RS1β†—2
(by Theorem 7.1.6) but RM(π‘š) = RM(2)2β†—π‘š (by Theorem 7.1.8), and so in this case the
theorem follows immediately from Theorem 7.2.1, with π’ž0 = RS, 𝑑 = 2, and π’ž1 = RM(2). If
π‘ž<
8
𝛿
then the theorem follows from Corollary 6.1.2.
Overview of Proof of Theorem 7.2.1. We illustrate the idea for the case where 𝑑 = 2,
π’ž0 is the Reed-Solomon code, and π’ž1 is the bivariate Reed-Muller code of the same degree.
The generalization to arbitrary 𝑑 and codes π’ž0 , π’ž1 is straightforward. If π‘Ÿ is far from the
lifted code, then on random planes π‘Ÿ will be far from the bivariate lifted code and hence also
from the bivariate Reed-Muller code. So the remaining case is when π‘Ÿ is close to the lifted
code. If the nearest function is a Reed-Muller codeword, then the theorem follows from the
robustness of the lifted code. Otherwise, if the nearest function 𝑐 is not Reed-Muller, then
we show (through Corollary 6.2.8) that on many planes 𝑐 is not a bivariate Reed-Muller
codeword, and so π‘Ÿ (being close to 𝑐) is not close to a bivariate Reed-Muller codeword (by
the distance of the code).
Proof of Theorem 7.2.1. Observe that, since (π’ž1 )π‘‘β†—π‘š ⊂ (π’ž0 )1β†—π‘š , we have 𝛿(π‘Ÿ, (π’ž0 )1β†—π‘š ) ≤
103
𝛿(π‘Ÿ, (π’ž1 )π‘‘β†—π‘š ). If 𝛿(π‘Ÿ, (π’ž0 )1β†—π‘š ) ≥ min
{︁
}︁
, then we are done since
𝛿2
, 𝛿(π‘Ÿ, (π’ž1 )π‘‘β†—π‘š )
4
𝜌 = E𝑒 [𝛿 (π‘Ÿ|𝑒 , π’ž1 )]
[︁ (︁
)︁]︁
≥ E𝑒 𝛿 π‘Ÿ|𝑒 , (π’ž0 )1↗𝑑
≥ 𝛼2 · 𝛿(π‘Ÿ, (π’ž0 )1β†—π‘š )
}οΈ‚
{οΈ‚ 2
𝛿
π‘‘β†—π‘š
, 𝛿(π‘Ÿ, (π’ž1 )
)
≥ 𝛼2 · min
4
𝛼2 𝛿 2
· 𝛿(π‘Ÿ, (π’ž1 )π‘‘β†—π‘š ).
≥
4
1β†—π‘š
Therefore, suppose 𝛿(π‘Ÿ, (π’ž0 )
) < min
{︁
}︁
𝛿2
, 𝛿(π‘Ÿ, (π’ž1 )π‘‘β†—π‘š )
4
𝛿2
.
4
1↗𝑑
codeword to π‘Ÿ, so that 𝑓 ∈
/ (π’ž1 )π‘‘β†—π‘š and 𝛿(π‘Ÿ, 𝑓 ) <
which 𝑓 |𝑒 ∈
/ π’ž1 , then, since π’ž1 is a subcode of (π’ž0 )
. Let 𝑓 ∈ (π’ž0 )1β†—π‘š be the nearest
If 𝑒 is a 𝑑-dimensional subspace for
, 𝛿(π‘Ÿ|𝑒 , π’ž1 ) ≥ 𝛿 − 𝛿(π‘Ÿ|𝑒 , 𝑓 |𝑒 ). Since
𝛿2
E𝑒 [𝛿 (π‘Ÿ|𝑒 , 𝑓 |𝑒 )] = 𝛿(π‘Ÿ, 𝑓 ) < ,
4
by Markov,
[οΈ‚
]οΈ‚
𝛿
𝛿
Pr 𝛿(π‘Ÿ|𝑒 , 𝑓 |𝑒 ) ≥
≤
𝑒
2
2
By Corollary 6.2.8,
Pr [𝑓 |𝑒 ∈ π’ž1 ] ≤ 1 − 𝛿 + 2π‘ž −1 .
𝑒
By the union bound, it follows that for at least 2𝛿 −2π‘ž −1 fraction of the 𝑑-dimensional 𝑒 ⊆ Fπ‘š
π‘ž ,
it holds that
𝛿
𝛿(π‘Ÿ|𝑒 , π’ž1 ) ≥ 𝛿 − 𝛿(π‘Ÿ|𝑒 , 𝑓 |𝑒 ) ≥ .
2
Therefore,
(οΈ‚
𝜌 = E𝑒 [𝛿 (π‘Ÿ|𝑒 , π’ž1 )] ≥
104
)οΈ‚
𝛿
𝛿
−1
− 2π‘ž
· .
2
2
7.3
Nikodym Sets
In this section, we use the lifted Reed-Solomon code to improve upon the polynomial method
to show that Nikodym sets are large.
π‘š
Definition 7.3.1. A subset 𝑆 ⊆ Fπ‘š
π‘ž is a Nikodym set if for every x ∈ Fπ‘ž there exists a
nonzero a ∈ Fπ‘š
π‘ž such that x + 𝑑a ∈ 𝑆 for every nonzero 𝑑 ∈ Fπ‘ž .
Before we prove our lower bound, let us recall how the polynomial method yields a good
lower bound in the first place. Let 𝑆 be a Nikodym set. Let 𝑅 be the rate of the code
RM(π‘ž, π‘ž − 2, π‘š), which is approximately 𝑅 ≈
1
.
π‘š!
Now, assume for the sake of contradiction
that |𝑆| < π‘…π‘ž π‘š . Then there exists a nonzero 𝑓 with deg(𝑓 ) ≤ π‘ž − 2 that vanishes on every
point of 𝑆. This is because the coefficients of 𝑓 provide π‘…π‘ž π‘š degrees of freedom, but we only
have |𝑆| linear constraints. Now, we claim that 𝑓 actually vanishes everywhere, contradicting
the fact that it is nonzero. To see this, let x ∈ Fπ‘š
π‘ž be an arbitrary point. By the Nikodym
property of 𝑆, there is a line 𝐿 through 𝑆 such that (𝐿 βˆ– {x}) ⊆ 𝑆. Consider the univariate
polynomial 𝑓 |𝐿 . It vanishes on π‘ž − 1 points 𝐿 βˆ– {x} but has degree deg(𝑓 |𝐿 ) < π‘ž − 1, so 𝑓 |𝐿
is identically zero. But then 𝑓 (x) = 𝑓 |𝐿 (x) = 0. Therefore |𝑆| ≥ π‘…π‘ž π‘š ≈
π‘žπ‘š
.
