1967 lines
170 KiB
TeX
1967 lines
170 KiB
TeX
In this chapter, we present the first dynamic group signature scheme based on lattice assumptions.
|
|
This construction relies on a signature scheme with efficient protocols as in~\cref{ch:sigmasig}, which is used in a similar manner.
|
|
As a consequence, it is possible to design lattice-based anonymous credentials from this building block.
|
|
The group signature scheme relies on the Gentry, Peikert and Vaikuntanathan identity-based encryption~\cite{GPV08} with the Canetti, Halevi and Katz~\cite{CHK04} transform to obtain a CCA2-secure public key encryption scheme which will be used to provide full-anonymity.
|
|
|
|
The group signature is proven secure in the \ROM under the \SIS and \LWE assumptions, which are fixed-size and well-studied assumptions.
|
|
As of the security parameter $\lambda$ and groups of up to $\Ngs$ members, the scheme features public key size $\softO(\lambda^2) \cdot \log \Ngs$, user's secret key size $\softO(\lambda)$ and signature size $\softO(\lambda) \cdot \log \Ngs$.
|
|
Our scheme thus achieves a level of efficiency comparable to recent proposals based on standard (i.e. non-ideal) lattices~\cite{LLLS13,NZZ15,LNW15,LLNW16} in the static setting as depicted in \cref{table:lattice-gs-comparison}.
|
|
In particular, the cost of moving to dynamic group is reasonable: while using the scheme from~\cite{LNW15} as a building block, our construction lengthens the signature size only by a (small) constant factor.
|
|
|
|
\begin{table}[t]
|
|
\scriptsize \centering
|
|
\begin{tabular}{|c||c|c|c|c|c|c|c|}
|
|
\hline
|
|
% after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
|
|
Scheme & \cite{LLLS13} & \cite{NZZ15} & \cite{LNW15} & \cite{LLNW16} & Ours \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
Group PK & $\widetilde{\mathcal{O}}(\lambda^2)\cdot \log N_\mathsf{gs}$ & $\widetilde{\mathcal{O}}(\lambda^2)$ & $\widetilde{\mathcal{O}}(\lambda^2)\cdot \log N_\mathsf{gs}$ & $\widetilde{\mathcal{O}}(\lambda^2)$& $\widetilde{\mathcal{O}}(\lambda^2)\cdot \log N_\mathsf{gs}$ \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
User's SK & $\widetilde{\mathcal{O}}(\lambda^2)$ & $\widetilde{\mathcal{O}}(\lambda^2)$ & $\widetilde{\mathcal{O}}(\lambda)$ &$\widetilde{\mathcal{O}}(\lambda)\cdot \log N_\mathsf{gs} $ & $\widetilde{\mathcal{O}}(\lambda)$ \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
Signature & $\widetilde{\mathcal{O}}(\lambda)\cdot \log N_\mathsf{gs}$ & $\widetilde{\mathcal{O}}(\lambda + \log^2 N_\mathsf{gs})$ & $\widetilde{\mathcal{O}}(\lambda)\cdot \log N_\mathsf{gs}$ & $\widetilde{\mathcal{O}}(\lambda)\cdot \log N_\mathsf{gs}$ & $ \widetilde{\mathcal{O}}(\lambda)\cdot \log N_\mathsf{gs}$ \\
|
|
\hline
|
|
\end{tabular}
|
|
\caption[Comparison between recent lattice-based group signatures]{Efficiency comparison among recent lattice-based group signatures for static groups and our dynamic scheme. The evaluation is done with respect to $2$ governing parameters: security parameter $\lambda$ and the maximum expected group size $N_\mathsf{gs}$. We do not include the earlier schemes~\cite{GKV10,CNR12} that have signature size $\widetilde{\mathcal{O}}(\lambda^2)\cdot N_\mathsf{gs}$.}
|
|
\label{table:lattice-gs-comparison}
|
|
\end{table}
|
|
|
|
The signature scheme with efficient protocols is built upon the $\SIS$-based signature of Böhl \textit{et al.}~\cite{BHJ+15}, which is itself a variant of Boyen's signature~\cite{Boy10}.
|
|
The latter scheme involves a public key containing matrices $\mathbf{A}, \mathbf{A}_0, \ldots, \mathbf{A}_\ell \in \Zq^{n \times m}$ and signs an $\ell$-bit message $\mathfrak m \in \bit^\ell$ by computing a short vector $\mathbf{v} \in \ZZ^{2m}_{}$ such that ${[\mathbf{A} \mid \mathbf{A}_0 + \sum_{j=1}^\ell \mathfrak m[j] \mathbf{A}_j ]} \cdot \mathbf{v} = \mathbf 0^n \bmod q$.
|
|
The variant proposed by Böhl \textit{et al.} only uses a constant number of matrices $\mathbf{A}, \mathbf{A}_0, \mathbf{A}_1 \in \Zq^{n \times m}$ where each signature is assigned with a single-use tag $\tau$ and the public key involves an extra matrix $\mathbf{D} \in \Zq^{n \times m}$ and a vector $\mathbf{u} \in \Zq^n$.
|
|
A message $\mathfrak m$ is then signed by first applying a chameleon hash function $\mathbf{h} = \mathcal{H}(\mathfrak m, \mathbf{s}) \in \bit^m_{}$ and signing $\mathbf{h}$ by computing a short $\mathbf{v} \in \ZZ^{2m}_{}$ such that ${[\mathbf{A} \mid \mathbf{A}_0 + \tau \mathbf{A}_1 ]} \cdot \mathbf{v} = \mathbf{u} + \mathbf{D} \cdot \mathbf{h} \bmod q$.
|
|
|
|
Our scheme extends~\cite{BHJ+15} so that an $N$-block message $(\mathfrak m_1, \ldots, \mathfrak m_N) \in (\bit^L)^N$, for some $L \in \NN$, is signed by outputting a tag $\tau \in \bit^\ell$ and a short $\mathbf{v} \in \ZZ^{2m}$ such that ${[\mathbf{A} \mid \mathbf{A}_0 + \sum_{j=1}^\ell \tau[j] \mathbf{A}_j ]} \cdot \mathbf{v} = \mathbf{u} + \mathbf{D} \cdot \mathcal{H}(\mathfrak{m}_1, \ldots, \mathfrak{m}_N, \mathbf s) \bmod q$, where the chameleon hash function computes $\mathbf{c}_M = \mathbf D_0 \cdot \mathbf s + \sum_{k=1}^{N} \mathbf D_k \cdot \mathfrak{m}_k \bmod q$, for some short vector $\mathbf s$, before re-encoding $\mathbf c_M$ so as to enable multiplication by $\mathbf D$.
|
|
|
|
In order to obtain a signature scheme that possesses efficient protocols akin to Camenish and Lysyanskaya~\cite{CL02}, our idea is to have the tag $\tau \in \bit^\ell$ play the same role as the prime exponent in Strong-RSA-based schemes~\cite{CL02a}.
|
|
To adapt this idea in the context of signatures with efficient protocols, we have to overcome several difficulties.
|
|
The first one is to map $\mathbf c_M$ back in the domain of the chameleon hash function while preserving the compatibility with ZK proofs.
|
|
To solve this issue, we extend a technique used in~\cite{LLNW16} in order to build a ``zero-knowledge-friendly'' chameleon hash function.
|
|
This function hashes the message by outputting the coordinate-wise binary decomposition $\mathbf w$ of $\mathbf D_0 \cdot \mathbf s + \sum_{k=1}^{N} \mathbf D_k \cdot \mathfrak{m}_k$. Using the ``power-of-two'' matrix $\mathbf H = \mathbf I \otimes [ 1 \mid 2 \mid \cdots \mid 2^{\lceil \log q\rceil} ]$, we can prove that $\mathbf w = \mathcal{H}(\mathfrak{m}_1, \ldots, \mathfrak{m}_N, \mathbf s)$ by demonstrating the knowledge of short vectors $(\mathfrak{m}_1, \ldots, \mathfrak{m}_N, \mathbf s, \mathbf w)$ that verifies $\mathbf H \cdot \mathbf w = \mathbf D_0 \cdot \mathbf s + \sum_{k=1}^{N} \mathbf D_k \cdot \mathfrak{m}_k \bmod q$ which can be proven using the ZKAoK of \cref{sse:stern}.
|
|
|
|
The second problem is to prove knowledge of $(\tau,\mathbf{v},\mathbf{s})$ and $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$ satisfying $[\mathbf{A} \mid \mathbf{A}_0 + \sum_{j=1}^\ell
|
|
\tau[j] \cdot \mathbf{A}_j ] \cdot \mathbf{v} = \mathbf{u} + \mathbf{D} \cdot \mathsf{CMHash}(\mathfrak{m}_1,\ldots,\mathfrak{m}_N,\mathbf{s})$, without revealing any of the witnesses. To
|
|
this end, we provide a framework for proving all the involved statement (and many other relations that naturally arise in lattice-based cryptography) as
|
|
special cases. We reduce the statements to asserting that a short integer vector $\mathbf{x}$ satisfies an equation of the form $\mathbf{P} \cdot \mathbf{x} = \mathbf{v}
|
|
\bmod q$, for some public matrix $\mathbf{P}$ and vector~$\mathbf{v}$, and belongs to a set $\mathsf{VALID}$ of short vectors with a particular structure. While the
|
|
small-norm property of $\mathbf{x}$ is provable using standard techniques (e.g., \cite{Lyu08}), we argue its membership of $\mathsf{VALID}$ by leveraging
|
|
the properties of Stern-like protocols \cite{Ste96,KTX08,LNSW13}. In particular, we rely on the fact that their underlying permutations interact well with
|
|
combinatorial statements pertaining to $\mathbf{x}$, especially $\mathbf{x}$ being a bitstring with a specific pattern. We believe our framework to be of independent
|
|
interest as it provides a blueprint for proving many other intricate relations in a modular manner.
|
|
|
|
When we extend the scheme with a protocol for signing committed messages, we need the signer to re-randomize the user's commitment before signing the hidden
|
|
messages. This is indeed necessary to provide the reduction with a backdoor allowing to correctly answer the $i^\dagger$-th query by ``programming'' the
|
|
randomness of the commitment. Since we work with integers vectors, a straightforward simulation incurs a non-negligible statistical distance between the
|
|
simulated distributions of re-randomization coins and the real one (which both have a discrete Gaussian distribution). Camenisch and Lysyanskaya \cite{CL02}
|
|
address a similar problem by choosing the signer's randomness to be exponentially larger than that of the user's commitment so as to statistically ``drown''
|
|
the aforementioned discrepancy. Here, the same idea would require to work with an exponentially large modulus~$q$. Instead, we adopt a more efficient
|
|
solution, inspired by Bai \textit{et al.} \cite{BLL+15}, which is to apply an analysis based on the R\'enyi divergence rather than the statistical distance. In
|
|
short, the R\'enyi divergence's properties tell us that, if some event~$E$ occurs with noticeable probability in some probability space~$P$, so does it in a
|
|
different probability space~$Q$ for which the second order divergence $R_2(P||Q)$ is sufficiently small. In our setting, $R_2(P||Q)$ is precisely polynomially
|
|
bounded since the two probability spaces only diverge in one signing query.
|
|
|
|
Our dynamic group signature scheme avoids these difficulties because the group manager only signs known messages: instead of signing the user's secret key as
|
|
in anonymous credentials, it creates a membership certificate by signing the user's public key. Our zero-knowledge arguments accommodate the requirements of
|
|
the scheme in the following way. In the joining protocol that dynamically introduces new group members, the user $i$ chooses a membership secret consisting of
|
|
a short discrete Gaussian vector $\mathbf{z}_i $. This user generates a public syndrome $\mathbf{v}_i = \mathbf{F} \cdot \mathbf{z}_i \mod q$, for some public matrix
|
|
$\mathbf{F}$, which constitutes his public key. In order to certify $\mathbf{v}_i$, the group manager computes the coordinate-wise binary expansion
|
|
$\mathsf{bin}(\mathbf{v}_i) $ of $\mathbf{v}_i$. The vector $\mathsf{bin}(\mathbf{v}_i) $ is then signed using our signature scheme. Using the resulting signature
|
|
$(\tau,\mathbf{v},\mathbf{s}) $ as a membership certificate, the group member is able to sign a message by proving that: (i) He holds a valid signature
|
|
$(\tau,\mathbf{v},\mathbf{s})$ on some secret binary message $\mathsf{bin}(\mathbf{v}_i) $; (ii) The latter vector $\mathsf{bin}(\mathbf{v}_i) $ is the binary expansion of
|
|
some syndrome $\mathbf{v}_i$ of which he knows a GPV pre-image $\mathbf{z}_i $. We remark that condition (ii) can be proved by providing evidence that we have $
|
|
\mathbf{v}_i = \mathbf{H} \cdot \textsf{bin}(\mathbf{v}_i) = \mathbf{F} \cdot \mathbf{z}_i \bmod q$, for some short integer vector $\mathbf{z}_i $ and some binary $\mathsf{bin}(\mathbf{v}_i) $,
|
|
where $\mathbf{H}$ is the ``powers-of-$2$'' matrix. Our abstraction of Stern-like protocols \cite{Ste96,KTX08,LNSW13} allows us to efficiently argue such
|
|
statements. The fact that the underlying chameleon hash function smoothly interacts with Stern-like zero-knowledge arguments is the property that maintains
|
|
the user's capability of efficiently proving knowledge of the underlying secret key.
|
|
|
|
|
|
Given the state of $\NIZK$ proofs in the lattice setting, it seems hard to provide group signature schemes in the standard model.
|
|
|
|
|
|
In the forthcoming sections, we first provide the description of our signature with efficient protocols; then a description of our dynamic group signature will be given and finally, we will explain how to use the Stern abstraction of \cref{sse:stern} to provide the required zero-knowledge arguments.
|
|
|
|
\section{A Lattice-Based Signature with Efficient Protocols} \label{se:gs-lwe-sigep}
|
|
|
|
Our scheme can be seen as a variant of the B\"ohl \textit{et al.} signature \cite{BHJ+15}, where
|
|
each signature is a triple $(\tau,\mathbf{v},\mathbf{s})$, made of a tag $\tau \in \{0,1\}^\ell$ and integer vectors $(\mathbf{v},\mathbf{s})$ satisfying
|
|
$[\mathbf{A} \mid \mathbf{A}_0 + \sum_{j=1}^\ell \tau[j] \cdot \mathbf{A}_j ] \cdot \mathbf{v} = \mathbf{u} + \mathbf{D} \cdot \mathbf{h} \bmod q$,
|
|
where matrices $\mathbf{A}, \mathbf{A}_0, \ldots, \mathbf{A}_\ell, \mathbf{D} \in \Zq^{n \times m}$
|
|
are public random matrices and $\mathbf{h} \in \{0,1\}^m$ is a chameleon hash of the message which is computed using randomness $\mathbf{s}$.
|
|
A difference is that, while \cite{BHJ+15} uses a short single-use tag $\tau \in \Zq$,
|
|
we need the tag to be an $\ell$-bit string $\tau \in \{0,1\}^{\ell}$ which will assume the same role as the prime exponent of Camenisch-Lysyanskaya signatures
|
|
\cite{CL02a} in the security proof.
|
|
|
|
We show that a suitable chameleon hash function makes the scheme compatible with Stern-like zero-knowledge arguments \cite{LNSW13,LNW15} for arguing possession of a valid message-signature pair. \cref{sse:stern} shows how to translate such a statement into asserting that a short witness vector $\mathbf{x}$ with a particular structure satisfies
|
|
a relation of the form
|
|
$\mathbf{P} \cdot \mathbf{x} = \mathbf{v} \bmod q$, for some public matrix $\mathbf{P}$ and vector~$\mathbf{v}$.
|
|
The underlying chameleon hash can be seen as a composition of the chameleon hash of \cite[Se. 4.1]{CHKP10} with
|
|
a technique used in \cite{PSTY13,LLNW16}: on input of a message $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$, it outputs the binary decomposition of
|
|
$\mathbf{D}_0 \cdot \mathbf{s} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k$, for some discrete Gaussian vector $\mathbf{s}$.
|
|
|
|
\subsection{Description} \label{desc-sig-protoc}
|
|
|
|
We assume that messages are vectors of $N$ blocks $\mathsf{Msg}=(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$, where each
|
|
block is a $2m$-bit string $\mathfrak{m}_k = \mathfrak{m}_k[1] \ldots \mathfrak{m}_k[2m] \in \{0,1\}^{2m}$ for $k \in \{1,\ldots, N\}$.
|
|
|
|
For each vector $\mathbf{v} \in \Zq^L$, we denote by $\textsf{bin}(\mathbf{v}) \in \{0,1\}^{L \lceil \log q \rceil}$ the vector obtained by replacing each
|
|
coordinate of $\mathbf{v}$ by its binary representation.
|
|
|
|
|
|
|
|
\begin{description}
|
|
\item[\textsf{Keygen}$(1^\lambda,1^N)$:] Given a security parameter $\lambda>0$ and the number of blocks $N = \mathsf{poly}(\lambda)$, choose the following parameters: $n = \bigO(\lambda)$; a prime modulus $q = \widetilde{\bigO}(N\cdot n^{4})$; dimension $m =2n \lceil \log q \rceil $; an integer $\ell = \Theta(\lambda)$; and Gaussian parameters $\sigma = \Omega(\sqrt{n\log q}\log n)$, $\sigma_0 = 2\sqrt{2}(N+1) \sigma m^{3/2}$, and $\sigma_1 = \sqrt{\sigma_0^2 + \sigma^2}$. Define the message space as $(\{0,1\}^{2m})^N$.
|
|
\smallskip
|
|
\begin{itemize}
|
|
\item[1.] Run $\TrapGen(1^n,1^m,q)$ to get~$\mathbf{A} \in
|
|
\Zq^{n \times m}$ and a short basis $\mathbf{T}_{\mathbf{A}}$ of
|
|
$\Lambda_q^{\perp}(\mathbf{A}).$ This basis allows computing short vectors in $\Lambda_q^{\perp}(\mathbf{A})$ with a Gaussian parameter $\sigma$.
|
|
Next, choose $\ell+1$ random $\mathbf{A}_0,\mathbf{A}_1,\ldots,\mathbf{A}_{\ell} \sample \U(\Zq^{n \times m})$. %, where $\ell = \Theta(\lambda)$.
|
|
\item[2.] Choose random matrices $\mathbf{D} \sample \U(\Zq^{n \times m})$, $\mathbf{D}_0,\mathbf{D}_1,\ldots,\mathbf{D}_{N} \sample \U(\Zq^{2n \times 2m})$ as well as a random vector
|
|
$\mathbf{u} \sample \U(\Zq^n)$. \smallskip
|
|
\end{itemize}
|
|
The private key consists of $SK \coloneqq \mathbf{T}_{\mathbf{A}} \in \ZZ^{m \times m}$ and the public key is
|
|
$${PK}\coloneqq \big( \mathbf{A}, ~ \{\mathbf{A}_j \}_{j=0}^{\ell}, ~ \{\mathbf{D}_k\}_{k=0}^{N},~\mathbf{D}, ~\mathbf{u} \big).$$
|
|
% \smallskip
|
|
\item[\textsf{Sign}$\big(SK, \mathsf{Msg} \big)$:] To sign an $N$-block message
|
|
$\mathsf{Msg}=\left(\mathfrak{m}_1,\ldots,\mathfrak{m}_N \right) \in \left(\{0,1\}^{2m} \right)^N$,
|
|
\begin{enumerate}[1.]
|
|
\item Choose a random string $\tau \sample \U(\{0,1\}^\ell )$. Then, using $SK\coloneqq
|
|
\mathbf{T}_{\mathbf{A}}$, compute with $\ExtBasis$ a short delegated basis $\mathbf{T}_\tau \in \ZZ^{2m \times 2m}$
|
|
for the matrix
|
|
\begin{eqnarray} \label{tau-matrix}
|
|
\mathbf{A}_{\tau}=
|
|
[ \mathbf{A} \mid \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \tau[j] \mathbf{A}_j
|
|
] \in \Zq^{ n \times 2m}.