π‘š!
Observe that, in the above argument, the only property we actually used was the fact
that, on every line 𝐿, the restriction 𝑓 |𝐿 has degree deg(𝑓 |𝐿 ) < π‘ž − 1. So pulling 𝑓 from
the Reed-Muller code was needlessly restrictive. We could use a larger code if possible, and
in fact the largest code satisfying this property is none other than the lifted Reed-Solomon
code!
Theorem 7.3.2. Let 𝑝 be a prime and let π‘ž = 𝑝𝑠 . If 𝑆 ⊆ Fπ‘š
π‘ž is a Nikodym set, then
(︁
)︁
⌈οΈ€
⌉οΈ€
π‘šπ‘
|𝑆| ≥ 1 − 𝑒−(𝑠+𝑏)/(𝑏𝑝 ) · π‘ž π‘š where 𝑏 = log𝑝 π‘š + 1. In particular, if 𝑝 is fixed and π‘ž → ∞,
then |𝑆| ≥ (1 − π‘œ(1)) · π‘ž π‘š .
Proof. Let 𝑅 = 1−𝑒−(𝑠+𝑏)/(𝑏𝑝
π‘šπ‘ )
. Suppose, for the sake of contradiction, that |𝑆| < π‘…π‘ž π‘š . Let
π’ž = RS(π‘ž, π‘ž − 2)1β†—π‘š . This is the construction of Section 7.1 with 𝑐 = 𝑠. By Theorem 7.1.11,
dimFπ‘ž (π’ž) ≥ π‘…π‘ž π‘š > |𝑆|, so there exists a nonzero 𝑐 ∈ π’ž such that 𝑐(x) = 0 for every x ∈ 𝑆.
This follows from the fact that each equation of the form 𝑐(x) = 0 is a linear constraint on
105
𝑐, and there are |𝑆| linear constraints on 𝑐 which has dimFπ‘ž (π’ž) > |𝑆| degrees of freedom. We
proceed to show that 𝑐 = 0, a contradiction. Let x ∈ Fπ‘š
π‘ž . Since 𝑆 is a Nikodym set, there is
a nonzero a ∈ Fπ‘š
π‘ž such that x + 𝑑a ∈ 𝑆 for every nonzero 𝑑 ∈ Fπ‘ž . Define 𝑐a (𝑑) , 𝑐(x + 𝑑a).
Since 𝑐 ∈ π’ž, we have deg(𝑐a ) ≤ π‘ž − 2. However, 𝑐a (𝑑) = 0 for every 𝑑 ΜΈ= 0, so 𝑐a vanishes on
π‘ž − 1 > deg(𝑐a ) points, so 𝑐a = 0. In particular, 𝑐(x) = 𝑐a (0) = 0.
106
Appendix A
Algebra Background
This appendix contains well-known facts about arithmetic over finite fields and tensor codes
that are used in the thesis.
A.1
Arithmetic over finite fields
Much of our work involves manipulating polynomials over finite fields. Expanding multinomials over a finite field, which has positive characteristic, is different from expanding over a
field of zero characteristic, because binomial coefficients may vanish due to the characteristic
of the field. For example, over a field of characteristic 2, the binomial (𝑋 + π‘Œ )2 expands to
𝑋 2 + 2π‘‹π‘Œ + π‘Œ 2 = 𝑋 2 + π‘Œ 2 .
The key fact we will use is (a generalization of) Lucas’ Theorem, which tells us what a
multinomial coefficient looks like modulo a prime 𝑝.
Theorem A.1.1 (Generalized Lucas’ Theorem). Let π‘Ž0 , π‘Ž1 , . . . , π‘Žπ‘š ∈ N and let 𝑝 be a prime.
∑οΈ€
(𝑖) (𝑖)
Write π‘Žπ‘— = 𝑛𝑖=1 π‘Ž(𝑖) · 𝑝𝑖 where π‘Žπ‘— , 𝑏𝑗 ∈ J𝑝K for every 0 ≤ 𝑖 ≤ 𝑛 and 0 ≤ 𝑗 ≤ π‘š. Then
(οΈ‚
π‘Ž0
π‘Ž1 , . . . , π‘Ž π‘š
)οΈ‚
≡
(𝑖)
𝑛 (οΈ‚
∏︁
π‘Ž0
(𝑖)
𝑖=0
)οΈ‚
(𝑖)
π‘Ž1 , . . . , π‘Ž π‘š
Proof. Let 𝑆 = {1, . . . , π‘Ž0 }. Partition 𝑆 into 𝑆𝑖,𝑗
107
(mod 𝑝).
[︁ ]︁
(𝑖)
⊂ 𝑆, for each 𝑖 ∈ J𝑛 + 1K and 𝑗 ∈ π‘Ž0 ,
[︁ ]︁
(𝑖)
π‘Ž
(𝑖)
of size |𝑆𝑖,𝑗 | = 𝑝𝑖 . For each 𝑖 ∈ J𝑛 + 1K and 𝑗 ∈ π‘Ž0 , let 𝐺𝑖,𝑗 = Z𝑝𝑖0 act on 𝑆𝑖,𝑗 be cyclic
⨁︀𝑛 β¨οΈ€π‘Ž0(𝑖)
permutation. Let 𝐺 =
𝑗=1 𝐺𝑖,𝑗 , which acts on 𝑆. Let 𝐢 be the collection of all
𝑖=0
(οΈ€ π‘Ž0 )οΈ€
π‘š-colorings 𝑐 : 𝑆 → [π‘š] such that |𝑐−1 (π‘˜)| = π‘Žπ‘˜ . Observe that |𝐢| = π‘Ž1 ,...,π‘Ž
, and the
π‘š
action of 𝐺 on 𝑆 naturally induces an action of 𝐺 on 𝐢, as follows: if 𝑐 ∈ 𝐢, and 𝑔 ∈ 𝐺,
then for π‘₯ ∈ 𝑆, (𝑔𝑐)(π‘₯) = 𝑐(𝑔 −1 π‘₯). Since |𝐺| is a power of 𝑝, so is the size of any orbit of 𝐢
(οΈ€ π‘Ž0 )οΈ€
under 𝐺. So, to compute π‘Ž1 ,...,π‘Ž
(mod 𝑝), it suffices to count the number of fixed points
π‘š
of 𝐢 under 𝐺. Since 𝐺 fixes each 𝑆𝑖,𝑗 , it is easy to see that a coloring is fixed under 𝐺 if
and only if each 𝑆𝑖,𝑗 is monochromatic. It therefore suffices to show that for each 𝑖 ∈ J𝑛 + 1K
(𝑖)
and each π‘˜ ∈ [π‘š], there are exactly π‘Žπ‘˜ values for 𝑗 such that 𝑆𝑖,𝑗 has color π‘˜. Fix π‘˜ ∈ [π‘š].
∑οΈ€
(𝑖)
Observe that there are π‘Žπ‘˜ = 𝑛𝑖=1 π‘Žπ‘˜ elements of color π‘˜ in total, and each set 𝑆𝑖,𝑗 of color
π‘˜ contributes 𝑝𝑖 elements. The claim then follows easily by induction on 𝑛 − 𝑖.
In particular, it characterizes which multinomial coefficients are nonzero modulo 𝑝.
𝑑
𝑒1 ,...,𝑒𝑛
(οΈ€
Corollary A.1.2. If 𝑑, 𝑒1 , . . . , 𝑒𝑛 ∈ N, then
)οΈ€
ΜΈ≡ 0 (mod 𝑝) only if (𝑒1 , . . . , 𝑒𝑛 ) ≤𝑝 𝑑.
This allows us to expand multinomials over a finite field and know which terms of the
usual expansion disappear.
Corollary A.1.3. Let 𝑝 be a prime and let F be a field of characteristic 𝑝. Let 𝑑 ∈ N and
let π‘₯1 , . . . , π‘₯𝑛 ∈ F. Then
(οΈƒ
𝑛
∑︁
𝑖=1
)︃𝑑
π‘₯𝑖
=
∑︁
(𝑒1 ,...,𝑒𝑛 )≤𝑝 𝑑
(οΈ‚
)οΈ‚ ∏︁
𝑛
𝑑
π‘₯𝑒 𝑖
𝑒1 , . . . , 𝑒𝑛 𝑖=1 𝑖
Expanding multinomials over finite fields is particularly important for us since we frequently look at the restricting polynomials to affine subspaces, which entails composing with
an affine function.
Proposition A.1.4. Let 𝑝 be a prime and let F be a field of characteristic 𝑝. Let X =
(𝑋1 , . . . , π‘‹π‘š ) and let Y ∈ (π‘Œ1 , . . . , π‘Œπ‘‘ ). Let 𝑓 (X) ∈ F[X], and let A ∈ Fπ‘š×𝑑 and b ∈ Fπ‘š
π‘ž . If
108
𝑓 (X) =
∑οΈ€
d∈𝐷
𝑓d · Xd , then
𝑓 (AY + b) =
∑︁
d∈supp(𝑓 )
(οΈƒ
)οΈƒ 𝑑
π‘š
𝑑
∑︁ (οΈ‚ d )οΈ‚ ∏︁
∏︁
∏︁ β€–E β€–
𝑒
𝑓d ·
𝑏𝑒𝑖 𝑖0
π‘Žπ‘–π‘—π‘–π‘— ·
π‘Œπ‘— *𝑗
E 𝑖=1
𝑗=1
𝑗=1
E≤ d
𝑝
where E in the summation is a π‘š × (𝑑 + 1) matrix with rows indexed by [π‘š] and columns
indexed by J𝑑 + 1K.
Proof.
𝑓 (AY + b) =
∑︁
𝑓d · (AY + b)d
d∈𝐷
=
∑︁
d∈𝐷
𝑓d ·
(οΈƒ 𝑑
π‘š
∏︁
∑︁
𝑖=1
π‘Žπ‘–π‘— π‘Œπ‘— + 𝑏𝑖
𝑗=1
)οΈ‚
𝑑
∏︁
𝑑𝑖
𝑒
𝑒
𝑒𝑖0
(Corollary A.1.3) =
𝑓d ·
𝑏𝑖
π‘Žπ‘–π‘—π‘–π‘— π‘Œπ‘— 𝑖𝑗
𝑒𝑖0 , 𝑒𝑖1 , . . . , 𝑒𝑖𝑑
𝑖=1 (𝑒𝑖0 ,𝑒𝑖1 ,...,𝑒𝑖𝑑 )≤𝑝 𝑑𝑖
𝑗=1
d∈𝐷
)οΈƒ
(οΈƒ
(οΈ‚
)οΈ‚
𝑑
π‘š
𝑑
∏︁
∑︁ d ∏︁
∏︁
∑︁
β€–E β€–
𝑒
π‘Œπ‘— *𝑗
𝑓d ·
𝑏𝑒𝑖 𝑖0
π‘Žπ‘–π‘—π‘–π‘— ·
=
E
𝑗=1
𝑖=1
𝑗=1
E≤ d
d∈𝐷
∑︁
π‘š
∏︁
)︃𝑑𝑖
∑︁
(οΈ‚
𝑝
A.2
Tensor codes
The tensor product is a natural operation in linear algebra that, when applied to two linear
codes, produces a new linear code in a natural way. There are many equivalent ways to define
the tensor product of two codes. Since in this thesis we think of codes as linear subspaces
of functions in {Fπ‘š
π‘ž → Fπ‘ž }, we define the tensor product in this context.