|
|
\end{eqnarray}
|
|
\item Sample a vector $\mathbf{s} \sample D_{\ZZ^{2m},\sigma_1 }$. Compute $\mathbf{c}_M \in \Zq^{2n}$ as a chameleon hash of $\left(\mathfrak{m}_1,\ldots,\mathfrak{m}_N \right)$: i.e., compute
|
|
$\mathbf{c}_M = \mathbf{D}_{0} \cdot \mathbf{s} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k \in \mathbb{Z}_q^{2n} ,$
|
|
which is used to define $\mathbf{u}_M=\mathbf{u} + \mathbf{D} \cdot \textsf{bin}( \mathbf{c}_M) \in \Zq^n .$
|
|
Then,
|
|
using the delegated basis $\mathbf{T}_\tau \in \ZZ^{2m \times 2m}$, sample a short vector $\mathbf{v} \in \ZZ^{2m}$ in $D_{\Lambda_q^{\mathbf{u}_M}(\mathbf{A}_\tau), \sigma}$.
|
|
\end{enumerate}
|
|
Output the signature $sig=(\tau,\mathbf{v},\mathbf{s}) \in \{0,1\}^\ell \times \ZZ^{2m} \times \ZZ^{2m}$. \smallskip
|
|
\item[\textsf{Verify}$\big(PK,\mathsf{Msg},sig\big)$:] Given $PK$, a message $\mathsf{Msg}=(\mathfrak{m}_1,\ldots,\mathfrak{m}_N) \in (\{0,1\}^{2m})^N$ and a purported
|
|
signature $sig=(\tau,\mathbf{v},\mathbf{s}) \in \{0,1\}^\ell \times \ZZ^{2m} \times \ZZ^{2m}$,
|
|
return $1$ if
|
|
\begin{eqnarray} \label{ver-eq-block}
|
|
\mathbf{A}_{\tau} \cdot \mathbf{v} &=& \mathbf{u} + \mathbf{D} \cdot \textsf{bin}( \mathbf{D}_0 \cdot \mathbf{s} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k ) \bmod q.
|
|
\end{eqnarray}
|
|
and $\| \mathbf{v} \| < \sigma \sqrt{2m}$, $\| \mathbf{s} \| < \sigma_1 \sqrt{2m}$.
|
|
\end{description}
|
|
When the scheme is used for obliviously signing committed messages,
|
|
the security proof follows Bai \textit{et al.} \cite{BLL+15} in that it applies an argument based on the R\'enyi divergence in one signing query. This argument requires
|
|
to sample $\mathbf{s}$ from a Gaussian distribution whose standard deviation $\sigma_1$ is polynomially larger than $\sigma$.
|
|
|
|
We note that, instead of being included in the public key, the matrices $ \{\mathbf{D}_k\}_{k=0}^{N}$ can be part of common public parameters shared by many signers. Indeed,
|
|
only the matrices $(\mathbf{A},\{\mathbf{A}_i\}_{i=0}^\ell)$ should be specific to the user who holds the secret key $SK=\mathbf{T}_{\mathbf{A}}$. In Section \ref{commit-sig}, we use a variant where $ \{\mathbf{D}_k\}_{k=0}^{N}$
|
|
belong to public parameters.
|
|
|
|
|
|
\subsection{Security Analysis}
|
|
The security analysis in Theorem \ref{th:gs-lwe-security-cma-sig} requires that $q>\ell$.
|
|
|
|
|
|
\begin{theorem} \label{th:gs-lwe-security-cma-sig}
|
|
The signature scheme is secure under chosen-message attacks under the $\SIS$ assumption.
|
|
\end{theorem}
|
|
|
|
\begin{proof}
|
|
To prove the result, we will distinguish three kinds of attacks:
|
|
\begin{description}
|
|
\item[Type I attacks] are attacks where, in the adversary's forgery $sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$, $\tau^\star$ did not appear in any output
|
|
of the signing oracle.
|
|
\item[Type II attacks] are such that, in the adversary's forgery $sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$, $\tau^\star$ is recycled from an output
|
|
$sig^{(i^\star)}=(\tau^{(i^\star)},\mathbf{v}^{(i^\star)},\mathbf{s}^{(i^\star)})$ of the signing oracle, for some index $i^\star \in \{1,\ldots,Q\}$. However,
|
|
if $\mathsf{Msg}^\star=(\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star)$ and $\mathsf{Msg}^{(i^\star)}=(\mathfrak{m}_1^{(i^\star)},\ldots,\mathfrak{m}_N^{(i^\star)})$ denote the forgery
|
|
message and the $i^\star$-th signing query, respectively, we have
|
|
$\mathbf{D}_0 \cdot \mathbf{s}^\star + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^\star \neq \mathbf{D}_0 \cdot \mathbf{s}^{(i^\star)} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^{(i^\star)}. $
|
|
\item[Type III attacks] are those where the adversary's forgery $sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$ recycles $\tau^\star $ from an output
|
|
$sig^{(i^\star)}=(\tau^{(i^\star)},\mathbf{v}^{(i^\star)},\mathbf{s}^{(i^\star)})$ of the signing oracle (i.e.,
|
|
$\tau^{(i^\star)}= \tau^\star$ for some index $i^\star \in \{1,\ldots,Q\}$) and we have the collision
|
|
\begin{eqnarray} \label{collision}
|
|
\mathbf{D}_0 \cdot \mathbf{s}^\star + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^\star = \mathbf{D}_0 \cdot \mathbf{s}^{(i^\star)} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^{(i^\star)}.
|
|
\end{eqnarray}
|
|
\end{description}
|
|
Type III attacks imply a collision for the chameleon hash function of Kawachi \textit{et al.} \cite{KTX08}: if (\ref{collision}) holds,
|
|
a short vector
|
|
of $\Lambda_q^{\perp}([ \mathbf{D}_0 \mid \mathbf{D}_1 \mid \ldots \mid \mathbf{D}_N])$ is obtained as
|
|
$$\big({\mathbf{s}^\star}^T- {\mathbf{s}^{(i^\star)}}^T \mid {\mathfrak{m}_1^\star }^T - {\mathfrak{m}_1^{(i^\star)} }^T \mid \ldots \mid {\mathfrak{m}_N^\star }^T - {\mathfrak{m}_N^{(i^\star)} }^T \big)^T,$$ so that a collision breaks the $\mathsf{SIS}$ assumption.
|
|
|
|
The security against Type I attacks is proved by \cref{le:lwe-gs-type-I-attacks} which applies the same technique as in \cite{Boy10,MP12}. In particular, the prefix guessing technique
|
|
of \cite{HW09} allows keeping the modulus smaller than the number $Q$ of adversarial queries as in \cite{MP12}.
|
|
In order to deal with Type II attacks, we can leverage the technique of~\cite{BHJ+15}. In \cref{le:lwe-gs-type-II-attacks}, we prove that Type II attack would also contradict $\mathsf{SIS}$.
|
|
\end{proof}
|
|
|
|
\begin{lemma} \label{le:lwe-gs-type-I-attacks}
|
|
The scheme is secure against Type I attacks if the $\mathsf{SIS}_{n,m,q,\beta'}$ assumption holds for $\beta' = m^{3/2} \sigma^2 ( \ell+3) + m^{1/2} \sigma_1 $
|
|
\end{lemma}
|
|
|
|
\begin{proof}
|
|
Let $\adv$ be a $\ppt$ adversary that can mount a Type I attack with non-negligible success probability $\varepsilon$. We construct a $\ppt$
|
|
algorithm $\bdv$ that uses $\adv$ to break the~$\SIS_{n,m,q,\beta'}$ assumption. It takes as input~$\bar{\mathbf{A}} \in
|
|
\Zq^{n \times m}$ and computes $\mathbf{v} \in
|
|
\Lambda_q^{\perp}(\bar{\mathbf{A}})$ with~$0 < \|\mathbf{v}\| \leq \beta'$.
|
|
|
|
Algorithm~$\bdv$ first chooses the $\ell$-bit strings $\tau^{(1)},\ldots,\tau^{(Q)} \sample \U(\{0,1\}^\ell)$ to be used in signing queries. As in \cite{HW09}, it
|
|
guesses the shortest prefix such that the string $\tau^\star$ contained in $\adv$'s forgery differs from all prefixes of $\tau^{(1)},\ldots,\tau^{(Q)}$. To this
|
|
end, $\bdv$ chooses $i^\dagger \sample \U(\{1,\ldots, Q\})$ and $t^\dagger \sample \U(\{1,\ldots,\ell\})$ so that, with probability $1/(Q \cdot \ell)$, the longest
|
|
common prefix between $\tau^\star$ and one of the $\left\{\tau^{(i)} \right\}_{i=1}^Q$ is the string
|
|
$\tau^\star[1] \ldots \tau^\star[t^\dagger-1] =\tau^{(i^\dagger)}[1] \ldots \tau^{(i^\dagger)}[t^\dagger-1] \in \{0,1\}^{t^\dagger -1}$ comprised of the
|
|
first $(t^\dagger-1)$-th bits of $\tau^\star \in \{0,1\}^\ell$. We define $ \tau^\dagger \in \{0,1\}^{t^\dagger}$ as the $t^\dagger$-bit string $\tau^\dagger=\tau^\star[1] \ldots \tau^\star[t^\dagger] $. By construction, with probability $1/(Q \cdot \ell)$, we have $\tau^\dagger \not \in \left\{\tau^{(1)}_{|{t^\dagger}}, \ldots ,\tau^{(Q)}_{|{t^\dagger}} \right\}$, where $\tau^{(i)}_{|{t^\dagger}}$ denotes
|
|
the $t^\dagger$-th prefix of $\tau^{(i)}$ for each $i\in \{1,\ldots,Q\}$.
|
|
|
|
Then, $\bdv$ runs
|
|
$\TrapGen(1^n,1^m,q)$ to obtain $\mathbf{C} \in \Zq^{n \times m}$ and a
|
|
basis $\mathbf{T}_{\mathbf{C}}$ of~$\Lambda_q^{\perp}(\mathbf{C})$ with
|
|
$\|\widetilde{\mathbf{T}_{\mathbf{C}}}\| \leq \bigO(\sqrt{n \log q})$. Then,
|
|
it picks~$\ell+1$ matrices~$\mathbf{Q}_0,\ldots, \mathbf{Q}_{\ell} \in \ZZ^{m \times m}$, where
|
|
each matrix $\mathbf{Q}_i$ has its columns sampled independently from~$D_{\ZZ^m, \sigma}$. The
|
|
reduction $\bdv$ defines the matrices $\{ \mathbf{A}_j\}_{j=0}^{\ell}$ as
|
|
\begin{eqnarray*}
|
|
\left\{
|
|
\begin{array}{ll}
|
|
\mathbf{A}_0 = \bar{\mathbf{A}} \cdot \mathbf{Q}_0 + (\sum_{j=1}^{t^\dagger} {\tau^\star[j]}) \cdot
|
|
\mathbf{C} \\
|
|
\mathbf{A}_j = \bar{\mathbf{A}} \cdot \mathbf{Q}_j + (-1)^{\tau^\star[j]} \cdot
|
|
\mathbf{C}, \qquad \quad \text{ for } j \in
|
|
[1,t^\dagger] \\
|
|
\mathbf{A}_j = \bar{\mathbf{A}} \cdot \mathbf{Q}_j , \qquad \quad \qquad \quad~~ \qquad \quad \text{ for } j \in
|
|
[t^\dagger+1,\ell]
|
|
\end{array}
|
|
\right.
|
|
\end{eqnarray*}
|
|
It also sets $\mathbf{A}=\bar{\mathbf{A}}$.
|
|
We note that we have
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\tau^{(i)}} &=& \left[ \begin{array}{c|c} \bar{\mathbf{A}} & \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i)}[j] \mathbf{A}_j
|
|
\end{array} \right] \\
|
|
& = & \left[
|
|
\begin{array}{c|c}
|
|
\bar{\mathbf{A}} ~ & ~ \bar{\mathbf{A}} \cdot (\mathbf{Q}_0 +
|
|
\sum_{j=1}^{\ell} \tau^{(i)}[j] \mathbf{Q}_j) + (
|
|
\sum_{j=1}^{t^\dagger} \tau^\star[j] +(-1)^{\tau^\star[j]} \tau^{(i)}[j])\cdot \mathbf{C}
|
|
\end{array} \right]
|
|
\\
|
|
&=&
|
|
\left[
|
|
\begin{array}{c|c}
|
|
\bar{\mathbf{A}} ~ & ~ \bar{\mathbf{A}} \cdot (\mathbf{Q}_0 +
|
|
\sum_{j=1}^{\ell} \tau^{(i)}[j] \mathbf{Q}_j) + h_{\tau^{(i)}} \cdot \mathbf{C}
|
|
\end{array} \right]
|
|
\end{eqnarray*}
|
|
where $ h_{\tau^{(i)}} \in [1,t^\dagger] \subset [1,\ell]$ stands for the Hamming distance between
|
|
$\tau^{(i)}_{|t^\dagger}$ and $\tau^\star_{|t^\dagger}$. Note that, with probability $1/(Q \cdot \ell)$ and since $q>\ell$, we have
|
|
$ h_{\tau^{(i)}} \neq 0 \bmod q$ whenever $\tau^{(i)}_{|t^\dagger} \neq \tau^\star_{|t^\dagger}$.
|
|
|
|
Next, $\bdv$ chooses the matrices $\mathbf{D}_k \sample \U(\Zq^{2n \times 2m})$ uniformly at random for each $k \in [0,N]$. Then, it picks a random short matrix $\mathbf{R} \in \ZZ^{m \times m}$ which has its columns independently sampled from $D_{\ZZ^m,\sigma}$
|
|
and computes
|
|
\begin{eqnarray*}
|
|
\mathbf{D} &=& \bar{\mathbf{A}} \cdot \mathbf{R}. % \qquad \qquad \qquad \quad~ \forall k \in \{0,\ldots,N\}.
|
|
\end{eqnarray*}
|
|
Finally, $\bdv$ samples a short vector $\mathbf{e}_u \sample D_{\ZZ^m,\sigma_1}$ and computes the vector $\mathbf{u} \in \Zq^n$
|
|
as $\mathbf{u} = \bar{\mathbf{A}} \cdot \mathbf{e}_u \in \Zq^n$. The public key $${PK}\coloneqq \big( \mathbf{A}, ~
|
|
\{\mathbf{A}_j \}_{j=0}^{\ell}, ~ \{\mathbf{D}_k\}_{k=0}^{N},~\mathbf{D}, ~\mathbf{u} \big)$$
|
|
is given to $\adv$.
|
|
|
|
At the $i$-th signing query $\mathsf{Msg}^{(i)}=(\mathfrak{m}_1^{(i)},\ldots,\mathfrak{m}_N^{(i)}) \in (\{0,1\}^{2m})^N$, $\bdv$ can use the trapdoor $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$ to generate a signature.
|
|
To do this, $\bdv$ first samples $\mathbf{s}^{(i)} \sample D_{\ZZ^{2m},\sigma_1}$ and computes a vector $\mathbf{u}_M \in \Zq^m$ as
|
|
$$\mathbf{u}_M = \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i)} } + \mathbf{D}_{0} \cdot {\mathbf{s}^{(i)} } \bigr) ~~ \bmod q.$$
|
|
Using $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$, $\bdv$ can then sample a short vector $\mathbf{v}^{(i)} \in \ZZ^{2m}$ in $D^{\mathbf{u}_M}_{\Lambda^{\perp}(\mathbf{A}_{\tau^{(i)}}), \sigma}$ such
|
|
that $\big(\tau^{(i)},\mathbf{v}^{(i)},\mathbf{s}^{(i)} \big)$ satisfies the verification equation (\ref{ver-eq-block}).
|
|
|
|
When $\adv$ halts, it outputs a valid signature $sig^\star=\big(\tau^{(i^\dagger)},\mathbf{v}^\star,\mathbf{s}^\star \big)$ on a
|
|
message $\mathsf{Msg}^\star=(\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star)$ with $\| \mathbf{v}^\star \| \leq \sigma \sqrt{2m}$ and $\| \mathbf{s}^\star \| \leq \sigma_1 \sqrt{2m}$.
|
|
At this point, $\bdv$ aborts and declares failure if it was unfortunate in its choice of $i^\dagger \in \{1,\ldots,Q\}$ and $t^\dagger \in \{1,\ldots,\ell\}$. Otherwise,
|
|
with probability $1/(Q \cdot \ell)$, $\bdv$ correctly guessed $i^\dagger \in \{1,\ldots,Q\}$ and $t^\dagger \in \{1,\ldots,\ell\}$, in which case it can solve the given $\mathsf{SIS}$ instance as follows.
|
|
|
|
If we parse $\mathbf{v}^\star \in \ZZ^{2m}$ as $({\mathbf{v}_1^\star }^T \mid {\mathbf{v}_2^\star }^T )^T$ with $\mathbf{v}_1^\star,\mathbf{v}_2^\star \in \ZZ^m$, we have the equality
|
|
\begin{align*}
|
|
&\left[ \begin{array}{c|c} \bar{\mathbf{A}} ~&~ \bar{\mathbf{A}} \cdot (\mathbf{Q}_0 +
|
|
\sum_{j=1}^{\ell} \tau^\star[j] \mathbf{Q}_j)
|
|
\end{array} \right] \cdot
|
|
\left[\begin{array}{c} {\mathbf{v}_1^\star } \\ \hline {\mathbf{v}_2^\star } \end{array} \right] \\
|
|
& \hspace{3cm}= \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} } +
|
|
\sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \bigr) \bmod q \\
|
|
& \hspace{3cm}= \bar{\mathbf{A}} \cdot \Bigl( \mathbf{e}_u + \mathbf{R} \cdot \bit \bigl( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} } +
|
|
\sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \bigr) \Bigr) \bmod q ,
|
|
\end{align*}
|
|
which implies that the vector
|
|
\begin{eqnarray*}
|
|
\mathbf{w} &=& {\mathbf{v}_1^\star } + (\mathbf{Q}_0 +
|
|
\sum_{j=1}^{\ell} \tau^\star[j] \mathbf{Q}_j) \cdot {\mathbf{v}_2^\star } - \mathbf{e}_u - \mathbf{R} \cdot \bit \bigl( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} } +
|
|
\sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \bigr) \in \ZZ^m
|
|
\end{eqnarray*}
|
|
is in $\Lambda_q^{\perp}(\bar{\mathbf{A}})$. Moreover, with overwhelming probability, this vector is non-zero since, in $\adv$'s view, the distribution of
|
|
$\mathbf{e}_u \in \ZZ^m$ is $D_{\Lambda_q^{\mathbf{u}}(\bar{\mathbf{A}}),\sigma_1}$, which ensures that $\mathbf{e}_u$ is statistically hidden by
|
|
the syndrome $\mathbf{u} = \bar{\mathbf{A}} \cdot \mathbf{e}_u $. Finally, the norm of $\mathbf{w}$ is smaller than
|
|
$\beta' = m^{3/2} \sigma^2 ( \ell+3) + m^{1/2} \sigma_1 $
|
|
which yields a valid solution of the given $\mathsf{SIS}_{n,m,q,\beta'}$ instance
|
|
with overwhelming probability.
|
|
\end{proof}
|
|
|
|
\begin{lemma} \label{le:lwe-gs-type-II-attacks}
|
|
The scheme is secure against Type II attacks if the $\mathsf{SIS}_{n,m,q,\beta''}$ assumption holds for $\beta'' = \sqrt{2} (\ell+2) \sigma^2 m^{3/2} + m^{1/2} $.
|
|
\end{lemma}
|
|
|
|
\begin{proof}
|
|
We prove the result using a sequence of games. For each $i$, we denote by $W_i$ the event that the adversary wins by outputting a Type II forgery in \textsf{Game} $i$.
|
|
\medskip
|
|
|
|
\begin{description}
|
|
\item[\textsf{Game} 0:] This is the real game where, at the $i$-th signing query $\mathsf{Msg}^{(i)}=(\mathfrak{m}_1^{(i)},\ldots,\mathfrak{m}_N^{(i)})$,
|
|
the adversary obtains a signature $sig^{(i)}=(\tau^{(i)},\mathbf{v}^{(i)},\mathbf{s}^{(i)})$ for each $i \in \{1,\ldots,Q\}$ from the signing oracle. At the end of the game, the adversary
|
|
outputs a forgery $sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$ on a message $\mathsf{Msg}^{\star}=(\mathfrak{m}_1^{\star},\ldots,\mathfrak{m}_N^{\star})$.
|
|
By hypothesis, the adversary's advantage is $\varepsilon = \Pr[W_0]$. We assume without loss of generality that the random $\ell$-bit strings $\tau^{(1)}, \ldots, \tau^{(Q)}$ are chosen
|
|
at the very beginning of the game.
|
|
Since $(\mathsf{Msg}^\star,sig^\star)$ is a Type II forgery, there exists an index $i^\star \in \{1,\ldots,Q\}$ such that $\tau^\star =\tau^{(i^\star)} $.
|
|
|
|
\item[\textsf{Game} 1:] This game is identical to \textsf{Game} $0$ with the difference that the reduction aborts the experiment in the unlikely event that, in the adversary's forgery
|
|
$sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$, $\tau^\star$ coincides with more than one of the random $\ell$-bit strings $\tau^{(1)}, \ldots, \tau^{(Q)}$
|
|
used by the challenger. If we call $F_1$ the latter event, we have $\Pr[F_1] < Q^2/2^\ell$ since we are guaranteed to have $\neg F_1$ as long as no two $\tau^{(i)}$, $\tau^{(i')}$ collide.
|
|
Given that \textsf{Game} $1$ is identical to \textsf{Game} $0$ until $F_1$ occurs, we have $|\Pr[W_1]-\Pr[W_0]| \leq \Pr[F_1] < Q^2/2^\ell$.
|
|
|
|
\item[\textsf{Game} 2:] This game is like \textsf{Game} $1$ with the following difference. At the outset of the game, the challenger $\bdv$ chooses a random index
|
|
$i^\dagger \sample (\{1,\ldots,Q\})$ as a guess that $\adv$'s forgery will recycle the $\ell$-bit string $\tau^{(i^\dagger)} \in \{0,1\}^\ell$ of the $i^\dagger$-th signing query.
|
|
When $\adv$ outputs its Type II forgery $sig^\star=(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star)$, the challenger aborts
|
|
in the event that $\tau^{(i^\dagger)} \neq \tau^\star$ (i.e., $i^\dagger \neq i^\star$). Since the choice of $i^\dagger $ in $\{1,\ldots,Q\}$ is independent of $\adv$'s view, we
|
|
have $\Pr[W_2]=\Pr[W_1]/Q$.
|
|
|
|
\item[\textsf{Game} 3:] In this game, we modify the key generation phase and the way to answer signing queries.
|
|
First, the challenger $\bdv$ randomly picks $h_0,h_1,\ldots,h_\ell \in \Zq$ subject to the constraints
|
|
\begin{eqnarray*}
|
|
h_0 + \sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot h_j &=& 0 \bmod q \\
|
|
h_0 + \sum_{j=1}^\ell \tau^{(i)}[j] \cdot h_j & \neq & 0 \bmod q \qquad \qquad i \in \{1,\ldots,Q\} \setminus \{i^\dagger\}
|
|
\end{eqnarray*}
|
|
It runs $(\mathbf{C},\mathbf{T}_{\mathbf{C}}) \leftarrow \mathsf{TrapGen}(1^n,1^m,q)$,
|
|
$(\mathbf{D}_0,\mathbf{T}_{\mathbf{D}_0}) \leftarrow \mathsf{TrapGen}(1^{2n},1^{2m},q)$ so as to obtain statistically random matrices $\mathbf{C} \in \Zq^{n \times m} $, $\mathbf{D}_0 \in \Zq^{2n \times 2m}$ with
|
|
trapdoors $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$, $\mathbf{T}_{\mathbf{D}_0} \in \ZZ^{2m \times 2m}$ consisting of short bases of $\Lambda_q^{\perp}(\mathbf{C})$ and $\Lambda_q^{\perp}(\mathbf{D}_0)$, respectively. Then,
|
|
$\bdv$
|
|
chooses
|
|
a uniformly random $\mathbf{D} \sample (\Zq^{n \times m})$ and re-randomizes it using short matrices
|
|
$\mathbf{S},\mathbf{S}_0,\mathbf{S}_1,\ldots,\mathbf{S}_{\ell} \sample \ZZ^{m \times m}$, which are obtained
|
|
by sampling their columns from the distribution $D_{\ZZ^m,\sigma}$. Namely, from $\mathbf{D} \in \Zq^{n \times m}$, $\bdv$
|
|
defines
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A} &=& \mathbf{D} \cdot \mathbf{S} \\ \label{setup-sig3}
|
|
\mathbf{A}_0 &=& \mathbf{D} \cdot \mathbf{S}_0 + h_0 \cdot \mathbf{C} \\ \nonumber
|
|
\mathbf{A}_j &=& \mathbf{D} \cdot \mathbf{S}_j + h_j \cdot \mathbf{C} \qquad \qquad \forall j \in \{1,\ldots,\ell\} %\\ \nonumber
|
|
\end{eqnarray}
|
|
In addition, $\bdv$ picks random matrices $\mathbf{D}_1,\ldots,\mathbf{D}_N \sample (\Zq^{2n \times 2m})$ and a random vector $\mathbf{c}_M \sample (\Zq^{2n})$. It samples
|
|
short vectors $\mathbf{v}_1 ,\mathbf{v}_2 \sample D_{\ZZ^m,\sigma}$ and computes $\mathbf{u} \in \Zq^n$
|
|
as $\mathbf{u} = \mathbf{A}_{\tau^{(i^\dagger)}} \cdot
|
|
\left[
|
|
\begin{array}{c}
|
|
\mathbf{v}_1 \\ \hline \mathbf{v}_2
|
|
\end{array} \right]
|
|
- \mathbf{D} \cdot \textsf{bin}( \mathbf{c}_M ) \bmod q$, where
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\tau^{(i^\dagger)}} &=& \left[
|
|
\begin{array}{c|c} \mathbf{A} ~ & ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{A}_j
|
|
\end{array} \right] \\ &=& \left[
|
|
\begin{array}{c|c} \mathbf{D} \cdot \mathbf{S} ~ & ~ \mathbf{D}\cdot (\mathbf{S}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{S}_j)
|
|
\end{array} \right] .
|
|
\end{eqnarray*}
|
|
The adversary's signing queries are then answered as follows.
|
|
\begin{itemize}
|
|
\item At the $i$-th signing query $ (\mathfrak{m}_1^{(i)},\ldots,\mathfrak{m}_N^{(i)})$, whenever $i \neq i^\dagger$, we have
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\tau^{(i)}} &=& \left[
|
|
\begin{array}{c|c} \mathbf{A} ~& ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i)}[j] \cdot \mathbf{A}_j
|
|
\end{array} \right] \\
|
|
&=& \left[
|
|
\begin{array}{c|c} \mathbf{A} ~ & ~ \mathbf{D} \cdot (\mathbf{S}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i)}[j] \cdot \mathbf{S}_j) + h_{\tau^{(i)}} \cdot \mathbf{C}
|
|
\end{array} \right]
|
|
\in \Zq^{ n \times 2m},
|
|
\end{eqnarray*}
|
|
with $h_{\tau^{(i)}} = h_0 + \sum_{j=1}^\ell \tau^{(i)}[j] \cdot h_j \neq 0$. This implies that $\bdv$ can use the trapdoor $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$ to generate a signature.
|
|
To this end, $\bdv$ first samples a discrete Gaussian vector $\vec{s}^{(i)} \sample D_{\ZZ^{2m},\sigma_1}$ and computes $\mathbf{u}_M \in \Zq^n$ as
|
|
$$\mathbf{u}_M = \mathbf{u} + \mathbf{D} \cdot \textsf{bin}( \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i)} } + \mathbf{D}_{0} \cdot {\mathbf{s}^{(i)} } ) ~~ \bmod q.$$ Then,
|
|
using $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$, it samples a short vector $\mathbf{v}^{(i)} \in \ZZ^{2m}$ in $D^{\mathbf{u}_M}_{\Lambda^{\perp}(\mathbf{A}_{\tau^{(i)}}), \sigma}$ such
|
|
that $(\tau^{(i)},\mathbf{v}^{(i)},\mathbf{s}^{(i)})$ satisfies (\ref{ver-eq-block}).
|
|
\item At the $i^\dagger$-th signing query $ (\mathfrak{m}_1^{(i^\dagger)},\ldots,\mathfrak{m}_N^{(i^\dagger)})$, we have
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A}_{\tau^{(i^\dagger)}} &=& \left[
|
|
\begin{array}{c|c} \mathbf{A} ~& ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{A}_j
|
|
\end{array} \right] \\
|
|
\label{i-mat} &=& \left[
|
|
\begin{array}{c|c} \mathbf{D} \cdot \mathbf{S} ~&~ \mathbf{D} \cdot (\mathbf{S}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{S}_j)
|
|
\end{array} \right]
|
|
\in \Zq^{ n \times 2m} \quad
|
|
\end{eqnarray}
|
|
due to the constraint $h_0 + \sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot h_j = 0 \bmod q $.
|
|
To answer the query, $\bdv$ uses the trapdoor $\mathbf{T}_{\mathbf{D}_0} \in \ZZ^{2m \times 2m}$ of $\Lambda_q^{\perp}(\mathbf{D}_0)$ to sample a short vector
|
|
$\mathbf{s}^{(i^\dagger)} \in D_{\Lambda_q^{\mathbf{c}'_M} (\mathbf{D}_0), \sigma_1}$, where $\mathbf{c}'_M = \mathbf{c}_M - \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i^\dagger)} } \in \Zq^{2n}$.
|
|
The obtained vector $\mathbf{s}^{(i^\dagger)} \in \ZZ^{2m}$ thus verifies
|
|
|
|
\begin{eqnarray} \label{sim-s}
|
|
\mathbf{D}_0 \cdot {\mathbf{s}^{(i^\dagger)} } &=&
|
|
\mathbf{c}_M - \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i^\dagger)} } ~\bmod q,
|
|
\quad
|
|
\end{eqnarray}
|
|
and $\adv$ receives $sig^{(i^\dagger)}=(\tau^{(i^\dagger)},\mathbf{v}^{(i^\dagger)},\mathbf{s}^{(i^\dagger)})$, where $ \mathbf{v}^{(i^\dagger)} = (\mathbf{v}_1^T \mid \mathbf{v}_2^T)^T $.
|
|
By construction, the returned signature $sig^{(i^\dagger)}$ satisfies
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\tau^{(i^\dagger)}}
|
|
\cdot \left[ \begin{array}{c}
|
|
\mathbf{v}_1 \\ \hline \mathbf{v}_2 \end{array} \right]
|
|
&=& \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{D}_0 \cdot {\mathbf{s}^{(i^\dagger)} } + \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i^\dagger)} } \bigr) \quad \bmod q,
|
|
\end{eqnarray*}
|
|
and the distribution of $(\tau^{(i^\dagger)},\mathbf{v}^{(i^\dagger)},\mathbf{s}^{(i^\dagger)})$ is statistically the same as in \textsf{Game} $2$.
|
|
\end{itemize}
|
|
\end{description}
|
|
We conclude that $\Pr[W_2]$ is negligibly far apart from $\Pr[W_3]$ since, by the Leftover Hash Lemma (see \cite[Le. 13]{ABB10}), the public key $PK$ in \textsf{Game} $3$ is statistically close to its distribution in \textsf{Game} $2$.
|
|
\medskip
|
|
|
|
In \textsf{Game} $3$, we claim that the challenger $\bdv$ can use $\adv$ to solve the $\mathsf{SIS}$ problem by finding a short vector of $\Lambda_q^\perp(\mathbf{D})$ with probability $\Pr[W_3]$. Indeed,
|
|
with proba\-bility $\Pr[W_3]$, the adversary outputs a valid signature $sig^\star=(\tau^{(i^\dagger)},\mathbf{v}^\star,\mathbf{s}^\star)$ on a message $\mathsf{Msg}^\star=(\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star)$ with $\| \mathbf{v}^\star \| \leq \sigma \sqrt{2m}$ and $\| \mathbf{s}^\star \| \leq \sigma_1 \sqrt{2m}$.
|
|
If we parse $\mathbf{v}^\star \in \ZZ^{2m}$ as $({\mathbf{v}_1^\star }^T \mid {\mathbf{v}_2^\star }^T )^T$ with $\mathbf{v}_1^\star,\mathbf{v}_2^\star \in \ZZ^m$, we have
|
|
the equality
|
|
\begin{eqnarray} \label{first-sol}
|
|
\mathbf{A}_{\tau^{(i^\dagger)}} \cdot \left[ \begin{array}{c}
|
|
\mathbf{v}_1^\star \\ \hline \mathbf{v}_2^\star
|
|
\end{array} \right]
|
|
&=& \mathbf{u} + \mathbf{D} \cdot \textsf{bin}( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } ) \quad \bmod q.
|
|
\end{eqnarray}
|
|
|
|
Due to the way $\mathbf{u} \in \Zq^n$ was defined at the outset of the game, $\bdv$ also knows short vectors $\mathbf{v}^{(i^\dagger)}=(\mathbf{v}_1^T \mid \mathbf{v}_2^T)^T \in \ZZ^{2m}$
|
|
such that
|
|
\begin{eqnarray} \label{second-sol} \mathbf{A}_{\tau^{(i^\dagger)}} \cdot
|
|
\left[\begin{array}{c}
|
|
\mathbf{v}_1 \\ \hline \mathbf{v}_2
|
|
\end{array} \right] = \mathbf{u} + \mathbf{D} \cdot \textsf{bin}( \mathbf{c}_M ) \bmod q. \end{eqnarray}
|
|
Relation (\ref{sim-s}) implies that $ \mathbf{c}_M \neq \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \bmod q$ by hypothesis. It follows that $\textsf{bin}(\mathbf{c}_M) - \textsf{bin}( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } ) $ is a non-zero vector in $\{-1,0,1\}^m$. Subtracting (\ref{second-sol}) from (\ref{first-sol}), we get
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\tau^{(i^\dagger)}} \cdot \left[\begin{array}{c}
|
|
\mathbf{v}_1^\star - \mathbf{v}_1 \\ \hline \mathbf{v}_2^\star - \mathbf{v}_1
|
|
\end{array} \right]
|
|
&=& \mathbf{D} \cdot \bigl( \textsf{bin}(\mathbf{c}_M) - \textsf{bin}( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } ) \bigr) \mod q,
|
|
\end{eqnarray*}
|
|
which implies
|
|
\begin{multline} \label{eq-un}
|
|
\left[
|
|
\begin{array}{c|c} \mathbf{D} \cdot \mathbf{S} ~ &~ \mathbf{D} \cdot (\mathbf{S}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{S}_j)
|
|
\end{array} \right] \cdot \left[ \begin{array}{c} {\mathbf{v}_1^\star -\mathbf{v}_1 } \\ \hline {\mathbf{v}_2^\star - \mathbf{v}_2 } \end{array} \right] \\ = \mathbf{D} \cdot \bigl( \textsf{bin}(\mathbf{c}_M) - \textsf{bin}( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } ) \bigr) \mod q .
|
|
\end{multline}
|
|
The above implies that the vector
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{w} &=&
|
|
\mathbf{S} \cdot (\mathbf{v}_1^\star - \mathbf{v}_1) + (\mathbf{S}_0 + \sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{S}_j ) \cdot (\mathbf{v}_2^\star - \mathbf{v}_2) \\
|
|
\nonumber && \hspace{2.75cm} ~+ \bit \big( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} } + \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \big) - \textsf{bin}(\mathbf{c}_M)
|
|
\end{eqnarray}
|
|
is a short integer vector of $\Lambda_q^{\perp}(\mathbf{D})$. Indeed, its norm can be bounded as $\| \mathbf{w} \| \leq \beta'' = \sqrt{2} (\ell+2) \sigma^2 m^{3/2} + m^{1/2} $. We argue that it is non-zero with overwhelming probability. We already observed that
|
|
$ \textsf{bin}( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } ) - \textsf{bin}(\mathbf{c}_M)$ is a non-zero vector of $\{-1,0,1\}^m$, which rules out the event that
|
|
$({\mathbf{v}_1^\star}, {\mathbf{v}_2^\star} ) =({\mathbf{v}_1} , {\mathbf{v}_2}) $. Hence, we can only have $\mathbf{w}=\mathbf{0}^m$ when the equality
|
|
\begin{multline} \label{final-eq}
|
|
\mathbf{S} \cdot (\mathbf{v}_1^\star - \mathbf{v}_1) + (\mathbf{S}_0 +
|
|
\sum_{j=1}^\ell \tau^{(i^\dagger)}[j] \cdot \mathbf{S}_j ) \cdot (\mathbf{v}_2^\star - \mathbf{v}_2) \\ = \textsf{bin}(\mathbf{c}_M) - \bit \big( \mathbf{D}_0 \cdot {\mathbf{s}^{\star} }
|
|
+ \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{\star} } \big) \qquad
|
|
\end{multline}
|
|
holds over $\ZZ$. However, as long as either $\mathbf{v}_1^\star \neq \mathbf{v}_1$ or $\mathbf{v}_2^\star \ne \mathbf{v}_2$, the left-hand-side member of (\ref{final-eq})
|
|
is information theoretically unpredictable since the columns of matrices $\mathbf{S}$ and $\{\mathbf{S}_j\}_{j=0}^\ell$ are statistically hidden in the view of $\adv$.
|
|
Indeed, conditionally on the public key, each column of $\mathbf{S}$ and $\{\mathbf{S}_j\}_{j=0}^\ell$ has at least $n$ bits
|
|
of min-entropy, as shown by, e.g., \cite[Le. 2.7]{MP12}.
|
|
\end{proof}
|
|
|
|
|
|
\subsection{Protocols for Signing a Committed Value and Proving Possession of a Signature} \label{commit-sig}
|
|
|
|
|
|
|
|
We first show a two-party protocol whereby a user can interact with the signer in order to obtain a signature on a committed message.
|
|
|
|
In order to prove that the scheme still guarantees unforgeability for obliviously signed messages,
|
|
we will assume that each message block $\mathfrak{m}_k \in \{0,1\}^{2m}$ is obtained by encoding
|
|
the actual message $M_k =M_k[1] \ldots M_k[m] \in \{0,1\}^m$ as $\mathfrak{m}_k= \mathsf{Encode}(M_k)=( \bar{M}_k[1] , M_k[1],\ldots, \bar{M}_k[m] , M_k[m] ) $. Namely,
|
|
each $0$ (respectively each $1$) is encoded as a pair $(1,0)$ (resp. $(0,1)$). The reason for this encoding is that the proof of Theorem \ref{commit-thm} requires that at least one block
|
|
$\mathfrak{m}_k^\star $ of the forgery message is $1$ while the same bit is $0$ at some specific signing query. We will show (see \cref{se:gs-lwe-stern}) that the correctness of this encoding can
|
|
be efficiently proved using Stern-like~\cite{Ste96} protocols.
|
|
|
|
To sign committed messages, a first idea is exploit the fact that our signature of Section \ref{desc-sig-protoc} blends well with the $\mathsf{SIS}$-based commitment scheme suggested by Kawachi \textit{et al.}~\cite{KTX08}.