Definition A.2.1. Let 𝑛 ≥ 2, let 𝑑1 , . . . , 𝑑𝑛 ≥ 1 and π‘š =
∑︀𝑛
𝑖=1 𝑑𝑖 ,
and for each 𝑖 ∈ [𝑛],
let the code π’žπ‘– ⊆ {Fπ‘‘π‘žπ‘– → Fπ‘ž } be linear and let 𝑉𝑖,a ⊆ Fπ‘š
π‘ž be the 𝑑𝑖 dimensional subspace
consisting of all points where the 𝑖-th block (of 𝑑𝑖 coordinates) is free and all the [𝑛] βˆ– {𝑖}
∏οΈ€
𝑑
blocks are fixed to a ∈ 𝑗̸=𝑖 Fπ‘žπ‘— . The tensor product code π’ž1 ⊗ · · · ⊗ π’žπ‘› ⊆ {Fπ‘š
π‘ž → Fπ‘ž } is the
109
code
βƒ’
}οΈƒ
βƒ’
∏︁
βƒ’
𝑓 : Fπ‘š
Fπ‘‘π‘žπ‘—
π‘ž → Fπ‘ž βƒ’ 𝑓 |𝑉𝑖,a ∈ π’žπ‘– for every 𝑖 ∈ [𝑛] and a ∈
βƒ’
{οΈƒ
π’ž1 ⊗ · · · ⊗ π’žπ‘› ,
𝑗̸=𝑖
𝑛
Define π’ž ⊗𝑛
⏞
⏟
, π’ž ⊗ · · · ⊗ π’ž.
The following characterization of tensor product codes will be helpful.
Proposition A.2.2. Let 𝑛 ≥ 2, let 𝑑1 , . . . , 𝑑𝑛 ≥ 1 and π‘š =
∑︀𝑛
𝑖=1 𝑑𝑖 ,
and for each 𝑖 ∈ [𝑛], let
the code π’žπ‘– ⊆ {Fπ‘‘π‘žπ‘– → Fπ‘ž } be linear, and let X𝑖 = (𝑋𝑖1 , . . . , 𝑋𝑖𝑑𝑖 ) be variables. Then
π’ž1 ⊗ · · · ⊗ π’žπ‘› = spanFπ‘ž
{οΈƒ 𝑛
∏︁
𝑖=1
βƒ’
}οΈƒ
βƒ’
βƒ’
𝑓𝑖 (X𝑖 ) βƒ’ 𝑓𝑖 ∈ π’žπ‘–
βƒ’
Corollary A.2.3. If π’ž ⊆ {Fπ‘‘π‘ž → Fπ‘ž } has a degree set Deg(π’ž), and 𝑛 ≥ 1, then π’ž ⊗𝑛 has
degree set Deg(π’ž ⊗𝑛 ) = Deg(π’ž)𝑛 . In particular, if π’ž is linear affine-invariant, and Fπ‘ž has
characteristic 𝑝, then π’ž ⊗𝑛 has a 𝑝-shadow-closed degree set.
Proposition A.2.4. Let π’ž1 and π’ž2 be codes with distance 𝛿1 and 𝛿2 repectively. Then 𝛿(π’ž1 ⊗
π’ž2 ) is at least 𝛿1 𝛿2 . In particular, 𝛿 (π’ž ⊗𝑛 ) ≥ 𝛿(π’ž)𝑛 .
The following is a statement about the erasure decoding properties of tensor product
codes.
{οΈ€
}οΈ€
Proposition A.2.5. Let π’ž = π’ž1 ⊗ . . . π’žπ‘› ∈ Fπ‘š
and 𝑆 ⊆ Fπ‘š
π‘ž → Fπ‘ž
π‘ž be a subset such that
∏οΈ€
𝑑
for every 𝑖 ∈ [𝑛] and a ∈ 𝑗̸=𝑖 Fπ‘žπ‘— satisfy |𝑆 ∩ 𝑉𝑖,a | ≥ (1 − 𝛿(π’žπ‘– ))π‘ž 𝑑𝑖 . Let π‘Ÿ : 𝑆 → Fπ‘ž be such
∏οΈ€
𝑑
that for every 𝑖 ∈ [𝑛] and a ∈ 𝑗̸=𝑖 Fπ‘žπ‘— satisfy that π‘Ÿ|𝑆∩𝑉𝑖,a can be extended into a codeword
of π’žπ‘– on 𝑉𝑖,a . Then there exists a unique π‘Ÿ′ ∈ π’ž such that π‘Ÿ′ |𝑆 = π‘Ÿ.
110
Appendix B
Finite field geometry
This appendix describes some basic structural facts about affine maps, as well as basic
geometry of affine subspaces over finite fields.
B.1
Affine maps
The following proposition allows us to decompose an arbitrary affine map (not necessarily
injective) into a composition of a linear map and an injective affine map.
Proposition B.1.1. Let 𝑑 ≤ π‘š. For every affine map 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž , there exists a linear
′′
′
map 𝐴′ : Fπ‘‘π‘ž → Fπ‘‘π‘ž and injective affine map 𝐴′′ : Fπ‘‘π‘ž → Fπ‘š
π‘ž such that 𝐴 = 𝐴 ∘ 𝐴 .
π‘š
′′
′
Proof. If 𝐴 : x ↦→ 𝐴𝐿 (x) + b for some linear 𝐴𝐿 : Fπ‘‘π‘ž → Fπ‘š
π‘ž and b ∈ Fπ‘ž , and if 𝐴𝐿 = 𝐴𝐿 ∘ 𝐴
′
𝑑
𝑑
′′
′
for some injective affine 𝐴′′𝐿 : Fπ‘‘π‘ž → Fπ‘š
π‘ž and linear 𝐴 : Fπ‘ž → Fπ‘ž , then 𝐴 = 𝐴 ∘ 𝐴 where
𝐴′′ : x ↦→ 𝐴′′𝐿 (x) + b. Therefore, we reduce to the case where 𝐴 is linear, i.e. 𝐴(0) = 0.
Fix a basis e1 , . . . , e𝑑 ∈ Fπ‘‘π‘ž . For 𝑗 ∈ [𝑑], let v𝑗 , 𝐴(e𝑖 ). After re-labeling, we may
assume without loss of generality that, for some 0 ≤ π‘Ÿ ≤ 𝑑, the vectors v1 , . . . , vπ‘Ÿ are linearly
independent and vπ‘Ÿ+1 , . . . , v𝑑 ∈ span{v1 , . . . , vπ‘Ÿ }. There exist (unique) π‘Žπ‘–π‘— ∈ Fπ‘ž , for 𝑖 ∈ [π‘Ÿ]
∑οΈ€
and π‘Ÿ + 1 ≤ 𝑗 ≤ 𝑑, such that v𝑗 = π‘Ÿπ‘–=1 π‘Žπ‘–π‘— v𝑗 . Define 𝐴′ : Fπ‘‘π‘ž → Fπ‘‘π‘ž as the unique linear map
∑οΈ€
such that 𝐴′ (e𝑖 ) = e𝑖 for 𝑖 ∈ [π‘Ÿ] and 𝐴′ (e𝑗 ) = π‘Ÿπ‘–=1 π‘Žπ‘–π‘— e𝑖 for π‘Ÿ + 1 ≤ 𝑗 ≤ 𝑑. For 𝑖 ∈ [π‘Ÿ], set
111
w𝑖 , v𝑖 , and extend w1 , . . . , wπ‘Ÿ to a set of linearly independent vectors w1 , . . . , w𝑑 ∈ Fπ‘š
π‘ž .
′′
Define 𝐴′′ : Fπ‘‘π‘ž → Fπ‘š
π‘ž to be the unique linear map such that 𝐴 (e𝑗 ) = w𝑗 .
By construction, 𝐴′′ is injective since its image is span{w1 , . . . , w𝑑 }. Moreover, for every
𝑖 ∈ [π‘Ÿ],
(𝐴′′ ∘ 𝐴′ )(e𝑖 ) = 𝐴′′ (𝐴′ (e𝑖 )) = 𝐴′′ (e𝑖 ) = w𝑖 = v𝑖 = 𝐴(e𝑖 )
and for every π‘Ÿ + 1 ≤ 𝑗 ≤ 𝑑,
(οΈƒ
′′
′
′′
′
(𝐴 ∘ 𝐴 )(e𝑗 ) = 𝐴 (𝐴 (e𝑗 )) = 𝐴
′′
π‘Ÿ
∑︁
)οΈƒ
π‘Žπ‘–π‘— e𝑖
=
𝑖=1
π‘Ÿ
∑︁
′′
π‘Žπ‘–π‘— 𝐴 (e𝑖 ) =
𝑖=1
π‘Ÿ
∑︁
π‘Žπ‘–π‘— v𝑖 = v𝑗 = 𝐴(e𝑗 ),
𝑖=1
so, by linearity, 𝐴′′ ∘ 𝐴′ = 𝐴.
The next three propositions describe the behavior of random affine maps. We are particularly interested in how 𝛿(𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴) behaves, where 𝑓, 𝑔 are distinct functions and 𝐴
is a random subspace. Proposition B.1.3 covers the case where 𝐴(0) is fixed, while Proposition B.1.4 allows 𝐴 to be truly free.
π‘š
𝑑
Proposition B.1.2. Let 𝑑 ≤ π‘š, let x ∈ Fπ‘š
π‘ž , and let 𝐴 : Fπ‘ž → Fπ‘ž be a random affine map
such that 𝐴(0) = x. For any fixed nonzero y ∈ Fπ‘‘π‘ž , 𝐴(y) is a uniformly random point in Fπ‘š
π‘ž .
Proof. We can choose 𝐴 by choosing π‘Žπ‘–π‘— ∈ Fπ‘ž for 𝑖 ∈ [π‘š], 𝑗 ∈ [𝑑] independently and uniformly
at random, and setting
⎑
⎀
π‘Ž
· · · π‘Ž1𝑑
⎒ 11
βŽ₯
.. βŽ₯
⎒ ..
.
.
𝐴(y) = x + ⎒ .
.
. βŽ₯y
⎣
⎦
π‘Žπ‘š1 · · · π‘Žπ‘šπ‘‘
The 𝑖-th coordinate of 𝐴(y) is therefore π‘₯𝑖 +
∑︀𝑑
𝑗=1
π‘Žπ‘–π‘— 𝑦𝑗 which is uniformly random in Fπ‘ž ,
and the coordinates of 𝐴(y) are independent since the π‘Žπ‘–π‘— are independent.
𝑑
π‘š
Proposition B.1.3. Let 𝑑 ≤ π‘š, let x ∈ Fπ‘š
π‘ž , and let 𝐴 : Fπ‘ž → Fπ‘ž be a random affine map
−𝑑
such that 𝐴(0) = x. If 𝑓, 𝑔 : Fπ‘š
π‘ž → Fπ‘ž , then E𝐴 [𝛿(𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴)] ≤ 𝛿(𝑓, 𝑔) + π‘ž .
112
Proof. We have
[︁
[οΈ€
]οΈ€]︁
E𝐴 [𝛿(𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴)] = E𝐴 Ey∈Fπ‘‘π‘ž 1𝑓 (𝐴(y))ΜΈ=𝑔(𝐴(y))
[οΈ€ [οΈ€
]οΈ€]οΈ€
= Ey∈Fπ‘‘π‘ž E𝐴 1𝑓 (𝐴(y))ΜΈ=𝑔(𝐴(y))
[οΈ€ [οΈ€
]οΈ€]οΈ€
≤ π‘ž −𝑑 + Ey∈Fπ‘‘π‘ž βˆ–{0} E𝐴 1𝑓 (𝐴(y))ΜΈ=𝑔(𝐴(y))
(B.1)
(Proposition B.1.2) = π‘ž −𝑑 + 𝛿(𝑓, 𝑔).
(B.4)
(B.2)
(B.3)
Proposition B.1.4. Let 𝑑 ≤ π‘š, and let 𝐴 : Fπ‘‘π‘ž → Fπ‘š
π‘ž be a uniformly random affine map. If
−𝑑
𝑓, 𝑔 : Fπ‘š
≤
π‘ž → Fπ‘ž and 𝛿 , 𝛿(𝑓, 𝑔), then 𝛿(𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴) has mean 𝛿 and variance 𝛿(1 − 𝛿)π‘ž
π‘ž −𝑑 /4.
𝑑
π‘š
Proof. Observe that for any fixed y ∈ Fπ‘š
π‘ž , if 𝐴 : Fπ‘ž → Fπ‘ž is uniformly random, then
𝐴(y) is a uniformly random point in Fπ‘š
π‘ž , and for distinct y, the 𝐴(y) are pairwise independent. Therefore, 1𝑓 (𝐴(y))ΜΈ=𝑔(𝐴(y)) has mean 𝛿 and variance 𝛿(1 − 𝛿). Since 𝛿(𝑓 ∘ 𝐴, 𝑔 ∘ 𝐴) =
∑οΈ€
π‘ž −𝑑 y∈Fπ‘‘π‘ž 1𝑓 (𝐴(y))ΜΈ=𝑔(𝐴(y)) , it follows that the mean is 𝛿 and variance is 𝛿(1−𝛿)π‘ž −𝑑 ≤ π‘ž −𝑑 /4.