|
|
In the latter scheme, the commitment key consists of matrices $(\mathbf{D}_0,\mathbf{D}_1) \in \Zq^{2n \times 2m} \times \Zq^{2n \times 2m}$, so that message
|
|
$\mathfrak{m} \in \{0,1\}^{2m}$ can be committed to by sampling a Gaussian vector $\mathbf{s} \sample D_{\ZZ^{2m},\sigma}$ and computing
|
|
$\mathbf{C}= \mathbf{D}_0 \cdot \mathbf{s} + \mathbf{D}_1 \cdot \mathfrak{m} \in \Zq^{2n}$. This scheme extends to commit to multiple messages $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$ at once by computing
|
|
$\mathbf{C}=\mathbf{D}_0 \cdot \mathbf{s} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k \in \Zq^{2n}$ using a longer
|
|
commitment key $(\mathbf{D}_0,\mathbf{D}_1,\ldots,\mathbf{D}_N) \in (\Zq^{2n \times 2m})^{N+1} $. It is easy to see that the resulting commitment remains statistically hiding and computationally
|
|
binding under the $\mathsf{SIS}$ assumption.
|
|
|
|
In order to make our construction usable in the definitional framework of Camenisch \textit{et al.} \cite{CKL+15}, we assume common public parameters
|
|
(i.e., a common reference string) and encrypt all witnesses of which knowledge is being proved under a public key included in the common reference string. The resulting ciphertexts thus serve as statistically binding commitments
|
|
to the witnesses.
|
|
To enable this, the common public parameters comprise public keys $\mathbf{G}_0 \in \Zq^{n \times \ell}$, $\mathbf{G}_1 \in \Zq^{n \times 2m}$
|
|
for multi-bit variants of the dual Regev cryptosystem \cite{GPV08} and all parties are denied access to the underlying private keys. The flexibility of Stern-like protocols allows us to prove that the content of a perfectly hiding commitment $ \mathbf{c}_{\mathfrak{m}}$ is consistent with
|
|
encrypted values.%, the protocols of Ling \textit{et al.} \cite{LNW15} come in handy.
|
|
|
|
|
|
\begin{description}
|
|
\item[\textsf{Global}\textrm{-}\textsf{Setup}:] Let $B = \sqrt{n} \omega(\log n)$ and let $\chi$ be a $B$-bounded distribution.
|
|
Let $p = \sigma \cdot \omega(\sqrt{m})$ upper-bound entries of vectors sampled from the distribution~$D_{\ZZ^{2m},\sigma}$.
|
|
Generate two public keys for the dual Regev encryption scheme
|
|
in its multi-bit variant. These keys consists of a public random matrix
|
|
$\mathbf{B} \sample (\Zq^{n \times m})$ and random matrices $\mathbf{G}_0 = \mathbf{B} \cdot \mathbf{E}_0 \in \Zq^{n \times \ell }$, $\mathbf{G}_1 = \mathbf{B} \cdot \mathbf{E}_1 \in \Zq^{n \times 2m}$,
|
|
where $\mathbf{E}_0 \in \ZZ^{ m \times \ell}$ and $\mathbf{E}_1 \in \ZZ^{m \times 2m}$ are short Gaussian matrices with columns sampled from $D_{\ZZ^{m},\sigma}$. These matrices will be
|
|
used to encrypt integer vectors of dimension $\ell$ and $2m$, respectively. Finally, generate public parameters $CK\coloneqq \{ \mathbf{D}_k \}_{k=0}^N$ consisting of uniformly
|
|
random matrices $\mathbf{D}_k \sample (\Zq^{2n \times 2m})$ for a statistically hiding commitment
|
|
to vectors in $(\{0,1\}^{2m})^N$.
|
|
Return public parameters consisting of
|
|
\[ \mathsf{par}\coloneqq \{ \mathbf{B} \in \Zq^{n \times m} ,\mathbf{G}_0 \in \Zq^{n \times \ell},\mathbf{G}_1 \in \Zq^{n \times 2m},CK \}. \]
|
|
|
|
\item[\textsf{Issue} $\leftrightarrow$ \textsf{Obtain} :] The signer $S$, who holds a key pair $PK\coloneqq \{ \mathbf{A} , ~\{\mathbf{A}_j\}_{j=0}^\ell,~\mathbf{D},~\mathbf{u} \}$, $SK\coloneqq \mathbf{T}_{\mathbf{A}}$, interacts with the user $U$
|
|
who has a message $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$, in the following interactive protocol. \smallskip
|
|
\begin{itemize}
|
|
\item[1.] $U$ samples $\mathbf{s}' \sample D_{\ZZ^{2m},\sigma} $ and computes $ \mathbf{c}_{\mathfrak{m}} = \mathbf{D}_0 \cdot \mathbf{s}' + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k \in \mathbb{Z}_q^{2n}$
|
|
which is sent to $S$ as a commitment to $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$. In addition, $U$ encrypts $\{\mathfrak{m}_k\}_{k=1}^N$ and $\mathbf{s}'$ under the dual-Regev public key $(\mathbf{B},\mathbf{G}_1)$
|
|
by computing for all $k \in \{1,\ldots,N\}$:
|
|
\begin{align} \label{enc-Mk} \nonumber
|
|
\mathbf{c}_{k} & = (\mathbf{c}_{k,1},\mathbf{c}_{k,2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{k} + \mathbf{e}_{k,1} ,~ \mathbf{G}_1^T \cdot \mathbf{s}_{k} + \mathbf{e}_{k,2} + \mathfrak{m}_k \cdot \lfloor q/2 \rfloor \big) \in \Zq^m \times \Zq^{2m}
|
|
\end{align}
|
|
for randomly chosen $\mathbf{s}_{k} \sample \chi^n$, $\mathbf{e}_{k,1} \sample \chi^m$, $\mathbf{e}_{k,2} \sample \chi^{2m}$,
|
|
and
|
|
\begin{align} \label{enc-s} \nonumber
|
|
\mathbf{c}_{s'} & = (\mathbf{c}_{s',1},\mathbf{c}_{s',2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{0} + \mathbf{e}_{0,1} ,~ \mathbf{G}_1^T \cdot \mathbf{s}_{0} + \mathbf{e}_{0,2} + \mathbf{s}' \cdot \lfloor q/p \rfloor \big) \in \Zq^m \times \Zq^{2m}
|
|
\end{align}
|
|
where $\mathbf{s}_{0} \sample \chi^n$, $\mathbf{e}_{0,1} \sample \chi^m$, $\mathbf{e}_{0,2} \sample \chi^{2m}$. The ciphertexts $\{\mathbf{c}_k\}_{k=1}^N$ and $\mathbf{c}_{s'}$ are
|
|
sent to $S$ along with $\mathbf{c}_{\mathfrak{m}}$.
|
|
|
|
Then, $U$ generates an interactive zero-knowledge argument to convince~$S$ that
|
|
$ \mathbf{c}_{\mathfrak{m}}$ is a commitment to $(\mathfrak{m}_1, \ldots, \mathfrak{m}_N)$ with the randomness $\mathbf{s}'$ such that $\{\mathfrak{m}_k\}_{k=1}^N$ and
|
|
$\mathbf{s}'$ were honestly encrypted to $\{ \mathbf{c}_{k} \}_{i=1}^N$ and $\mathbf{c}_{s'}$, as in~(\ref{enc-Mk}) and~(\ref{enc-s}).
|
|
For convenience, this argument system will be described in Section~\ref{subsection:zk-for-commitments}, where we demonstrate that, together with other zero-knowledge protocols used in this work, it can be derived from a Stern-like~\cite{Ste96} protocol constructed in \cref{se:gs-lwe-stern}.
|
|
|
|
\item[2.] If the argument of step 1 properly verifies, $S$ samples $\mathbf{s}'' \sample D_{\ZZ^{2m},\sigma_0}$ and computes
|
|
a vector $\mathbf{u}_{\mathfrak{m}}= \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{c}_{\mathfrak{m}} + \mathbf{D}_0 \cdot \mathbf{s}'' \bigr) \in \Zq^n$.
|
|
Next, $S$ randomly picks $\tau \sample \{0,1\}^\ell$ and
|
|
uses $\mathbf{T}_{\mathbf{A}}$ to compute a delegated basis $\mathbf{T}_{\tau} \in \ZZ^{2m \times 2m}$ for the matrix $\mathbf{A}_{\tau} \in \Zq^{n \times 2m}$ of (\ref{tau-matrix}).
|
|
Using $\mathbf{T}_\tau \in \ZZ^{2m \times 2m}$, $S$ samples a short vector $\mathbf{v} \in \ZZ^{2m}$ in $D^{\mathbf{u}_M}_{\Lambda^{\perp}(\mathbf{A}_\tau), \sigma}$. It returns
|
|
the vector $( \tau,\mathbf{v},\mathbf{s}'') \in \{0,1\}^\ell \times \ZZ^{2m} \times \ZZ^{2m} $ to $U$.
|
|
\item[3.] $U$ computes $\mathbf{s} = \mathbf{s}'+\mathbf{s}''$ over $\ZZ$ and verifies that $$\mathbf{A}_{\tau} \cdot \mathbf{v} = \mathbf{u} + \mathbf{D} \cdot \bit
|
|
\bigl( \mathbf{D}_0 \cdot \mathbf{s} + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k \bigr) \bmod q.$$ If so, it outputs $(\tau,\mathbf{v},\mathbf{s})$. Otherwise, it outputs $\perp$.
|
|
\end{itemize}
|
|
\end{description}
|
|
Note that, if both parties faithfully run the protocol, the user obtains a valid signature $(\tau,\mathbf{v},\mathbf{s})$ for which the distribution of $\mathbf{s}$ is $D_{\ZZ^{2m},\sigma_1}$,
|
|
where $\sigma_1=\sqrt{\sigma^2 + \sigma_0^2}$.
|
|
|
|
The following protocol allows proving possession of a message-signature pair.
|
|
|
|
\begin{description}
|
|
\item[\textsf{Prove}:] On input of a signature $(\tau,\mathbf{v}=(\mathbf{v}_1^T \mid \mathbf{v}_2^T)^T,\mathbf{s}) \in \{0,1\}^\ell \times \ZZ^{2m} \times \ZZ^{2m}$ on the message $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$, the user
|
|
does the following. \smallskip \smallskip
|
|
\begin{itemize}
|
|
\item[1.] Using $(\mathbf{B},\mathbf{G}_0)$ and $(\mathbf{B},\mathbf{G}_1)$ generate perfectly binding commitments to $\tau \in \{0,1\}^\ell$, $\{\mathfrak{m}_k \}_{k=1}^N$,
|
|
$\mathbf{v}_1,\mathbf{v}_2 \in \ZZ^m$ and $\mathbf{s} \in \ZZ^{2m}$. Namely, compute
|
|
\begin{align*}
|
|
\mathbf{c}_{\tau} & = (\mathbf{c}_{\tau,1},\mathbf{c}_{\tau,2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{\tau} + \mathbf{e}_{\tau,1} ,~ \mathbf{G}_0^T \cdot \mathbf{s}_{\tau} + \mathbf{e}_{\tau,2} + \tau
|
|
\cdot \lfloor q/2 \rfloor \big) \in \Zq^m \times \Zq^\ell, \\
|
|
%
|
|
\mathbf{c}_{k} & = (\mathbf{c}_{k,1},\mathbf{c}_{k,2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{k} + \mathbf{e}_{k,1} ,~ \mathbf{G}_1^T \cdot \mathbf{s}_{k} + \mathbf{e}_{k,2} + \mathfrak{m}_k \cdot \lfloor q/2 \rfloor \big) \in \Zq^m \times \Zq^{2m} \\
|
|
& \hspace{7.6cm} \forall k\in \{1,\ldots,N\}
|
|
\end{align*}
|
|
where $\mathbf{s}_{\tau}, \mathbf{s}_{k} \sample \chi^n$, $\mathbf{e}_{\tau,1} , \mathbf{e}_{k,1} \sample \chi^m$, $\mathbf{e}_{\tau,2} \sample \chi^\ell$, $\mathbf{e}_{k,2} \sample \chi^{2m}$,
|
|
as well as
|
|
\begin{align*}
|
|
\mathbf{c}_{\mathbf{v}} & = (\mathbf{c}_{\mathbf{v},1},\mathbf{c}_{\mathbf{v},2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{\mathbf{v}} + \mathbf{e}_{\mathbf{v},1} ,~ \mathbf{G}_1^T \cdot \mathbf{s}_{\mathbf{v}} + \mathbf{e}_{\mathbf{v},2} + \mathbf{v} \cdot \lfloor q/p \rfloor \big) \in \Zq^m \times \Zq^{2m} \\
|
|
\mathbf{c}_{s} & = (\mathbf{c}_{s,1},\mathbf{c}_{s,2}) \\
|
|
& = \big( \mathbf{B}^T \cdot \mathbf{s}_{0} + \mathbf{e}_{0,1} ,~ \mathbf{G}_1^T \cdot \mathbf{s}_{0} + \mathbf{e}_{0,2} + \mathbf{s} \cdot \lfloor q/p \rfloor \big) \in \Zq^m \times \Zq^{2m} ,
|
|
\end{align*}
|
|
where $\mathbf{s}_{\mathbf{v}}, \mathbf{s}_{0} \sample \chi^n$, $\mathbf{e}_{\mathbf{v},1},\mathbf{e}_{0,1} \sample \chi^m$,
|
|
$\mathbf{e}_{\mathbf{v},2},\mathbf{e}_{0,2}\sample \chi^{2m}$.
|
|
\item[2.] Prove in zero-knowledge that $\mathbf{c}_{\tau}$, $\mathbf{c}_{s}$, $\mathbf{c}_{\mathbf{v} }$, $\{\mathbf{c}_k\}_{k=1}^N$ encrypt a valid message-signature pair. In Section~\ref{subsection:zk-for-signature}, we show that this involved zero-knowledge protocol can be derived from the statistical zero-knowledge argument of knowledge for a simpler, but more general relation that we explicitly present in \cref{se:gs-lwe-stern}. The proof system can be made statistically ZK for a malicious verifier using standard techniques (assuming a common reference string, we can use \cite{Dam00}). In the random oracle model, it can
|
|
be made non-interactive using the Fiat-Shamir heuristic \cite{FS86}.
|
|
|
|
\end{itemize}
|
|
\end{description}
|
|
|
|
We require that the adversary be unable to prove possession of a signature of a message $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$ for which it did not legally
|
|
obtain a credential by interacting with the issuer. Note that the messages that are blindly signed by the issuer are uniquely defined since, at each signing
|
|
query, the adversary is required to supply perfectly binding commitments $\{\mathbf{c}_k\}_{k=1}^N$ to $(\mathfrak{m}_1,\ldots,\mathfrak{m}_N)$.
|
|
|
|
In instantiations using non-interactive proofs, we assume that these can be bound to a verifier-chosen nonce to prevent replay attacks, as suggested in \cite{CKL+15}.
|
|
|
|
The security proof (in Theorem \ref{commit-thm}) makes crucial use of the R\'enyi divergence using arguments in the spirit of Bai \textit{et al.} \cite{BLL+15}. The
|
|
reduction has to guess upfront the index $i^\star \in \{1,\ldots,Q\}$ of the specific signing query for which the adversary will re-use $\tau^{(i^\star)}$. For
|
|
this query, the reduction will have to make sure that the simulation trapdoor of Agrawal \textit{et al.} \cite{ABB10} (used by the $\mathsf{SampleRight}$ algorithm
|
|
of Lemma \ref{lem:sampler}) vanishes: otherwise, the adversary's forgery would not be usable for solving $\mathsf{SIS}$. This means that, as in the proof of
|
|
\cite{BHJ+15}, the reduction must answer exactly one signing query in a different way, without using the trapdoor. While B\"ohl \textit{et al.} solve this
|
|
problem by exploiting the fact that they only need to prove security against non-adaptive forgers, we directly use a built-in chameleon hash function mechanism
|
|
which is implicitly realized by the matrix $\mathbf{D}_0$ and the vector $\mathbf{s}$. Namely, in the signing query for which the Agrawal \textit{et al.}
|
|
trapdoor~\cite{ABB10} cancels, we assign a special value to the vector $\mathbf{s} \in \ZZ^{2m}$, which depends on the adaptively-chosen signed message
|
|
$(\mathsf{Msg}_1^{(i^\star)},\ldots,\mathsf{Msg}_N^{(i^\star)})$ and some Gaussian matrices $\{\mathbf{R}_k\}_{k=1}^N$ hidden behind $\{\mathbf{D}_k\}_{k=1}^N$.
|
|
|
|
One issue is that this results in a different distribution for the vector $\mathbf{s} \in \ZZ^m$. However, we can still view $\mathbf{s}$ as a vector sampled from a
|
|
Gaussian distribution centered away from $\mathbf{0}^{2m}$. Since this specific situation occurs only once during the simulation, we can apply a result proved in
|
|
\cite{LSS14} which upper-bounds the R\'enyi divergence between two Gaussian distributions with identical standard deviations but different centers. By
|
|
choosing the standard deviation $\sigma_1$ of $\mathbf{s} \in \ZZ^{2m}$ to be polynomially larger than that of the columns of matrices $\{\mathbf{R}_k\}_{k=1}^N$, we can
|
|
keep the R\'enyi divergence between the two distributions of $\mathbf{s}$ (i.e., the one of the simulation and the one of the real game) sufficiently small to apply
|
|
the probability preservation property (which still gives a polynomial reduction since the argument must only be applied on one signing query). Namely, the
|
|
latter implies that, if the R\'enyi divergence $R_2(\mathbf{s}^{\mathsf{real}}||\mathbf{s}^{\mathsf{sim}})$ is polynomial, the probability that the simulated vector
|
|
$\mathbf{s}^{\mathsf{sim}} \in \ZZ^{2m}$ passes the verification test will only be polynomially smaller than in the real game and so will be the adversary's
|
|
probability of success.
|
|
|
|
Another option would have been to keep the statistical distance between $\mathbf{s}^{\mathsf{real}}$ and $\mathbf{s}^{\mathsf{sim}}$ negligible using the smudging
|
|
technique of \cite{AJL+12}. However, this would have implied to use an exponentially large modulus $q$ since $\sigma_1$ should have been exponentially larger
|
|
than the standard deviations of the columns of $\{\mathbf{R}_k\}_{k=1}^N$.
|
|
|
|
\begin{theorem} \label{commit-thm}
|
|
Under the $\mathsf{SIS}_{n,2m, q, \hat{\beta}}$ assumption, where $\hat{\beta} = N \sigma (2m)^{3/2} + 4 \sigma_1 m^{3/2}$\hspace*{-1.5pt}, the above
|
|
protocols are secure protocols for obtaining a signature on a committed message and proving possession of a valid message-signature pair.
|
|
\end{theorem}
|
|
|
|
In the following proof, we make use of the Rényi divergence in a similar way to~\cite{BLL+15}:
|
|
instead of the classical statistical distance we sometimes use the R\'enyi divergence, which is a measurement of the distance between two distributions.
|
|
Its use in security proofs for lattice-based systems was first considered by Bai {\em et al.}~\cite{BLL+15} and further improved by Prest~\cite{Pre17}. We first recall its definition.
|
|
|
|
\defRenyi*
|
|
|
|
|
|
We will focus on the following properties of the R\'enyi divergence, the proofs can be found in~\cite{LSS14}.
|
|
|
|
\begin{lemma}[{\cite[Le. 2.7]{BLL+15}}]
|
|
\label{lem:renyi}
|
|
Let $a \in [1, +\infty]$. Let $P$ and $Q$ denote distributions with $\Supp(P)
|
|
\subseteq \Supp(Q)$. Then the following properties hold:
|
|
\begin{description}
|
|
\item[Log. Positivity:] $R_a(P||Q) \geq R_a(P||P) = 1$
|
|
\item[Data Processing Inequality:] $R_a(P^f || Q^f) \leq R_a(P||Q)$ for any
|
|
function $f$, where $P^f$ denotes the distribution of $f(y)$ induced by
|
|
sampling $y \sample P$ (resp. $y \sample Q$)
|
|
\item[Multiplicativity:] Assume $P$ and $Q$ are two distributions of a pair
|
|
of random variables $(Y_1, Y_2)$. For $i \in \{1,2\}$, let $P_i$ (resp.
|
|
$Q_i$) denote the marginal distribution of $Y_i$ under $P$ (resp. $Q$),
|
|
and let $P_{2|1}(\cdot|y_1)$ (resp. $Q_{2|1}(\cdot|y_1)$) denote the conditional distribution of $Y_2$ given that $Y_1 = y_1$. Then we have:
|
|
\begin{itemize} \renewcommand\labelitemi{$\bullet$}
|
|
\item $R_a(P||Q) = P_a(P_1 || Q_1) \cdot R_a(P_2||Q_2)$ if $Y_B$ and $Y_2$ are independent;
|
|
\item $R_a(P||Q) \leq R_\infty (P_1 || Q_1) \cdot max_{y_1 \in X} R_a\left( P_{2|1}(\cdot | y_1) || Q_{2|1}(\cdot | y_1) \right)$.
|
|
\end{itemize}
|
|
\item[Probability Preservation:] Let $A \subseteq \Supp(Q)$ be an arbitrary
|
|
event. If $a \in ]1, +\infty[$, then $Q(A) \geq
|
|
P(A)^{\frac{a}{a-1}}/R_a(P||Q)$. Further we have:
|
|
\[ Q(A) \geq P(A) / R_\infty(P||Q) \]
|
|
\item[Weak Triangle Inequality:] Let $P_1, P_2, P_3$ be three distributions
|
|
with \[\Supp(P_1) \subseteq \Supp(P_2) \subseteq \Supp(P_3).\]
|
|
Then we have:
|
|
\[ R_a(P_1||P_3) \leq \begin{cases}
|
|
R_a(P_1 || P_2) \cdot R_\infty(P_2 || P_3),\\[2mm]
|
|
R_\infty(P_1||P_2)^{\frac{a}{a-1}} \cdot R_a(P_2||P_3) & \mbox{if } a \in ]1, +\infty[.
|
|
\end{cases}\]
|
|
\end{description}
|
|
\end{lemma}
|
|
|
|
In our proofs, we mainly use the probability preservation to bound the
|
|
probabilities during hybrid games where the two distributions are not close in terms of statistical distance.
|
|
|
|
%--------- PROOF ----------
|
|
\begin{proof} The proof is very similar to the proof of \cref{th:gs-lwe-security-cma-sig} and we will only explain the changes.
|
|
|
|
Let us assume that an adversary $\adv$ can prove possession of a signature on a message $(\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star)$ which has not been blindly signed by the issuer,
|
|
we outline an algorithm $\bdv$ that solves a $\mathsf{SIS}_{n,2m,q,\beta}$ instance $\bar{\mathbf{A}}$, where $\bar{\mathbf{A}} =
|
|
[ \bar{\mathbf{A}}_1 \mid \bar{\mathbf{A}}_2 ] \in \ZZ_q^{ n \times 2m}$ with matrices
|
|
$\bar{\mathbf{A}}_1, \bar{\mathbf{A}}_2 \sample \U(\ZZ_q^{n \times m})$.
|
|
|
|
At the outset of the game, $\bdv$ generates the common parameters $\mathsf{par}$ by choosing
|
|
$\mathbf{B} \in_R \ZZ_q^{n \times m}$ and defining $\mathbf{G}_0 = \mathbf{B} \cdot \mathbf{E}_0 \in \ZZ_q^{n \times \ell }$, $\mathbf{G}_1 = \mathbf{B} \cdot \mathbf{E}_1 \in \ZZ_q^{n \times 2m}$.
|
|
The short Gaussian matrices $\mathbf{E}_0 \in \ZZ^{ m \times \ell}$ and $\mathbf{E}_1 \in \ZZ^{m \times 2m}$ are retained for later use. Also, $\bdv$ flips a coin $coin \in \{0,1,2\}$ as
|
|
a guess for the kind of attack that $\adv$ will mount. If $coin=0$, $\bdv$ expects a Type I forgery, where $\adv$'s forgery involves a new $\tau^\star \in \{0,1\}^\ell$ that
|
|
was never used by the signing oracle. If $coin=1$, $\bdv$ expects $\adv$ to recycle a tag $\tau^\star$ involved in some signing query in its forgery. Namely,
|
|
if $coin=1$, $\bdv$ expects an attack which is either a Type II forgery or a Type III forgery.
|
|
If $coin=2$, $\bdv$ rather bets that $\adv$ will break the soundness of the interactive argument systems used in the signature issuing protocol or the $\mathsf{Prove}$ protocol.
|
|
Depending on the value of $coin \in \{0,1,2 \}$, $\bdv$ generates the issuer's public key $PK$ and simulates $\adv$'s view in different ways.
|
|
|
|
\begin{itemize}
|
|
\item If $coin=0$, $\bdv$ undertakes to find a short non-zero vector of $\Lambda_q^{\perp}(\bar{\mathbf{A}}_1)$, which in turn yields a short non-zero vector
|
|
of $\Lambda_q^{\perp}(\bar{\mathbf{A}})$. To this end, it defines $\mathbf{A}=\bar{\mathbf{A}}_1$ and
|
|
generates $PK$ by computing $\{\mathbf{A}_j\}_{j=0}^\ell$ as re-randomizations of $\mathbf{A} \in \ZZ_q^{n \times m}$ as in the proof of Lemma \ref{le:lwe-gs-type-I-attacks}. This implies that $\bdv$ can always answer signing queries using the trapdoor $\mathbf{T}_{\mathbf{C}}
|
|
\in \ZZ^{m \times m}$ of the matrix $\mathbf{C}$ without even knowing the messages hidden in the commitments $ \mathbf{c}_{\mathfrak{m}}$ and $\{\mathbf{c}_k\}_{k=1}^N$, $\mathbf{c}_{s'}$.
|
|
When the adversary generates a proof of possession of its own at the end of the game, $\bdv$ uses the matrices $\mathbf{E}_0 \in \ZZ^{ m \times \ell}$ and $\mathbf{E}_1 \in \ZZ^{m \times 2m}$
|
|
as an extraction trapdoor to extract a plain message-signature pair $\big( (\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star), (\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star) \big)$
|
|
from the ciphertexts
|
|
$\{ \mathbf{c}_k^\star\}_{k=1}^N$ $(\mathbf{c}_{\mathbf{v}_1}^\star,\mathbf{c}_{\mathbf{v}_2^\star})$, $\mathbf{c}_{\tau}^\star$, $\mathbf{c}_{\mathbf{s}}^\star$ produced by $\adv$ as part of its forgery.
|
|
If the extracted $\tau^\star$ is not a new tag, then $\bdv$ aborts. Otherwise, it can solve the given $\mathsf{SIS}$ instance exactly as in the proof of Lemma \ref{le:lwe-gs-type-I-attacks}.
|
|
|
|
\item If $coin=1$, the proof proceeds as in the proof of Lemma \ref{le:lwe-gs-type-II-attacks} with one difference in \textsf{Game} $3$. This difference is that \textsf{Game} $3$ is no longer statistically
|
|
indistinguishable from \textsf{Game} $2$: instead, we rely on an argument based on the R\'enyi divergence.
|
|
In \textsf{Game} $3$, $\bdv$ generates $PK$ exactly as in the proof of Lemma \ref{le:lwe-gs-type-II-attacks}. This implies that $\bdv$ takes a guess $i^\dagger \leftarrow U(\{1,\ldots,Q\})$
|
|
with the hope that $\adv$ will choose to recycle the tag $\tau^{(i^\dagger)} $ of the $i^\dagger$-th signing query (i.e., $ \tau^\star =\tau^{(i^\dagger)} $).
|
|
As in the proof of Lemma \ref{le:lwe-gs-type-II-attacks}, $\bdv$ defines $\mathbf{D}=\bar{\mathbf{A}}_1 \in \ZZ_q^{n \times m}$ and $\mathbf{A}= \bar{\mathbf{A}}_1 \cdot \mathbf{S} $ for a small-norm
|
|
matrix $\mathbf{S} \in \ZZ^{m \times m}$ with Gaussian entries. It also ``programs'' the matrices $\{ \mathbf{A}_j\}_{j=0}^\ell$ in such a way that
|
|
the trapdoor precisely vanishes at the $i^\dagger$-th signing query: in other words,
|
|
the sum $$\mathbf{A}_0 + \sum_{j=1}^\ell \tau^{(i)} [j] \mathbf{A}_j = \bar{\mathbf{A}}_1 \cdot (\mathbf{S}_0 + \sum_{j=1}^\ell \tau^{(i)} [j] \cdot \mathbf{S}_j) + (h_0 + \sum_{j=1}^\ell \tau^{(i)} [j] \cdot h_j ) \cdot \mathbf{C} $$ does not depend on the matrix $\mathbf{C} \in \ZZ_q^{n \times m}$
|
|
(of which a trapdoor $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$ is known to $\bdv$) when $\tau^{(i)} = \tau^{(i^\dagger)}$, but it does for all other tags $\tau^{(i)} \neq \tau^{(i^\dagger)}$. In the setup phase,
|
|
$\bdv$ also sets up a random matrix $\mathbf{D}_0 \in U(\ZZ_q^{2n \times 2m})$ which it obtains by choosing
|
|
$\mathbf{A}' \sample (\ZZ_q^{n \times 2m})$ to define
|
|
\begin{eqnarray} \label{def-D0}
|
|
\mathbf{D}_0=\begin{bmatrix} \bar{\mathbf{A}} \\ \hline \mathbf{A}' \end{bmatrix} \in \ZZ_q^{2n \times 2m}.
|
|
\end{eqnarray}
|
|
Then, it computes $\mathbf{c}_M = \mathbf{D}_0 \cdot \mathbf{s}_0 \in \ZZ_q^{2n}$ for a short Gaussian vector
|
|
$\mathbf{s}_0 \sample D_{\ZZ^{2m},\sigma_0}$, which will be used in the $i^\dagger$-th query.
|
|
Next, it samples short vectors $\mathbf{v}_1,\mathbf{v}_2 \sample D_{\ZZ^m,\sigma}$ to define
|
|
$$\mathbf{u}= \mathbf{A}_{\tau^{(i^\dagger)}} \cdot \begin{bmatrix} \mathbf{v}_1 \\ \mathbf{v}_2 \end{bmatrix} - \mathbf{D} \cdot \textsf{bin}(\mathbf{c}_M) ~ \in \ZZ_q^n.$$
|
|
In addition, $\bdv$ picks extra small-norm matrices $\mathbf{R}_1,\ldots,\mathbf{R}_N \in \ZZ^{2m \times 2m}$ whose columns are sampled from $D_{\ZZ^m,\sigma}$, which
|
|
are used to define randomizations of $\mathbf{D}_0$ by computing $\mathbf{D}_k = \mathbf{D}_0 \cdot \mathbf{R}_k$ for each $k \in \{1,\ldots,N\}$.
|
|
The adversary is given public parameters $\mathsf{par}\coloneqq \{\mathbf{B},\mathbf{G}_0,\mathbf{G}_1,CK\}$, where $CK=\{\mathbf{D}_k\}_{k=0}^N$, and the public key $PK\coloneqq \big( \mathbf{A}, \{\mathbf{A}_j\}_{j=0}^\ell, \mathbf{D},\mathbf{u} \big)$.
|
|
\end{itemize}
|
|
|
|
Using $\mathbf{T}_{\mathbf{C}}$,
|
|
$\bdv$ can perfectly emulate the signing oracle at all queries, except the $i^\dagger$-th query where the
|
|
vector ${\mathbf{s}''}^{(i^\dagger)}$ chosen by $\bdv$ is sampled from a distribution that departs from $D_{\ZZ^{2m},\sigma_0}$. At the $i^\dagger$-th query,
|
|
$\bdv$ uses the extraction trapdoor $\mathbf{E}_1 \in \ZZ^{m \times 2m}$ to obtain $ {\mathbf{s}' }^{(i^\dagger)} \in \ZZ^{2m}$ and $\{\mathfrak{m}_k\}_{k=1}^N$ -- which form a valid opening
|
|
of $\mathbf{c}_{\mathfrak{m}}$ unless the soundness of the proof system is broken (note that the latter case is addressed by the situation $coin=3$) -- from the ciphertexts
|
|
$\mathbf{c}_{s'}^{(i^\dagger)} $ and $\{ \mathbf{c}_k\}_{k=1}^N$ sent by $\adv$ at step 1 of the signing protocol. Then, $\bdv$
|
|
computes the vector ${\mathbf{s}''}^{(i^\dagger)}$ as
|
|
\begin{eqnarray} \label{sim-s-prime}
|
|
{\mathbf{s}'' }^{(i^\dagger)} = \mathbf{s}_0 - \sum_{k=1}^N \mathbf{R}_k \cdot \mathfrak{m}_k^{(i^\dagger)} - {\mathbf{s}' }^{(i^\dagger)} \in \ZZ^{2m},
|
|
\end{eqnarray}
|
|
which satisfies $\mathbf{c}_M=\sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^{(i^\dagger)} + \mathbf{D}_0 \cdot ({\mathbf{s}' }^{(i^\dagger)} + {\mathbf{s}'' }^{(i^\dagger)} ) $ and
|
|
allows returning $(\tau^{(i^\dagger)},\mathbf{v}^{(i^\dagger)}, {\mathbf{s}'' }^{(i^\dagger)} )$ such that
|
|
$(\tau^{(i^\dagger)},\mathbf{v}^{(i^\dagger)}, {\mathbf{s}' }^{(i^\dagger)} + {\mathbf{s}'' }^{(i^\dagger)} )$ satisfies the verification
|
|
equation of the signature scheme. Moreover, we argue that, with noticeable probability, the integer
|
|
vector ${\mathbf{s} }^{(i^\dagger)} ={\mathbf{s}' }^{(i^\dagger)} + {\mathbf{s}'' }^{(i^\dagger)}$ will be accepted by the verification algorithm since the R\'enyi divergence
|
|
between the simulated distribution of ${\mathbf{s}'' }^{(i^\dagger)}$ and its distribution in the real game will be sufficiently small. Indeed, its distribution
|
|
is now that of a Gaussian vector $D_{\ZZ^{2m},\sigma_0,\mathbf{z}^\dagger }$ centered in $$\mathbf{z}^\dagger = - \sum_{k=1}^N
|
|
\mathbf{R}_k \cdot {\mathfrak{m}_k^{(i^\dagger)} }
|
|
- {\mathbf{s}' }^{(i^\dagger)} \in \ZZ^{2m} ,$$ whose norm is at most $\| \mathbf{z}^\dagger \|_2 \leq N \sigma ({2m})^{3/2} + \sigma (2m)^{1/2}$. By choosing the standard deviation $\sigma_0$ to
|
|
be at least
|
|
$\sigma_0> N \sigma (2m)^{3/2} + \sigma (2m)^{1/2} $, the R\'enyi divergence between the simulated
|
|
distribution of ${\mathbf{s}'' }^{(i^\dagger)}$ (in \textsf{Game} $3$) and its real distribution (which is the one of \textsf{Game} $2$) can be kept constant: we have
|
|
\begin{eqnarray} \label{r-bound}
|
|
R_2( {\mathbf{s}'' }^{(i^\dagger),2} ||{\mathbf{s}'' }^{(i^\dagger),3} ) \leq \exp \big( 2\pi \cdot \frac{ \| \mathbf{z}^\dagger \|_2^2}{\sigma_0^2} \big) \leq \exp(2 \pi).
|
|
\end{eqnarray}
|
|
This ensures that, with noticeable
|
|
probability, $(\tau^{(i^\dagger)},\mathbf{v}^{(i^\dagger)}, {\mathbf{s} }^{(i^\dagger)} )$ will pass the verification test and lead $\adv$ to eventually output a valid forgery.
|
|
So, the success probability of $\adv$ in \textsf{Game} $3$ remains noticeable as (\ref{r-bound}) implies $\Pr[W_3] \geq \Pr[W_2]^2 / \exp(2\pi)$.
|
|
|
|
When $W_3$ occurs in \textsf{Game} $3$, $\bdv$ uses the matrices $(\mathbf{E}_0,\mathbf{E}_1)$ to extract a plain message-signature pair $\big((\mathfrak{m}_1^\star,\ldots,\mathfrak{m}_N^\star),(\tau^\star,\mathbf{v}^\star,\mathbf{s}^\star) \big)$ from the extractable commitments
|
|
$\{ \mathbf{c}_k^\star\}_{k=1}^N$ $(\mathbf{c}_{\mathbf{v}_1}^\star,\mathbf{c}_{\mathbf{v}_2}^\star)$, $\mathbf{c}_{\tau}^\star$, $\mathbf{c}_{\mathbf{s}}^\star$ generated by $\adv$.
|
|
At this point, two cases can be distinguished. First, if $\mathbf{c}_M \neq \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^\star + \mathbf{D}_0 \cdot \mathbf{s}^\star \bmod q$, then algorithm
|
|
$\bdv$ can
|
|
find a short vector of $\Lambda_q^{\perp}(\bar{\mathbf{A}}_1)=\Lambda_q^{\perp}( {\mathbf{D}})$ exactly as in the proof of Lemma~\ref{le:lwe-gs-type-II-attacks}. In the event that $\mathbf{c}_M = \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k^\star + \mathbf{D}_0 \cdot \mathbf{s}^\star $,
|
|
$\bdv$ can use the fact that the collision $\mathbf{c}_M = \sum_{k=1}^N \mathbf{D}_k \cdot {\mathfrak{m}_k^{(i^\dagger)} } + \mathbf{D}_0 \cdot {\mathbf{s}^{(i^\dagger)} } $ allows computing
|
|
$$ \mathbf{w}= \mathbf{s}^\star -{\mathbf{s}^{(i^\dagger)}} + \sum_{k=1}^N \mathbf{R}_k \cdot \left(\mathfrak{m}_k^\star - \mathfrak{m}_k^{(i^\dagger)} \right) ~ \in \ZZ^{2m} , $$
|
|
which belongs to $\Lambda_q^{\perp}(\mathbf{D}_0)$ and has norm $\| \mathbf{w} \|_2 \leq N \sigma (2m)^{3/2} + 4 \sigma_1 m^{3/2} $. Moreover, it
|
|
is non-zero with overwhelming probability. Indeed, there exists at least one $k \in [1,N]$ such that $\mathfrak{m}_k^{(i^\dagger)} \neq \mathfrak{m}_k^\star$. Let us assume w.l.o.g.
|
|
that they differ in their first two bits where $\mathfrak{m}_k^{(i^\dagger)}$ contains a $0$ and $\mathfrak{m}_k^\star$ contains a $1$ (recall that each bit $b$
|
|
is encoded as $(\bar{b},b)$ in both messages).
|
|
This implies that $ {\mathbf{s}'' }^{(i^\dagger)} $ (as computed in (\ref{sim-s-prime})) does not depend on the first column of $\mathbf{R}_k$ but $\mathbf{w}$ does.
|
|
Hence, given that the columns of $\mathbf{R}_k$ have at least $n$ bits of min-entropy conditionally on $\mathbf{D}_k =\mathbf{D}_0 \cdot \mathbf{R}_k$, the vector
|
|
$\mathbf{w} \in \ZZ^{2m}$ is unpredictable to the adversary.
|
|
|
|
Due to the definition of $\mathbf{D}_0 \in \ZZ_q^{2n \times 2m}$ in (\ref{def-D0}), we finally note that
|
|
$\mathbf{w} \in \ZZ^{2m}$ is also a short non-zero vector of $\Lambda_q^{\perp}(\bar{\mathbf{A}})$.