Finally, we prove that if 𝑓, 𝑔 are distinct polynomials on a plane, then they cannot agree
on too many lines.
Proposition B.1.5. Let 𝑓, 𝑔 : F2π‘ž → Fπ‘ž be distinct. For any a, b ∈ F2π‘ž , define 𝑓a,b , 𝑔a,b :
Fπ‘ž → Fπ‘ž by 𝑓a,b (𝑇 ) , 𝑓 (a𝑇 + b) and similarly 𝑔a,b (𝑇 ) , 𝑔(a𝑇 + b). Then 𝑓a,b = 𝑔a,b for at
most 2π‘ž 3 pairs (a, b) ∈ F2π‘ž × F2π‘ž .
Proof. Let β„Ž = 𝑓 − 𝑔 and β„Ža,b = 𝑓a,b − 𝑔a,b . We have four cases to consider.
1. π‘Ž1 , π‘Ž2 ΜΈ= 0: if β„Ža,b = 0, then the polynomial β„Ž(𝑋, π‘Œ ) is divisible by π‘Œ − π‘Žπ‘Ž21 𝑋 − π‘Žπ‘Ž21𝑏1 − 𝑏2 .
There are π‘ž(π‘ž − 1) pairs (a, b) which correspond to this factor (π‘ž − 1 choices for a and
then π‘ž choices for b given a).
2. π‘Ž1 ΜΈ= 0, π‘Ž2 = 0: if β„Ža,b = 0, then the polynomial β„Ž(𝑋, π‘Œ ) is divisible by π‘Œ − 𝑏2 . There
are π‘ž 2 pairs corresponding to this factor (π‘ž choices for π‘Ž1 and π‘ž choices for 𝑏1 ).
113
3. π‘Ž1 = 0, π‘Ž2 ΜΈ= 0: same as previous case, by symmetry.
4. π‘Ž1 = π‘Ž2 = 0: there are at most π‘ž 2 such pairs (π‘ž 2 choices for b).
Say a pair (a, b) is bad if β„Ža,b = 0. Trivially, deg(β„Ž) ≤ 2(π‘ž − 1), so β„Ž has at most 2(π‘ž − 1)
linear factors. Each factor from cases 1, 2, and 3 corresponds to at most π‘ž 2 pairs (a, b),
resulting in at most 2π‘ž 2 (π‘ž − 1) bad pairs (a, b) from those cases, and there are at most π‘ž 2
bad pairs from case 4. So, the total number of bad pairs is at most 2π‘ž 2 (π‘ž − 1) + π‘ž 2 ≤ 2π‘ž 3 .
B.2
Affine subspaces
The next two lemmas analyze the behavior of random affine subspaces in Fπ‘š
π‘ž . Lemma B.2.1
shows that a low-dimensional subspace is disjoint from almost all low-dimensional affine
subspaces.
Lemma B.2.1. Let 𝑑 ≤ π‘˜ < π‘š. Let 𝑒 ⊆ Fπ‘š
π‘ž be a fixed affine subspace of dimension 𝑑, and
let 𝑣 ⊆ Fπ‘š
π‘ž be a uniformly random affine subspace of dimension π‘˜. Then Pr𝑣 [𝑒 ∩ 𝑣 ΜΈ= ∅] <
π‘ž −(π‘š−π‘˜−𝑑) .
Proof. By affine symmetry, we may assume that 𝑣 is fixed and 𝑒 is random. Furthermore,
we can assume that 𝑣 = (0, 𝐡), where 𝐡 is a basis and hence |𝐡| = π‘˜. We choose 𝑒 by
choosing random x ∈ Fπ‘š
π‘ž , random basis 𝐴 = {a1 , . . . , a𝑑 }, and setting 𝑒 , (x, 𝐴). Let 𝐸 be
the event that 𝑒 ∩ 𝑣 ΜΈ= ∅.
Re-arrange a1 , . . . , a𝑑 so that for some 0 ≤ 𝑠 ≤ 𝑑 − 1, a𝑖 ∈ span(𝐡) if and only if 𝑖 ≤ 𝑠.
Note that 𝐸 holds if and only if there exist 𝑐1 , . . . , 𝑐𝑑 ∈ Fπ‘š
π‘ž such that x + 𝑐1 a1 + · · · + 𝑐𝑑 a𝑑 ∈
π‘š−π‘˜
span(𝐡). Let 𝑃 : Fπ‘š
be the linear map that projects onto the last π‘š−π‘˜ coordinates.
π‘ž → Fπ‘ž
Note that ker(𝑃 ) = 𝐡. For each 𝑖 ∈ [𝑑], let a′𝑖 , 𝑃 a𝑖 ∈ Fπ‘š−π‘˜
. Then 𝐸 holds if and only if
π‘ž
𝑃 x ∈ span(a′𝑠+1 , . . . , a′𝑑 ). Therefore, there are at most π‘ž 𝑑 choices for 𝑃 x, hence at most π‘ž π‘˜+𝑑
choices for x, out of π‘ž π‘š total choices for x, so Pr[𝐸] ≤
π‘ž π‘˜+𝑑
π‘žπ‘š
= π‘ž −(π‘š−π‘˜−𝑑) .
Lemma B.2.2 shows that if π‘˜ β‰ͺ π‘š, then π‘˜ random vectors are likely to be linearly
independent, and in particular two random low-dimensional subspaces through a fixed point
114
x are likely to intersect only at x.
Lemma B.2.2. Let π‘˜ < π‘š and a1 , . . . , aπ‘˜ ∈ Fπ‘š
π‘ž be uniformly chosen vectors. Then the
probability that {a𝑖 }π‘˜π‘–=1 are linearly independent is at least 1 − π‘ž −(π‘š−π‘˜) . In particular, the
probability that two 𝑑-dimensional subspaces through a point x ∈ Fπ‘š
π‘ž will intersect only on x
is at least 1 − π‘ž −(π‘š−2𝑑) .