|
|
|
|
\begin{itemize}
|
|
\item If $coin=2$, $\bdv$ faithfully generates $\mathsf{par}$ and $PK$, but it retains the extraction trapdoor $(\mathbf{E}_0,\mathbf{E}_1)$ associated with the dual Regev public keys
|
|
$(\mathbf{G}_0,\mathbf{G}_1)$. Note that $\adv$ can break the soundness of the proof system by either: (i) Generating ciphertexts
|
|
$\{\mathbf{c}_k\}_{k=1}^N$ and $\mathbf{c}_{s'}$ that do not encrypt an opening of $\mathbf{c}_{\mathfrak{m}}$ in the signature issuing protocol; (ii) Generating ciphertexts
|
|
$\{\mathbf{c}_k\}_{k=1}^N$, $\mathbf{c}_{\tau}$, $\mathbf{c}_{\mathbf{v}_1}$, $\mathbf{c}_{\mathbf{v}_2}$ and $\mathbf{c}_{s}$ that do not encrypt a valid signature in the $\mathsf{Prove}$ protocol.
|
|
In either case, the reduction $\bdv$ is able to detect the event by decrypting dual Regev ciphertext using $(\mathbf{E}_0,\mathbf{E}_1)$ and create a breach in the
|
|
soundness of the argument system.
|
|
\end{itemize}
|
|
|
|
It it easy to see that, since $coin \in \{0,1,2 \}$ is chosen independently of $\adv$'s view, it turns out to be correct with probability $1/3$. As a consequence, if $\adv$'s advantage
|
|
is non-negligible, so is $\bdv$'s.
|
|
\end{proof}
|
|
|
|
\begin{theorem} \label{anon-cred}
|
|
The scheme provides anonymity under the $\mathsf{LWE}_{n,q,\chi}$ assumption.
|
|
\end{theorem}
|
|
|
|
\begin{proof}
|
|
The proof is rather straightforward and consists of a sequence of three games.
|
|
\medskip
|
|
|
|
\begin{description}
|
|
\item[\textsf{Game} 0:] This is the real game. Namely, the adversary is given common public parameters $\mathsf{par}$ and comes up with a public key $PK$ of its own.
|
|
The adversary can run oblivious signing protocols with honest users. At each query, the adversary chooses a user index $i$ and triggers an execution of the signing protocol
|
|
with the challenger emulating the honest users. At some point, the adversary chooses some user index $i^\star$ for which the execution of the signing protocol ended successfully.
|
|
At this point, the challenger $\bdv$ runs the real $\mathsf{Prove}$ protocol on behalf of user $i$. At the end of the game, the adversary outputs
|
|
a bit $b' \in \{0,1\}$. We define $W_0$ to be the event that
|
|
$b'=1$.
|
|
\smallskip
|
|
|
|
\item[\textsf{Game} 1:] This game is like \textsf{Game} $0$ with the difference that, at each execution of the $\mathsf{Prove}$ protocol, the challenger runs the zero-knowledge
|
|
simulator of the interactive proof system. The latter simulator uses either a trapdoor hidden in the common reference string (if Damg\aa rd's technique \cite{Dam00} is used) or
|
|
proceeds by programming the random oracle which allows implementing the Fiat-Shamir heuristic. In either case, the statistical zero-knowledge property ensures that the
|
|
adversary cannot distinguish \textsf{Game} $1$ from \textsf{Game} $0$ and $|\Pr[W_1] - \Pr[W_0] | \in \mathsf{negl}(\lambda)$.
|
|
\smallskip
|
|
|
|
\item[Game 3:] This game is like \textsf{Game} $1$ except that, at each execution of the $\mathsf{Prove}$ protocol, the ciphertexts $\{\mathbf{c}_k\}_{k=1}^N$, $\mathbf{c}_s$, $\mathbf{c}_{\tau}$,
|
|
and $\mathbf{c}_{\mathbf{v}_1}$, $\mathbf{c}_{\mathbf{v}_2}$ encrypt random messages instead of the actual witnesses. The semantic security of the dual Regev cryptosystem ensures that,
|
|
under the $\LWE_{n,q,\chi}$ assumption, the adversary is unable to see the difference. Hence, we have $|\Pr[W_2] - \Pr[W_1]| \leq \mathbf{Adv}_{\bdv}^{\mathsf{LWE}}(\lambda)$.
|
|
\end{description}
|
|
\medskip
|
|
|
|
In \textsf{Game} $2$, we can notice that the adversary is interacting with a simulator that emulates the user in the $\mathsf{Prove}$ protocol \textit{without} using
|
|
any message-signature pair. We thus conclude that, under the $\LWE_{n,q,\chi}$ assumption, $\adv$'s view cannot distinguish a real proof of signature possession from a simulated proof
|
|
produced without any witness.
|
|
\end{proof}
|
|
|
|
\section{A Dynamic Lattice-Based Group Signature} \label{see:lwe-gs-desc}
|
|
|
|
In this section, the signature scheme of Section \ref{se:gs-lwe-sigep} is used to design a group signature for dynamic groups using the syntax and the security model of Kiayias and Yung \cite{KY06}, which is recalled in \cref{sse:gs-definitions}.
|
|
|
|
In the notations hereunder, for any positive integers $\mathfrak{n}$, and $q \geq 2$, we define the ``powers-of-2'' matrix $\mathbf{H}_{\mathfrak{n} \times \mathfrak{n}\lceil\log q\rceil} \in \ZZ_q^{\mathfrak{n} \times \mathfrak{n}\lceil\log q\rceil}$ to be:
|
|
\begin{eqnarray*}
|
|
\mathbf{H}_{\mathfrak{n} \times \mathfrak{n} \lceil\log q\rceil } &=& \mathbf{I}_{\mathfrak{n}} \otimes [1 \mid 2 \mid 4 \mid \ldots \mid 2^{\lceil\log q\rceil-1} ] .
|
|
\end{eqnarray*}
|
|
Also, for each vector $\mathbf{v} \in \ZZ_q^{\mathfrak{n}}$, we define $\textsf{bin}(\mathbf{v}) \in \{0,1\}^{\mathfrak{n}\lceil\log q\rceil}$ to be the vector obtained by replacing each entry of $\mathbf{v}$ by its binary expansion.
|
|
Hence, we have $\mathbf{v}=\mathbf{H}_{\mathfrak{n} \times \mathfrak{n}\lceil\log q\rceil} \cdot \textsf{bin}(\mathbf{v})$ for any $\mathbf{v} \in \ZZ_q^{\mathfrak{n}}$.
|
|
|
|
In our scheme, each group membership certificate is a
|
|
signature generated by the group manager on the user's public key. Since the group manager only needs to sign known (rather than committed) messages, we can
|
|
use a simplified version of the signature, where the chameleon hash function does not need to choose
|
|
the discrete Gaussian vector $\mathbf{s}$ with a larger standard deviation than other vectors.
|
|
|
|
A key component of the scheme is the two-message joining protocol whereby the group manager admits new group members by signing their public key. The first message is sent by
|
|
the new user $\mathcal{U}_i$ who samples a membership secret consisting of a short vector $\mathbf{z}_i \sample D_{\ZZ^{4m},\sigma}$ (where $m= 2n \lceil\log q\rceil$), which is used to compute a
|
|
syndrome $\mathbf{v}_i = \mathbf{F} \cdot \mathbf{z}_i \in \ZZ_q^{4n}$ for some public matrix $\mathbf{F} \in \ZZ_q^{4n \times 4m} $. This syndrome $\mathbf{v}_i \in \ZZ_q^{4n}$ must be signed by $\mathcal{U}_i$ using his long term secret key $\mathsf{usk}[i]$ (as in
|
|
\cite{KY06,BSZ05}, we assume that each user has a long-term key $\mathsf{upk}[i]$ for a digital signature, which is registered in some PKI) and will uniquely
|
|
identify $\mathcal{U}_i$.
|
|
In order to generate a membership certificate for $\mathbf{v}_i \in \ZZ_q^{4n}$, the group manager $\mathsf{GM}$ signs its binary expansion
|
|
$\mathsf{bin}(\mathbf{v}_i) \in \{0,1\}^{4n \lceil \log q \rceil }$ using the scheme of Section \ref{se:gs-lwe-sigep}.
|
|
|
|
Equipped with his membership
|
|
certificate $(\tau,\mathbf{d},\mathbf{s}) \in \{0,1\}^\ell \times \ZZ^{2m} \times \ZZ^{2m}$, the new group member $\mathcal{U}_i$ can sign a message using a Stern-like protocol for
|
|
demonstrating his knowledge of
|
|
a valid certificate for which he also knows the secret key associated with the certified public key $\mathbf{v}_i \in \ZZ_q^{4n}$. This boils down to
|
|
providing evidence that the membership certificate is a valid signature on some binary message $\mathsf{bin}(\mathbf{v}_i) \in \{0,1\}^{4n \lceil \log q \rceil }$
|
|
for which he also knows a short $\mathbf{z}_i \in \ZZ^{4m}$
|
|
such that
|
|
$ \mathbf{v}_i = \mathbf{H}_{4n \times 2m} \cdot \textsf{bin}(\mathbf{v}_i) = \mathbf{F} \cdot \mathbf{z}_i \in \mathbb{Z}_q^{4n}$.
|
|
|
|
Interestingly, the process does not require any proof of knowledge of the membership secret $\mathbf{z}_i$ during the joining phase, which is round-optimal. Analogously to the Kiayias-Yung technique \cite{KY05} and constructions based on structure-preserving signatures
|
|
\cite{AFG+10}, the joining protocol thus remains secure in environments where many users want
|
|
to register at the same time in concurrent sessions.
|
|
|
|
We remark that a similar Stern-like protocol could also be directly used to prove knowledge of a Boyen signature \cite{Boy10} on a binary expansion of the
|
|
user's syndrome $\mathbf{v}_i \in \ZZ_q^{4n}$ while preserving the user's ability to prove knowledge of a short $\mathbf{z}_i \in \ZZ^{4m}$ such that $\mathbf{F} \cdot \mathbf{z}_i =
|
|
\mathbf{v}_i \bmod q$. However, this would require considerably longer private keys containing $ 4n \cdot \log q$ matrices $\{\mathbf{A}_j\}_{j=0}^\ell$ of dimension $n \times
|
|
m$ each (i.e., we would need $\ell= \Theta(n \cdot \log q)$). In contrast, by using the signature scheme of Section \ref{se:gs-lwe-sigep}, we only need the group public key
|
|
$\mathcal{Y}$ to contain $\ell=\log N_{\mathsf{gs}}$ matrices in $\ZZ_q^{n \times m}$. Since the number of users $N_{\mathsf{gs}}$ is polynomial, we have $\log
|
|
N_{\mathsf{gs}} \ll n$, which results in a much more efficient scheme.
|
|
|
|
|
|
|
|
|
|
\subsection{Description of the Scheme}
|
|
|
|
\begin{description}
|
|
\item[\textsf{Setup}$(1^\lambda,1^{N_{\mathsf{gs}}})$:] Given a security parameter $\lambda>0$
|
|
and the maximal expected number of group members ${N_{\mathsf{gs}}}=2^{\ell} \in
|
|
\mathsf{poly}(\lambda)$, choose lattice parameter
|
|
$n = \mathcal{O}(\lambda)$; prime modulus $q = \widetilde{\mathcal{O}}(\ell n^3)$; dimension $m =2 n\lceil \log q\rceil$; Gaussian parameter $\sigma = \Omega(\sqrt{n\log q}\log n)$; infinity norm bounds $\beta = \sigma\omega({\log m})$ and $B = \sqrt{n} \omega(\log n)$. Let $\chi$ be a $B$-bounded distribution.
|
|
Choose a hash function $H:\{0,1\}^*
|
|
\rightarrow \{1,2,3\}^t$ for some $t = \omega(\log n)$,
|
|
which will be modeled as a random oracle in the security analysis.
|
|
Then, do the following. \smallskip \smallskip
|
|
\begin{itemize}
|
|
\item[1.] Generate a key pair for the signature of Section \ref{desc-sig-protoc} for signing single-block messages. Namely, run $\TrapGen(1^n,1^m,q)$ to get~$\mathbf{A} \in
|
|
\ZZ_q^{n \times m}$ and a short basis $\mathbf{T}_{\mathbf{A}}$ of
|
|
$\Lambda_q^{\perp}(\mathbf{A})$, which allows computing short vectors in $\Lambda_q^{\perp}(\mathbf{A})$ with Gaussian parameter $\sigma$.
|
|
Next, choose matrices
|
|
$\mathbf{A}_0,\mathbf{A}_1,\ldots,\mathbf{A}_{\ell},\mathbf{D} \sample (\ZZ_q^{n \times m})$, $ \mathbf{D}_0,\mathbf{D}_1 \sample (\ZZ_q^{2n \times 2m})$ and a vector $\mathbf{u} \sample (\ZZ_q^n)$.
|
|
\item[2.] Choose an additional random matrix $\mathbf{F} \sample (\ZZ_q^{4n \times 4m})$ uniformly. Looking ahead, this matrix will be used to ensure security against framing attacks.
|
|
\item[3.]
|
|
Generate a master key pair for the Gentry-Peikert-Vaikuntanathan IBE scheme
|
|
in its multi-bit variant. This key pair consists of a statistically uniform matrix
|
|
$\mathbf{B} \in \ZZ_q^{n \times m}$ and a short basis $\mathbf{T}_{\mathbf{B}} \in
|
|
\ZZ^{m \times m}$ of $\Lambda_q^{\perp}(\mathbf{B})$. This basis will allow us to compute GPV private keys with a
|
|
Gaussian parameter $\sigma_{\mathrm{GPV}} \geq \| \widetilde{\mathbf{T}}_{\mathbf{B}} \| \cdot
|
|
\sqrt{\log m}$.
|
|
\item[4.] Choose a one-time signature scheme $\Pi^\mathrm{OTS}=(\mathcal{G},\mathcal{S},\mathcal{V})$ and a hash function $H_0:\{0,1\}^* \rightarrow \ZZ_q^{ n \times 2m}$,
|
|
that will be modeled as random oracles.
|
|
\end{itemize}
|
|
The group public key is defined
|
|
as $$\mathcal{Y}\coloneqq \big( \mathbf{A}, ~
|
|
\{\mathbf{A}_j \}_{j=0}^{\ell},~\mathbf{B}, ~\mathbf{D},~ \mathbf{D}_0,~\mathbf{D}_1,~\mathbf{F}, ~\mathbf{u} , ~\Pi^\mathrm{OTS}, ~ H,~H_0 \big).$$
|
|
The opening authority's private key is $\mathcal{S}_{\OA}\coloneqq
|
|
\mathbf{T}_{\mathbf{B}} $ and the private key of the group manager consists of $\mathcal{S}_{\GM}\coloneqq \mathbf{T}_{\mathbf{A}}$. The algorithm outputs
|
|
$\big( \mathcal{Y},\mathcal{S}_{\GM},\mathcal{S}_{\OA} \big)$.
|
|
|
|
\bigskip
|
|
|
|
\item[\textsf{Join}$^{(\mathsf{GM},\mathcal{U}_i)}$:] the group manager $\GM$ and the prospective user $\mathcal{U}_i$ run the following interactive protocol: \smallskip
|
|
$\left\langle \mathsf{J}_{\user}(\lambda,\mathcal{Y}),\mathsf{J}_{\GM}(\lambda,St,\mathcal{Y},\mathcal{S}_{\GM}) \right\rangle$
|
|
\begin{itemize}
|
|
\item[1.] $\mathcal{U}_i$ samples a discrete Gaussian vector $\mathbf{z}_{i} \leftarrow D_{\ZZ^{4m},\sigma}$ and computes $\mathbf{v}_{i} = \mathbf{F} \cdot \mathbf{z}_{i} \in \ZZ_q^{ 4n}$.
|
|
He sends the vector $\mathbf{v}_{i} \in \ZZ_q^{4n}$, whose binary representation $\mathsf{bin}(\mathbf{v}_i)$ consists of $4n\lceil\log q\rceil = 2m$ bits, together with an ordinary digital signature $sig_i = \mathrm{Sign}_{\usk[i]}(\mathbf{v}_i)$ to $\GM$.
|
|
\item[2.] $\mathsf{J}_{\GM}$ verifies that $\mathbf{v}_i$ was not previously used by a registered user and that
|
|
$sig_i$ is a valid signature on $ \mathbf{v}_i $ w.r.t. $\upk[i]$. It aborts if this is not the case. Otherwise, $\GM$ chooses a fresh $\ell$-bit identifier $\mathsf{id}_i=\mathsf{id}_i[1]\ldots \mathsf{id}_i[\ell]
|
|
\in \{0,1\}^{\ell}$ and
|
|
uses $\mathcal{S}_{\GM}=\mathbf{T}_{\mathbf{A}}$ to certify
|
|
$\mathcal{U}_i$ as a new group member. To this end, $\GM$
|
|
defines the matrix
|
|
\begin{eqnarray} \label{matr}
|
|
\mathbf{A}_{\mathsf{id}_i}= \left[ \begin{array}{c|c} \mathbf{A} ~& ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \mathsf{id}_i[j] \mathbf{A}_j
|
|
\end{array} \right] \in \ZZ_q^{ n \times 2m}.
|
|
\end{eqnarray}
|
|
Then, $\GM$ runs $\mathbf{T}_{\mathsf{id}_i}' \leftarrow
|
|
\ExtBasis(\mathbf{A}_{\mathsf{id}_i},\mathbf{T}_{\mathbf{A}})$ to obtain a short delegated basis
|
|
$\mathbf{T}_{\mathsf{id}_i}'$ of $\Lambda_q^{\perp}(\mathbf{A}_{\mathsf{id}_i}) \in \ZZ^{ 2m \times 2m }$.
|
|
Finally, $\GM$ samples a short vector $\mathbf{s}_i \sample D_{\ZZ^{2m},\sigma }$ and uses the obtained delegated basis $\mathbf{T}_{\mathsf{id}_i}' $ to compute a short vector
|
|
$\mathbf{d}_i = \begin{bmatrix} \mathbf{d}_{i,1} \\ \hline \mathbf{d}_{i,2} \end{bmatrix} \in \ZZ^{2m}$ such that
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A}_{\mathsf{id}_i} \cdot \mathbf{d}_i &=& \left[ \begin{array}{c|c} \mathbf{A} ~& ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \mathsf{id}_i[j] \mathbf{A}_j
|
|
\end{array} \right] \cdot \mathbf{d}_i\\
|
|
\label{rel-cert} &=& \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}_i) + \mathbf{D}_1 \cdot \mathbf{s}_i \bigr) \bmod q. \quad
|
|
\end{eqnarray}
|
|
The triple $(\mathsf{id}_i,\mathbf{d}_i,\mathbf{s}_i)$ is sent to $\mathcal{U}_i$. Then,
|
|
$\mathsf{J}_{\user}$ verifies that the received $(\mathsf{id}_i,\mathbf{d}_i,\mathbf{s}_i)$ satisfies (\ref{rel-cert}) and that
|
|
$\| \mathbf{d}_i \|_\infty \leq \beta$, $\| \mathbf{s}_i \|_\infty \leq \beta $. If these conditions are not satisfied, $\mathsf{J}_{\user}$ aborts.
|
|
Otherwise,
|
|
$\mathsf{J}_{\user}$ defines the membership
|
|
certificate as
|
|
$ \crt_{i }=( \mathsf{id}_i, \mathbf{d}_i,\mathbf{s}_i )$.