Proof. The probability that a𝑖+1 ∈
/ span {a1 , . . . , a𝑖 } given that the latter are linearly independent is 1 − π‘ž −(π‘š−𝑖) . Therefore the probability that all of them are independent is
π‘˜−1
∏︁
(1 − π‘ž
𝑖=0
−(π‘š−𝑖)
)≥1−
π‘˜−1
∑︁
π‘ž
−(π‘š−𝑖)
=1−π‘ž
−(π‘š−π‘˜)
π‘˜
∑︁
π‘ž −𝑖 ≥ 1 − π‘ž −(π‘š−π‘˜) .
𝑖=1
𝑖=0
For the last part, observe that choosing two 𝑑-dimensional subspaces through x is equivalent
to choose 2𝑑 basis vectors, given that each 𝑑 are linearly independent. So the probability that
they intersect only on x, is the same as that those vectors are linearly independent. Hence,
by the first part, this probability is at least 1 − π‘ž −(π‘š−2𝑑) .
115
116
Bibliography
[AKK+ 05]
Noga Alon, Tali Kaufman, Michael Krivelevich, Simon Litsyn, and Dana
Ron. Testing Reed-Muller codes. IEEE Transactions on Information Theory,
51(11):4032–4039, 2005.
[ALM+ 98]
Sanjeev Arora, Carsten Lund, Rajeev Motwani, Madhu Sudan, and Mario
Szegedy. Proof verification and the hardness of approximation problems. Journal of the ACM, 45(3):501–555, May 1998.
[Aro94]
Sanjeev Arora. Probabilistic checking of proofs and the hardness of approximation problems. PhD thesis, University of California at Berkeley, 1994.
[AS98]
Sanjeev Arora and Shmuel Safra. Probabilistic checking of proofs: A new
characterization of NP. Journal of the ACM, 45(1):70–122, January 1998.
[AS03]
Sanjeev Arora and Madhu Sudan. Improved low degree testing and its applications. Combinatorica, 23(3):365–426, 2003. Preliminary version in Proceedings
of ACM STOC 1997.
[BET10]
Avraham Ben-Aroya, Klim Efremenko, and Amnon Ta-Shma. Local list decoding with a constant number of queries. In 51st IEEE Symposium on Foundations
of Computer Science (FOCS), 2010.
[BGK+ 13]
Eli Ben-Sasson, Ariel Gabizon, Yohay Kaplan, Swastik Kopparty, and Shubhangi Saraf. A new family of locally correctable codes based on degree-lifted
algebraic geometry codes. In Dan Boneh, Tim Roughgarden, and Joan Feigenbaum, editors, Symposium on Theory of Computing Conference, STOC’13, Palo
Alto, CA, USA, June 1-4, 2013, pages 833–842. ACM, 2013.
[BGM+ 11a] Eli Ben-Sasson, Elena Grigorescu, Ghid Maatouk, Amir Shpilka, and Madhu
Sudan. On sums of locally testable affine invariant properties. In APPROXRANDOM, pages 400–411, 2011.
[BGM+ 11b] Eli Ben-Sasson, Elena Grigorescu, Ghid Maatouk, Amir Shpilka, and Madhu
Sudan. On sums of locally testable affine invariant properties. Electronic Colloquium on Computational Complexity (ECCC), 18:79, 2011.
117
[BK09]
K. Brander and S. Kopparty. List-decoding Reed-Muller over large fields upto
the Johnson radius. Manuscript, 2009.
[BKS+ 10]
Arnab Bhattacharyya, Swastik Kopparty, Grant Schoenebeck, Madhu Sudan,
and David Zuckerman. Optimal testing of Reed-Muller codes. In FOCS, pages
488–497. IEEE Computer Society, 2010.
[BLR93]
Manuel Blum, Michael Luby, and Ronitt Rubinfeld. Self-testing/correcting with
applications to numerical problems. Journal of Computer and System Sciences,
47(3):549–595, 1993.
[BSGH+ 04] Eli Ben-Sasson, Oded Goldreich, Prahladh Harsha, Madhu Sudan, and Salil
Vadhan. Robust PCPs of proximity, shorter PCPs and applications to coding.
In Proceedings of the 36th Annual ACM Symposium on Theory of Computing,
pages 1–10, New York, 2004. ACM Press.
[BSMSS11] Eli Ben-Sasson, Ghid Maatouk, Amir Shpilka, and Madhu Sudan. Symmetric LDPC codes are not necessarily locally testable. In IEEE Conference on
Computational Complexity, pages 55–65. IEEE Computer Society, 2011.
[BSS06]
Eli Ben-Sasson and Madhu Sudan. Robust locally testable codes and products
of codes. Random Structures and Algorithms, 28(4):387–402, 2006.
[BSV09a]
Eli Ben-Sasson and Michael Viderman. Composition of semi-LTCs by twowise tensor products. In Irit Dinur, Klaus Jansen, Joseph Naor, and José
D. P. Rolim, editors, APPROX-RANDOM, volume 5687 of Lecture Notes in
Computer Science, pages 378–391. Springer, 2009.
[BSV09b]
Eli Ben-Sasson and Michael Viderman. Tensor products of weakly smooth codes
are robust. Theory of Computing, 5(1):239–255, 2009. Preliminary version in
Proc. APPROX-RANDOM 2008.
[DKSS09]
Zeev Dvir, Swastik Kopparty, Shubhangi Saraf, and Madhu Sudan. Extensions
to the method of multiplicities, with applications to kakeya sets and mergers.
In FOCS, pages 181–190. IEEE Computer Society, 2009.
[DR04]
Irit Dinur and Omer Reingold. Assignment testers: Towards a combinatorial
proof of the PCP-theorem. In Proceedings of the 45th Annual IEEE Symposium
on Foundations of Computer Science, pages 155–164, Loc Alamitos, CA, USA,
2004. IEEE Press.
[DS07]
Zeev Dvir and Amir Shpilka. An improved analysis of linear mergers. Computational Complexity, 16(1):34–59, 2007.
[DSW06]
Irit Dinur, Madhu Sudan, and Avi Wigderson. Robust local testability of tensor
products of LDPC codes. In Josep Dı́az, Klaus Jansen, José D. P. Rolim,
118
and Uri Zwick, editors, APPROX-RANDOM, volume 4110 of Lecture Notes in
Computer Science, pages 304–315. Springer, 2006.
[Dvi08]
Zeev Dvir. On the size of Kakeya sets in finite fields. Journal of the American
Mathematical Society, (to appear), 2008. Article electronically published on
June 23, 2008.
[DW11]
Zeev Dvir and Avi Wigderson. Kakeya sets, new mergers, and old extractors.
SIAM J. Comput., 40(3):778–792, 2011.
[FS95]
Katalin Friedl and Madhu Sudan. Some improvements to total degree
tests. In Proceedings of the 3rd Annual Israel Symposium on Theory of
Computing and Systems, pages 190–198, Washington, DC, USA, 4-6 January 1995. IEEE Computer Society. Corrected version available online at
http://people.csail.mit.edu/madhu/papers/friedl.ps.
[GGR09]
Parikshit Gopalan, Venkatesan Guruswami, and Prasad Raghavendra. List
decoding tensor products and interleaved codes. In Proceedings of the 41st
Annual ACM Symposium on Theory of Computing, STOC 2009, Bethesda, MD,
USA, May 31 - June 2, 2009, pages 13–22, 2009.
[GHS15]
Alan Guo, Elad Haramaty, and Madhu Sudan. Robust testing of lifted codes
with applications to low-degree testing. Electronic Colloquium on Computational Complexity (ECCC), 22:34, 2015.
[GK14]
Alan Guo and Swastik Kopparty. List-decoding algorithms for lifted codes.
CoRR, abs/1412.0305, 2014.
[GKS13]
Alan Guo, Swastik Kopparty, and Madhu Sudan. New affine-invariant codes
from lifting. In Robert D. Kleinberg, editor, Innovations in Theoretical Computer Science, ITCS ’13, Berkeley, CA, USA, January 9-12, 2013, pages 529–
540. ACM, 2013.
[GS92]
Peter Gemmell and Madhu Sudan. Highly resilient correctors for multivariate
polynomials. Information Processing Letters, 43(4):169–174, September 1992.
[GS99]
Venkatesan Guruswami and Madhu Sudan. Improved decoding of ReedSolomon and algebraic-geometric codes. IEEE Transactions on Information
Theory, 45:1757–1767, 1999. Preliminary version appeared in Proc. of FOCS
1998.
[HRS13]
Elad Haramaty, Noga Ron-Zewi, and Madhu Sudan. Absolutely sound testing
of lifted codes. In Prasad Raghavendra, Sofya Raskhodnikova, Klaus Jansen,
and José D. P. Rolim, editors, Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques - 16th International Workshop,
APPROX 2013, and 17th International Workshop, RANDOM 2013, Berkeley,
119
CA, USA, August 21-23, 2013. Proceedings, volume 8096 of Lecture Notes in
Computer Science, pages 671–682. Springer, 2013.
[HSS11]
Elad Haramaty, Amir Shpilka, and Madhu Sudan. Optimal testing of multivariate polynomials over small prime fields. In Rafail Ostrovsky, editor, IEEE
52nd Annual Symposium on Foundations of Computer Science, FOCS 2011,
Palm Springs, CA, USA, October 22-25, 2011, pages 629–637. IEEE, 2011.
[JPRZ09]
Charanjit S. Jutla, Anindya C. Patthak, Atri Rudra, and David Zuckerman.
Testing low-degree polynomials over prime fields. Random Struct. Algorithms,
35(2):163–193, 2009.
[Kop12]
Swastik Kopparty. List-decoding multiplicity codes. Electronic Colloquium on
Computational Complexity (ECCC), 19:44, 2012.
[KR06]
Tali Kaufman and Dana Ron. Testing polynomials over general fields. SIAM
Journal of Computing, 36(3):779–802, 2006.
[KS07]
Tali Kaufman and Madhu Sudan. Algebraic property testing: The role of invariance. Electronic Colloquium on Computational Complexity (ECCC), 14(111),
2007.
[KS08]
Tali Kaufman and Madhu Sudan. Algebraic property testing: the role of invariance. In STOC, pages 403–412, 2008.
[KSY14]
Swastik Kopparty, Shubhangi Saraf, and Sergey Yekhanin. High-rate codes
with sublinear-time decoding. J. ACM, 61(5):28, 2014.
[MR06]
Dana Moshkovitz and Ran Raz. Sub-constant error low degree test of almostlinear size. In Proceedings of the 38th Annual ACM Symposium on Theory of
Computing, Seattle, WA, USA, May 21-23, 2006, pages 21–30, 2006.
[PW04]
Ruud Pellikaan and Xin-Wen Wu. List decoding of q-ary reed-muller codes.
IEEE Transactions on Information Theory, 50(4):679–682, 2004.
[RS96]
Ronitt Rubinfeld and Madhu Sudan. Robust characterizations of polynomials
with applications to program testing. SIAM Journal on Computing, 25(2):252–
271, April 1996.
[RS97]
Ran Raz and Shmuel Safra. A sub-constant error-probability low-degree test,
and a sub-constant error-probability PCP characterization of NP. In Proceedings of the Twenty-Ninth Annual ACM Symposium on Theory of Computing,
pages 475–484, New York, NY, 1997. ACM Press.
[SS08]
Shubhangi Saraf and Madhu Sudan. Improved lower bound on the size of
Kakeya sets over finite fields. ArXiv e-prints, August 2008.
120
[STV99]
Madhu Sudan, Luca Trevisan, and Salil Vadhan. Pseudorandom generators
without the XOR lemma. Proceedings of the 31st Annual ACM Symposium on
Theory of Computing, pages 537–546, 1999.
[Val05]
Paul Valiant. The tensor product of two codes is not necessarily robustly
testable. In Chandra Chekuri, Klaus Jansen, José D. P. Rolim, and Luca Trevisan, editors, APPROX-RANDOM, volume 3624 of Lecture Notes in Computer
Science, pages 472–481. Springer, 2005.
[Vid12]
Michael Viderman. A combination of testability and decodability by tensor
products. In Approximation, Randomization, and Combinatorial Optimization.
Algorithms and Techniques - 15th International Workshop, APPROX 2012, and
16th International Workshop, RANDOM 2012, Cambridge, MA, USA, August
15-17, 2012. Proceedings, pages 651–662, 2012.
[WB86]
Lloyd R. Welch and Elwyn R. Berlekamp. Error correction of algebraic block
codes. US Patent Number 4,633,470, December 1986.
121