|
|
The membership secret $\scr_{i }$ is defined to be $\scr_i=\mathbf{z}_i \in \ZZ^{4m}$. $\mathsf{J}_{\GM}$ stores
|
|
$\transcript_i=(\mathbf{v}_i, \crt_i, i,\mathsf{upk}[i],sig_i)$
|
|
in the database $St_{trans}$ of joining transcripts. \smallskip \smallskip
|
|
\end{itemize}
|
|
|
|
|
|
|
|
\item[\textsf{Sign}$(\mathcal{Y},\crt_i,\scr_i ,M)$:] To sign $M \in
|
|
\{0,1\}^*$ using $\crt_i=(\mathsf{id}_i,\mathbf{d}_i,\mathbf{s}_i)$, where $\mathbf{d}_i=[ \mathbf{d}_{i,1}^T \mid \mathbf{d}_{i,2}^T ]^T \in \ZZ^{2m}$ and $\mathbf{s}_i \in \ZZ^{2m}$, as
|
|
well as the membership secret $\scr_i=\mathbf{z}_i \in \ZZ^{4m}$, the group
|
|
member $\mathcal{U}_i$ generates a one-time signature key pair $(\mathsf{VK},\mathsf{SK}) \leftarrow \mathcal{G}(n)$ and conducts the following steps. \smallskip
|
|
|
|
\begin{itemize}
|
|
\item[1.] Compute $\mathbf{G}_0=H_0(\mathsf{VK}) \in \ZZ_q^{ n \times 2m}$ and use it as an IBE public key to encrypt
|
|
$\textsf{bin}(\mathbf{v}_i) \in \{0,1\}^{2m}$, where $\mathbf{v}_i=\mathbf{F} \cdot \mathbf{z}_i \in \ZZ_q^{4n}$ is the syndrome of
|
|
$\scr_i=\mathbf{z}_i \in \mathbb{Z}^{4m}$ for the matrix $\mathbf{F}$. Namely, compute $ \mathbf{c}_{\mathbf{v}_i} \in \ZZ_q^m \times \ZZ_q^{2m}$ as
|
|
\begin{eqnarray} \label{enc1}
|
|
\mathbf{c}_{\mathbf{v}_i}=(\mathbf{c}_1,\mathbf{c}_2) &=& \big( \mathbf{B}^T \cdot \mathbf{e}_0 + \mathbf{x}_1 ,~ \mathbf{G}_0^T \cdot \mathbf{e}_0 + \mathbf{x}_2 + \textsf{bin}(\mathbf{v}_i) \cdot \lfloor q/2 \rfloor \big) \qquad
|
|
\end{eqnarray}
|
|
for randomly chosen $\mathbf{e}_0 \sample \chi^n$, $\mathbf{x}_1 \sample \chi^m, \mathbf{x}_2 \sample \chi^{2m} $.
|
|
Notice that, as in the construction of \cite{LNW15}, the columns of $\mathbf{G}_0$ can be interpreted as public keys for the multi-bit version
|
|
of the dual Regev encryption scheme.
|
|
\item[2.] Run the protocol in Section~\ref{subsection:zk-for-group-signature} to prove the knowledge of $\mathsf{id}_i
|
|
\in \{0,1\}^{\ell}$,
|
|
vectors $\mathbf{s}_i \in \ZZ^{2m}, \mathbf{d}_{i,1},\mathbf{d}_{i,2} \in \ZZ^{m},\mathbf{z}_i \in \ZZ^{4m}$ with infinity norm bound $\beta $; $\mathbf{e}_0 \in \ZZ^n$, $\mathbf{x}_1 \in \ZZ^m, \mathbf{x}_2 \in \ZZ^{2m} $ with infinity norm bound $B$
|
|
and $\textsf{bin}(\mathbf{v}_i) \in \{0,1\}^{2m}, \mathbf{w}_{i} \in \{0,1\}^m$, that satisfy
|
|
\eqref{enc1} as well as
|
|
\begin{eqnarray} \label{rel-deux}
|
|
\mathbf{A} \cdot \mathbf{d}_{i,1} + \mathbf{A}_0 \cdot \mathbf{d}_{i,2} + \sum_{j=1}^{\ell} ( \mathsf{id}_i[j] \cdot \mathbf{d}_{i,2}) \cdot \mathbf{A}_j
|
|
- \mathbf{D} \cdot \mathbf{w}_i = \mathbf{u} \in \ZZ_q^n
|
|
\end{eqnarray}
|
|
and
|
|
|
|
\begin{eqnarray} \label{eq:rel-3}
|
|
\left\{
|
|
\begin{array}{l}
|
|
\mathbf{H}_{2n \times m} \cdot \mathbf{w}_{i} = \mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}_i) + \mathbf{D}_1 \cdot \mathbf{s}_i \in \ZZ_q^{2n} \\
|
|
\mathbf{F} \cdot \mathbf{z}_i = \mathbf{H}_{4n \times 2m} \cdot \textsf{bin}(\mathbf{v}_i) \in \ZZ_q^{4n}.
|
|
\end{array}
|
|
\right.
|
|
\end{eqnarray}
|
|
|
|
The protocol is repeated $t = \omega(\log n)$ times in parallel to achieve negligible soundness error, and then made non-interactive using the Fiat-Shamir
|
|
heuristic~\cite{FS86} as a triple $\pi_K=(
|
|
\{\mathsf{Comm}_{K,j}\}_{j=1}^t,\mathsf{Chall}_K,\{\mathsf{Resp}_{K,j}\}_{j=1}^t)$,
|
|
where $\mathsf{Chall}_K = H(M, \vk, \mathbf{c}_{\mathbf{v}_i},
|
|
\{ \mathsf{Comm}_{K,j}\}_{j=1}^t) \in \{1,2,3\}^t$
|
|
|
|
\item[3.] Compute a one-time signature $sig=\mathcal{S}(\mathsf{SK},(\mathbf{c}_{\mathbf{v}_i} , \pi_K))$. \smallskip
|
|
|
|
|
|
\end{itemize}
|
|
Output the signature that consists of
|
|
|
|
\begin{equation} \label{eq:sig-final} \Sigma=\big( \mathsf{VK} ,\mathbf{c}_{\mathbf{v}_i}, \pi_K,sig \big).
|
|
\end{equation}
|
|
|
|
\smallskip
|
|
|
|
\item[\textsf{Verify}$(\mathcal{Y},M,\Sigma)$:] Parse the signature $\Sigma$ as in
|
|
(\ref{eq:sig-final}). Then, return $1$ if and only if:
|
|
(i) $\mathcal{V}(\mathsf{VK},(\mathbf{c}_{\mathbf{v}_i},\mathbf{c}_{\mathbf{s}_i},\mathbf{c}_{\mathsf{id}},\pi_K),sig)=1$;
|
|
(ii) The proof $\pi_K$ properly verifies. \smallskip %Otherwise, return $0$. \smallskip
|
|
|
|
\item[\textsf{Open}$(\mathcal{Y},\mathcal{S}_{\OA},M,\Sigma)$:] Parse~$\mathcal{S}_{\OA}$ as~$
|
|
\mathbf{T}_{\mathbf{B}} \in \ZZ^{m \times m}$ and $\Sigma$ as
|
|
in~(\ref{eq:sig-final}). \smallskip
|
|
\begin{itemize}
|
|
\item[1.]
|
|
Compute $\mathbf{G}_0=H_0(\mathsf{VK}) \in \ZZ_q^{n \times 2m}$. Then, using $\mathbf{T}_{\mathbf{B}}$
|
|
to compute a small-norm matrix
|
|
$\mathbf{E}_{0,\mathsf{VK}} \in \ZZ^{m \times 2m }$ such that $ \mathbf{B} \cdot \mathbf{E}_{0,\mathsf{VK}} = \mathbf{G}_0 \bmod q $.
|
|
\item[2.] Using $\mathbf{E}_{0,\mathsf{VK}}$, decrypt $\mathbf{c}_{\mathbf{v}_i}$ to obtain a string $\textsf{bin}(\mathbf{v} ) \in \{0,1\}^{2m}$
|
|
(i.e., by computing $\lfloor (\mathbf{c}_2 - \mathbf{E}_{0,\mathsf{VK}}^T \cdot \mathbf{c}_1) / (q/2) \rceil$). \smallskip
|
|
\item[3.] Determine whether the $\textsf{bin}(\mathbf{v} ) \in \{0,1\}^{2m} $ obtained at step~2 corresponds to a vector $\mathbf{v} = \mathbf{H}_{4n \times 2m} \cdot \textsf{bin}(\mathbf{v} ) \bmod q$ that appears in a record $\transcript_i=(\mathbf{v} , \crt_i, i,\mathsf{upk}[i],sig_i)$ of the database $St_{trans}$ for some $i$. If so,
|
|
output the corresponding $i$ (and, optionally, $\mathsf{upk}[i]$). Otherwise, output $\perp$.
|
|
\end{itemize}
|
|
\end{description}
|
|
|
|
We remark that the scheme readily extends to provide a mechanism whereby the opening authority can efficiently prove that signatures were correctly opened at each opening operation.
|
|
The difference between the dynamic group signature models suggested by Kiayias and Yung \cite{KY06} and Bellare \textit{et al.} \cite{BSZ05} is that, in the latter, the opening authority
|
|
($\mathsf{OA}$) must be able to convince a judge that the $\mathsf{Open}$ algorithm was run correctly.
|
|
Here, such a mechanism can be realized using the techniques of public-key encryption with non-interactive opening \cite{DHKT08}. Namely, since
|
|
$\textsf{bin}(\mathbf{v}_i)$ is encrypted using an IBE scheme for the identity $\vk$, the $\mathsf{OA}$ can simply reveal the decryption matrix $\mathbf{E}_{0,\mathsf{VK}} $,
|
|
that satisfies $\mathbf{B} \cdot \mathbf{E}_{0,\vk} = \mathbf{G}_0 \bmod q$ (which corresponds to the verification of a GPV signature) and allows the verifier to perform step 2 of the opening
|
|
algorithm himself. The resulting construction is easily seen to satisfy the notion of opening soundness of Sakai \textit{et al.} \cite{SSE+12}.
|
|
|
|
\subsection{Efficiency and Correctness}
|
|
\textsc{Efficiency.} The given dynamic group signature scheme can be implemented in polynomial time. The group public key has total bit-size $\mathcal{O}(\ell n m \log q) = \widetilde{\mathcal{O}}(\lambda^2)\cdot \log N_\textsf{gs}$. The secret signing key of each user consists of a small constant number of low-norm vectors, and has bit-size $\widetilde{\mathcal{O}}(\lambda)$.
|
|
|
|
The size of each group signature is largely dominated by that of the non-interactive argument $\pi_K$, which is obtained from the Stern-like protocol of Section~\ref{subsection:zk-for-group-signature}. Each round of the protocol has communication cost $\widetilde{\mathcal{O}}(m \cdot \log q) \cdot \log N_\textsf{gs}$. Thus, the bit-size of $\pi_K$ is $t\hspace*{-1pt}\cdot\hspace*{-1pt} \widetilde{\mathcal{O}}(m \hspace*{-1pt}\cdot\hspace*{-1pt} \log q) \hspace*{-1pt}\cdot\hspace*{-1pt} \log N_\textsf{gs} = \widetilde{\mathcal{O}}(\lambda)\hspace*{-1pt}\cdot \hspace*{-1pt}\log N_\textsf{gs}$. This is also the asymptotic bound on the size of the group signature.
|
|
|
|
|
|
\smallskip
|
|
\textsc{Correctness.} The correctness of algorithm \textsf{Verify}$(\mathcal{Y},M,\Sigma)$ follows from the facts that every certified group member is able to compute valid witness vectors satisfying equations~(\ref{enc1}), (\ref{rel-deux}) and (\ref{eq:rel-3}), and that the underlying argument system is perfectly complete. Moreover, the scheme parameters are chosen so that the GPV IBE~\cite{GPV08} is correct, which implies that algorithm \textsf{Open}$(\mathcal{Y},\mathcal{S}_{\OA},M,\Sigma)$ is also correct.
|
|
|
|
|
|
\subsection{Security Analysis}
|
|
|
|
Due to the fact that the number of public matrices $\{\mathbf{A}_j\}_{j=0}^\ell$ is only logarithmic in ${N_{\mathsf{gs}}}=2^\ell$ instead of being linear in the security parameter $\lambda$,
|
|
the proof of security against misidentification attacks (as defined in \cref{sse:gs-sec-notions}) cannot rely on the security of our signature scheme in a modular manner.
|
|
The reason is that, at each run of the $\mathsf{Join}$ protocol, the group manager maintains a state and, instead of choosing the $\ell$-bit identifier $\mathsf{id}$ uniformly in
|
|
$\{0,1\}^{\ell}$, it chooses an identifier that has not been used yet. Since $\ell \ll \lambda$ (given that ${N_{\mathsf{gs}}}=2^\ell$ is polynomial in $\lambda$), we thus have
|
|
to prove security from scratch. However, the strategy of the reduction is exactly the same as in the security proof of the signature scheme.
|
|
|
|
|
|
\begin{theorem} \label{traceability-thm}
|
|
The scheme is secure against misidentification attacks under the $\SIS_{n,2m,q,\beta'}$ assumption, for $\beta' \hspace*{-1pt}=\hspace*{-1pt} \mathcal{O}(\ell \sigma^2 m^{3/2})$.
|
|
\end{theorem}
|
|
|
|
\begin{proof}
|
|
We prove that any adversary $\adv$ with non-negligible success probability $\varepsilon$ implies an algorithm $\bdv$ solving the \textsf{SIS} problem
|
|
in the random oracle model.
|
|
|
|
Let $\adv$ be such a $\ppt$ adversary.
|
|
We then build a $\ppt$ reduction~$\bdv$ that uses the adversary~$\adv$ to
|
|
solve~$\SIS_{n,2m,q,\beta'}$: specifically, $\bdv$ takes as input~$\bar{\mathbf{A}} = \begin{bmatrix} \bar{\mathbf{A}}_1 | \bar{\mathbf{A}}_2 \end{bmatrix} \in
|
|
\Zq^{n \times 2m}$, where $\bar{\mathbf{A}}_1,\bar{\mathbf{A}}_2 \in \Zq^{n \times m}$, and finds $\mathbf{w} \in
|
|
\Lambda_q^{\perp}(\bar{\mathbf{A}})$ with~$0 < \|\mathbf{w}\| \leq \beta'$.
|
|
\medskip
|
|
|
|
|
|
\noindent \textbf{Initialization.} Algorithm~$\bdv$ first chooses a random $coin \sample
|
|
U(\{0,1,2\})$ as a guess for the kind of misidentification attack that $\adv$ will mount. Also, $\bdv$
|
|
chooses a random $\ell$-bit string $\mathsf{id}^\dagger \sample (\{0,1\}^\ell)$.
|
|
In
|
|
addition, $\bdv$
|
|
samples~$i^\star
|
|
\sample ([1,Q_a])$. \\
|
|
\indent
|
|
Looking ahead, $coin=0$ corresponds to the case where, after repeated executions of $\adv$, the knowledge extractor of the proof system
|
|
reveals witnesses containing a new identifier $\mathsf{id}^\star \in \{0,1\}^\ell$ that does not belong to any user in $U^a$.
|
|
In this case, $\bdv$ will be able to exploit $\adv$'s forgery when $\mathsf{id}^\star=\mathsf{id}^\dagger$.
|
|
The case $coin=1$ corresponds to $\bdv$'s expectation that the knowledge extractor will obtain the identifier $ \mathsf{id}^\star = \mathsf{id}^\dagger$ of a group member in
|
|
$ U^a$ (i.e., a group member that was legitimately introduced at the $i^\star$-th $\mathcal{Q}_{\ajoin}$-query, for some $i^\star \in \{1,\ldots,Q_a\}$, where the identifier
|
|
$\mathsf{id}^\dagger$ is used by $\mathcal{Q}_{\ajoin}$),
|
|
but $\textsf{bin}( \mathbf{v}^\star ) \in \{0,1\}^{2m}$ (which is encrypted in in $\mathbf{c}_{\mathbf{v}_i}^\star$ as part of the forgery $\Sigma^\star$) and the extracted $\mathbf{s}^\star \in \ZZ^{2m}$ are such that $ \bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}^\star ) + \mathbf{D}_1 \cdot \mathbf{s}^\star \bigr) \in \{0,1\}^m $
|
|
does not match
|
|
the string $ \bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}_{i^\star} ) + \mathbf{D}_1 \cdot \mathbf{s}_{i^\star} \bigr) \in \{0,1\}^{2m} $ for which
|
|
user $i^\star$ obtained a membership certificate at the $i^\star$-th $\mathcal{Q}_{\ajoin}$-query. When $coin=1$, the choice of $i^\star$ corresponds to a guess that the knowledge
|
|
extractor will reveal an $\ell$-bit identifier that coincides with the identifier $\mathsf{id}^\dagger$ assigned to the user introduced at the $i^\star$-th $\mathcal{Q}_{\ajoin}$-query.
|
|
The last case $coin=2$ corresponds to $\bdv$'s expectation that decrypting $\mathbf{c}_{\mathbf{v}_i}^\star$ (which is part of $\Sigma^\star$) and running
|
|
the knowledge extractor on $\adv$ will uncover vectors $\textsf{bin}( \mathbf{v}^\star ) \in \{0,1\}^{2m}$, $\mathbf{w}^\star \in \{0,1\}^m$ and $\mathbf{s}^\star \in \ZZ^{2m}$
|
|
such that $\mathbf{w}^\star= \textsf{bin}(\mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}^\star) + \mathbf{D}_1 \cdot \mathbf{s}^\star )$ and
|
|
\begin{eqnarray} \label{collide}
|
|
\bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}^\star ) + \mathbf{D}_1 \cdot \mathbf{s}^\star \bigr) = \bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}_{i^\star} ) + \mathbf{D}_1 \cdot \mathbf{s}_{i^\star} \bigr)
|
|
\end{eqnarray}
|
|
but $(\textsf{bin}( \mathbf{v}^\star ), \mathbf{s}^\star) \neq ( \textsf{bin}( \mathbf{v}_{i^\star} ), \mathbf{s}_{i^\star} ) $, where $ \mathbf{v}_{i^\star} \in \Zq^{4n}$ and $\mathbf{s}_{i^\star} \in \ZZ^{2m}$ are the vectors
|
|
involved in the $i^\star$-th $\mathcal{Q}_{\ajoin}$-query.
|
|
\\
|
|
\indent
|
|
Depending on $coin \in \{0,1,2\}$, the group public key $\mathcal{Y}$ is
|
|
generated using different methods. \smallskip
|
|
|
|
\noindent $\bullet$ If $coin=0$, algorithm~$\bdv$ first randomly chooses $\mathsf{id}^\dagger \sample (\{0,1\}^\ell)$ as a guess for the $\ell$-bit string
|
|
that will be revealed by the knowledge extractor of the proof system after repeated executions of the adversary $\adv$.
|
|
Then, it runs
|
|
$\TrapGen(1^n,1^m,q)$ to obtain $\mathbf{C} \in \Zq^{n \times m}$ and a
|
|
basis $\mathbf{T}_{\mathbf{C}}$ of~$\Lambda_q^{\perp}(\mathbf{C})$ with
|
|
$\|\widetilde{\mathbf{T}_{\mathbf{C}}}\| \leq \bigO(\sqrt{n \log q})$. Then,
|
|
it chooses~$\ell+2$ matrices~$ \mathbf{Q}_0,\ldots,\mathbf{Q}_{\ell},\mathbf{Q}_D \in \ZZ^{m \times m}$,
|
|
each matrix having its columns sampled independently from~$D_{\ZZ^m,\sigma}$. Then, $\bdv$ defines the matrices $\{ \mathbf{A}_i\}_{i=0}^{\ell}$ as
|
|
\begin{eqnarray*}
|
|
\left\{
|
|
\begin{array}{ll}
|
|
\mathbf{A}_0 = \bar{\mathbf{A}}_1 \cdot \mathbf{Q}_0 + (\sum_{i=1}^{\ell} {\mathsf{id}^\dagger[i]}) \cdot
|
|
\mathbf{C} \\
|
|
\mathbf{A}_j = \bar{\mathbf{A}}_1 \cdot \mathbf{Q}_i + (-1)^{\mathsf{id}^{\dagger}[j]} \cdot
|
|
\mathbf{C}, \quad \text{ for } j \in
|
|
[1,\ell]. \\
|
|
\mathbf{D} = \bar{\mathbf{A}}_1 \cdot \mathbf{Q}_D
|
|
\end{array}
|
|
\right.
|
|
\end{eqnarray*}
|
|
It also defines $\mathbf{A}=\bar{\mathbf{A}}_1$.
|
|
Next, it samples a vector $\mathbf{e}_u \sample D_{\ZZ,\sigma}^m$ and computes a syndrome $\mathbf{u} = \bar{\mathbf{A}}_1 \cdot \mathbf{e}_u \in \Zq^n$. It picks $\mathbf{D}_0,\mathbf{D}_1
|
|
\sample (\Zq^{2n \times 2m})$ at random and also faithfully generates the GPV master key pair $(\mathbf{B},\mathbf{T}_{\mathbf{B}})$ as in Step~3 of the real setup algorithm. The group
|
|
public key $\mathcal{Y}=\big(\mathbf{A},\{\mathbf{A}_j \}_{j=0}^{\ell}, \mathbf{B}, \mathbf{D},\mathbf{D}_0,\mathbf{D}_1,\mathbf{F}, \mathbf{u},\mathcal{OTS},H,H_0 \big)$
|
|
is finally given to~$\adv$. \\
|
|
\indent Note that, for each $\mathsf{id} \neq \mathsf{id}^\dagger$, we have
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A}_{\mathsf{id}} &=& \left[
|
|
\begin{array}{c|c} \bar{\mathbf{A}}_1 ~&~ \mathbf{A}_0 +
|
|
\sum_{i=1}^\ell \mathsf{id}[i] \mathbf{A}_i
|
|
\end{array} \right] \\ \nonumber & = & \left[
|
|
\begin{array}{c|c} \bar{\mathbf{A}}_1 ~&~ \bar{\mathbf{A}}_1 \cdot (\mathbf{Q}_0 +
|
|
\sum_{i=1}^{\ell} \mathsf{id}[i] \mathbf{Q}_i) + (
|
|
\sum_{i=1}^{\ell} \mathsf{id}^\dagger [i] +(-1)^{\mathsf{id}^\dagger[i]} \mathsf{id}[i])\cdot \mathbf{C}
|
|
\end{array} \right] \\ \label{sim-matr} &=&
|
|
\left[
|
|
\begin{array}{c|c} \bar{\mathbf{A}}_1 ~&~ \bar{\mathbf{A}}_1 + h_{\mathsf{id}} \cdot \mathbf{C}
|
|
\end{array} \right]
|
|
\end{eqnarray}
|
|
where $h_{\mathsf{id}} \in [1,\ell]$ denotes the Hamming distance between
|
|
the identifiers $\mathsf{id}$ and $\mathsf{id}^\dagger$. Since $q>\ell$, we have
|
|
$h_{\mathsf{id}_j} \neq 0 \bmod q$ whenever $\mathsf{id}_j \neq \mathsf{id}^\dagger$, so
|
|
that algorithm $\bdv$ is able to compute (see~\cite[Se.~4.2]{ABB10},
|
|
using the basis~$\mathbf{T}_{\mathbf{C}}$ of~$\Lambda_q^{\perp}(\mathbf{C})$ and
|
|
the refined $\GPVSample$ of Lemma~\ref{le:GPV}) a basis
|
|
$\mathbf{T}_{\mathsf{id}}$ of $\Lambda_q^{\perp}(\mathbf{A}_{\mathsf{id}})$
|
|
with~$\|\widetilde{\mathbf{T}_{\mathsf{id}}}\| \leq \Omega(\sqrt{n\log
|
|
q\log n})$. In contrast,
|
|
algorithm~$\bdv$ lacks a trapdoor for $\mathbf{A}_{\mathsf{id}^\dagger}$ as the
|
|
latter only depends on $\mathbf{A}$ and $\{\mathbf{Q}_k\}_{k=0}^{\ell}$.
|
|
Observe that, since the columns of the matrices~$\{\mathbf{Q}_k\}_{k=0}^\ell$ are sampled
|
|
from~$D_{\ZZ^m,\sigma}$, the
|
|
matrices~$ \mathbf{A}_0,\ldots,\mathbf{A}_{\ell}$ are within
|
|
statistical distance~$2^{-\Omega(m)}$ of~$U(\Zq^{n \times m})$.
|
|
\smallskip
|
|
|
|
|
|
\noindent $\bullet$ If $coin=1$, algorithm~$\bdv$ sets up $\mathcal{Y}$ by defining
|
|
$\mathbf{D}=\bar{\mathbf{A}}$. Initially, $\bdv$
|
|
chooses $Q_a-1$ distinct strings $\mathsf{id}_1, \ldots,\mathsf{id}_{i^\star-1}, \mathsf{id}_{i^\star+1},\ldots,\mathsf{id}_{Q_a} \in \{0,1\}^\ell$ such that, for each $i \in [1,Q_a] \backslash \{i^\star\}$, $\mathsf{id}_i$ will be embedded in the membership certificate
|
|
returned in the $i$-th $\mathcal{Q}_{\ajoin}$-query. Let also $\mathsf{id}^\dagger=\mathsf{id}_{i^\star}$ be the $\ell$-bit identifier
|
|
that will be used in the $i^\star$-th query.
|
|
The reduction $\bdv$ picks random $h_0,h_1,\ldots,h_\ell \in \Zq$ under the constraints
|
|
\begin{eqnarray*}
|
|
h_{\mathsf{id}^\dagger} = h_0 + \sum_{j=1}^\ell \mathsf{id}^\dagger[j] \cdot h_j &=& 0 \bmod q \\
|
|
h_{\mathsf{id}_i} = h_0 + \sum_{j=1}^\ell \mathsf{id}_i[j] \cdot h_j & \neq & 0 \bmod q \qquad \qquad i \in \{1,\ldots,Q_a\} \setminus \{i^\dagger\}
|
|
\end{eqnarray*}
|
|
Next, $\bdv$ runs $(\mathbf{C},\mathbf{T}_{\mathbf{C}}) \leftarrow \mathsf{TrapGen}(1^n,1^m,q)$, $(\mathbf{D}_1,\mathbf{T}_{\mathbf{D}_1}) \leftarrow \mathsf{TrapGen}(1^{2n},1^{2m},q)$ so as to obtain statistically random matrices $\mathbf{C} \in \Zq^{n \times m}$, $ \mathbf{D}_1 \in \Zq^{2n \times 2m}$ together with
|
|
trapdoors $\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m} $, $\mathbf{T}_{\mathbf{D}_1} \in \ZZ^{2m \times 2m}$ consisting of short bases of $\Lambda_q^{\perp}(\mathbf{C})$ and $\Lambda_q^{\perp}(\mathbf{D}_1)$, respectively. Then,
|
|
$\bdv$
|
|
picks a random $\mathbf{D}_0 \sample (\Zq^{2n \times 2m})$ and re-randomizes $\mathbf{D}=\bar{\mathbf{A}}_1 \in \Zq^{n \times m}$ using Gaussian matrices
|
|
$\mathbf{S},\mathbf{S}_0,\mathbf{S}_1,\ldots,\mathbf{S}_{\ell} \sample \ZZ^{m \times m}$ whose columns are sampled from the distribution $D_{\ZZ^m,\sigma}$.
|
|
Namely, from $\mathbf{D} =\bar{\mathbf{A}}_1 $, $\bdv$
|
|
defines
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A} &=& \bar{\mathbf{A}}_1 \cdot \mathbf{S} \\ \label{setup-sig2}
|
|
\mathbf{A}_0 &=& \bar{\mathbf{A}}_1 \cdot \mathbf{S}_0 + h_0 \cdot \mathbf{C} \\ \nonumber
|
|
\mathbf{A}_j &=& \bar{\mathbf{A}}_1 \cdot \mathbf{S}_j + h_j \cdot \mathbf{C} \qquad \qquad \forall j \in \{1,\ldots,\ell\} .
|
|
\end{eqnarray}
|
|
As part of the generation of
|
|
$\mathcal{Y}$, the vector $\mathbf{u} \in \Zq^n$ is obtained by picking short discrete Gaussian vectors
|
|
$ \mathbf{d}_{i^\star,1}, \mathbf{d}_{i^\star,2} \sample D_{\ZZ^m,\sigma} $
|
|
and computing
|
|
\begin{eqnarray} \label{def-u}
|
|
\mathbf{u} = [ \mathbf{A} ~\mid ~ \mathbf{A}_0 +
|
|
\sum_{j=1}^\ell \mathsf{id}^\dagger[j] \mathbf{A}_j
|
|
] \cdot
|
|
\begin{bmatrix}
|
|
\mathbf{d}_{i^\star,1} \\ \hline \mathbf{d}_{i^\star,2}
|
|
\end{bmatrix}
|
|
- \mathbf{D} \cdot \textsf{bin}(\mathbf{c}_M),
|
|
\end{eqnarray}
|
|
where
|
|
$\mathbf{c}_{M} \sample (\Zq^{2n})$ is a randomly chosen vector. Observe that, since $\mathbf{A}$ is statistically uniform over $\Zq^{n \times m}$ and $ \mathbf{d}_{i^\star,1}
|
|
\sample D_{\ZZ^m,\sigma}$, the distribution of
|
|
$\mathbf{u} $ is statistically close to $U(\Zq^n)$.
|
|
\medskip
|
|
|
|
\noindent $\bullet$ If $coin=2$, $\bdv$ picks $\bar{\mathbf{A}}' \sample (\Zq^{n \times 2m})$
|
|
and a random matrix $\mathbf{Q} \sample \ZZ^{2m \times 2m}$ whose columns are sampled from $D_{\ZZ^{2m},\sigma}$. These
|
|
are used to define $$\mathbf{D}_0= \begin{bmatrix} \bar{\mathbf{A}} \\ \hline \bar{\mathbf{A}}' \end{bmatrix} \in \Zq^{2n \times 2m} ,$$
|
|
and $\mathbf{D}_1=\mathbf{D}_0 \cdot \mathbf{Q} \bmod q$, which is statistically close to $U(\Zq^{2n \times 2m})$. All other components of $\mathcal{Y}$ are obtained by faithfully running the setup algorithm. \medskip
|
|
|
|
|
|
\indent For each value of $coin \in \{0,1,2\}$, the group public key
|
|
$$\mathcal{Y}=\big(\mathbf{A},\{\mathbf{A}_j \}_{j=0}^{\ell},\mathbf{B},\mathbf{D},\mathbf{D}_0,\mathbf{D}_1,\mathbf{F}, \mathbf{u},\mathcal{OTS},H,H_0 \big)$$ has a distribution which is statistically close to that of the real scheme and $\mathcal{Y}$ is given to $\adv$.
|
|
|
|
\medskip
|
|
|
|
|
|
\noindent \textbf{Queries.} The reduction~$\bdv$ starts interacting
|
|
with the adversary~$\adv$ and the way it handles~$\adv$'s queries to the $\mathcal{Q}_{\ajoin}$ oracle depends on the value of~$coin \in \{0,1,2\}$. \smallskip \smallskip
|
|
|
|
\noindent $\bullet$ If $coin=0$, answers $\mathcal{Q}_{\ajoin}$-queries as follows. When $\adv$ triggers an execution of the joining protocol, it chooses
|
|
a syndrome $\mathbf{v}_{i} \in \Zq^n$.
|
|
To answer the query, $\bdv$ chooses a fresh $\ell$-bit identifier $\mathsf{id}_i \in \{0,1\}^\ell$ such that
|
|
$\mathsf{id}_i \neq \mathsf{id}^\dagger$. If $\adv$ also provides a correct signature $sig_i$ such that
|
|
$\mathrm{Verify}_{\mathsf{upk}[i]}(\mathbf{v}_{i},sig_i)=1$, $\bdv$ samples $\mathbf{s}_i \sample D_{\ZZ^{2m},\sigma}$ and uses the trapdoor $\mathbf{T}_{\mathbf{C}}$ to compute a short vector
|
|
$\mathbf{d}_i=[\mathbf{d}_{i,1}^T ~|~\mathbf{d}_{i,2}^T]^T \in \ZZ^{2m}$ such that
|
|
\begin{eqnarray} \label{sim-cert}
|
|
\mathbf{A}_{\mathsf{id}_i} \cdot \begin{bmatrix} \mathbf{d}_{i,1} \\ \hline \mathbf{d}_{i,2} \end{bmatrix} = \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}_{i}) + \mathbf{D}_1 \cdot \mathbf{s}_i \bigr) ,
|
|
\end{eqnarray}
|
|
where $\mathbf{A}_{\mathsf{id}_i} \in \Zq^{n \times 2m}$ is the matrix in (\ref{sim-matr}). Note that $\bdv$ is able to compute such a vector using the $\mathsf{SampleRight}$
|
|
algorithm of \cite{ABB10} (since the Hamming distance $h_{\mathsf{id}_i}$ between $\mathsf{id}_i$ and $\mathsf{id}^\star$ is non-zero). The membership certificate
|
|
$\crt_i= (\mathsf{id}_i,\mathbf{d}_i,\mathbf{s}_i)$ is then returned to $\adv$.
|
|
\smallskip
|
|
|
|
\noindent $\bullet$ If $coin=1$, algorithm~$\bdv$ responds each $\mathcal{Q}_{\ajoin}$-query depending on the index $i \in \{1,\ldots,Q_a\}$ of the query. Specifically,
|
|
we distinguish two cases. \smallskip
|
|
|
|
\begin{itemize}
|
|
\item[-] If $i \neq i^\star$, $\bdv$ proceeds as in the previous case. Namely, it recalls the $\ell$-bit identifier $\mathsf{id}_i \in \{0,1\}^\ell$ (for which $\mathsf{id}_i \neq \mathsf{id}^\dagger$)
|
|
that was chosen in the setup phase and samples a short vector $\mathbf{s}_{i} \sample D_{\ZZ^{2m},\sigma}$. If $\adv$ also provides a correct signature $sig_i$ such that
|
|
$\mathrm{Verify}_{\mathsf{upk}[i]}(\mathbf{v}_{i},sig_i)=1$, generates a membership certificate $\crt_i$ for $\adv$ as in the case $coin=0$.
|
|
Note that
|
|
\begin{eqnarray} \nonumber
|
|
\mathbf{A}_{\mathsf{id}_i} &=& \left[
|
|
\begin{array}{c|c} \bar{\mathbf{A}} \cdot \mathbf{S} ~&~ \bar{\mathbf{A}} \cdot (\mathbf{S}_0 +
|
|
\sum_{j=1}^{\ell} \mathsf{id}_i[j] \mathbf{S}_j) + h_{\mathsf{id}_i} \mathbf{C}
|
|
\end{array} \right] \\ \label{sim-matr-coin1} &=&
|
|
\left[
|
|
\begin{array}{c|c} \bar{\mathbf{A}} \cdot \mathbf{S} ~&~ \bar{\mathbf{A}} + h_{\mathsf{id}_i} \cdot \mathbf{C}
|
|
\end{array} \right]
|
|
\end{eqnarray}
|
|
Since $h_{\mathsf{id}_i} \neq 0$, $\bdv$ can use the trapdoor
|
|
$\mathbf{T}_{\mathbf{C}} \in \ZZ^{m \times m}$ of $\Lambda_q^{\perp}(\mathbf{C})$ to compute a short vector $\mathbf{d}_i = [ \mathbf{d}_{i,1}^T ~|~\mathbf{d}_{i,2}^T ]^T \in \ZZ^{2m}$ such that
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\mathsf{id}_i} \cdot \begin{bmatrix} \mathbf{d}_{i,1} \\ \hline \mathbf{d}_{i,2} \end{bmatrix} = \mathbf{u} + \mathbf{D} \cdot \bit \bigl( \mathbf{D}_0 \cdot (\textsf{bin}(\mathbf{v}_{i}) + \mathbf{D}_1 \cdot \mathbf{s}_i \bigr) ,
|
|
\end{eqnarray*}
|
|
where $\mathbf{v}_{i} \in \Zq^{4n}$ is the syndrome chosen by $\adv$ at step 1 of the joining protocol.
|
|
\item[-] If $i = i^\star$, $\bdv$ undertakes to generate a membership certificate $\crt_{i^\star}$ for the $\ell$-bit identifier $\mathsf{id}^\dagger \in \{0,1\}^\ell$ that was
|
|
chosen at the outset of the game. To this end, $\bdv$ has to compute $\crt_{i^\star}$ without using the trapdoor $\mathbf{T}_{\mathbf{C}}$ since the matrix $\mathbf{A}_{\mathsf{id}^\dagger}$ does no longer
|
|
depend on $\mathbf{C}$ in (\ref{sim-matr-coin1} ). This can be done by recalling
|
|
the vector $\mathbf{d}_{i^\star,1},\mathbf{d}_{i^\star,2} \in \ZZ^m$ and $\mathbf{c}_M \in \Zq^{2n}$ that were used to define $\mathbf{u} \in \Zq^n$ in (\ref{def-u}) and using $\mathbf{T}_{\mathbf{D}_1}$. If $\adv$ provides a correct signature
|
|
$sig_{i^\star}$ such that
|
|
$\mathrm{Verify}_{\mathsf{upk}[i^\star]}(\mathbf{v}_{i^\star},sig_{i^\star})=1$,
|
|
$\bdv$ uses the trapdoor $\mathbf{T}_{\mathbf{D}_1}$ of $\Lambda_q^\perp (\mathbf{D}_1)$ to sample a short vector $\mathbf{s}_{i^\star} \in \ZZ^{2m}$ of $D_{\Lambda_q^{\mathbf{c}_{i^\star}}(\mathbf{D}_1),\sigma}$, where $\mathbf{c}_{i^\star} = \mathbf{c}_M - \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}_{i^\star}) \bmod q $,
|
|
satisfying
|
|
$$ \mathbf{D}_1 \cdot \mathbf{s}_{i^\star} = \mathbf{c}_M - \mathbf{D}_0 \cdot \textsf{bin}( \mathbf{v}_{i^\star}) ~\bmod q , $$
|
|
before returning $\crt_{i^\star}=(\mathsf{id}^\dagger,\mathbf{d}_{i^\star} =[ \mathbf{d}_{i^\star,1}^T \mid \mathbf{d}_{i^\star,2}^T]^T,\mathbf{s}_{i^\star})$
|
|
to $\adv$. From the definition of $\mathbf{u} \in \Zq^n$ (\ref{def-u}), it is easy to see that $\crt_{i^\star}=(\mathsf{id}^\dagger,\mathbf{d}_{i^\star} ,\mathbf{s}_{i^\star})$ forms a valid membership certificate for
|
|
any membership secret $\mathbf{z}_{i^\star} \in \ZZ^{4m}$ corresponding to the syndrome $\mathbf{v}_{i^\star} = \mathbf{F} \cdot \mathbf{z}_{i^\star} \bmod q$.
|
|
\end{itemize}
|
|
|
|
Regardless of the value of $coin$, queries to the random oracle~$H$
|
|
are handled by returning a uniformly chosen value in $\{1,2,3\}^t$. For
|
|
each $\kappa \leq Q_H$, we let~$r_{\kappa}$ denote the answer to the
|
|
$\kappa$-th $H$-query. Of course, if the adversary makes a given query
|
|
more than once, then~$\bdv$ consistently returns the previously defined
|
|
value. Queries to the random oracle $H_0$ are answered in the usual way, by returning a uniformly random value in the appropriate range. \medskip
|
|
|
|
\textbf{Forgery.} When $\adv$ halts, it outputs a
|
|
signature $ \Sigma^\star=\big( \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}_i}^\star, \pi_K^\star,sig^\star \big)$ on some message $M^\star$. At this point, $\bdv$ uses the
|
|
trapdoor $\mathbf{T}_{\mathbf{B}}$ to decrypt $\mathbf{c}_{\mathbf{v}_i}^\star$ and obtain an $m$-bit string $\textsf{bin}(\mathbf{v}^\star) \in \{0,1\}^m$.
|
|
|
|
If we parse the proof $\pi_K^\star$ as
|
|
$(\{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t,\mathsf{Chall}_K^\star,\{\mathsf{Resp}_{K,j}^\star \}_{j=1}^t)$,
|
|
the adversary $\adv$ must have invoked the random oracle~$H$ on the
|
|
input~$(M^\star, \mathsf{VK}^\star , \mathbf{c}_{\mathbf{v}_i}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$ with high probability.
|
|
Otherwise, the probability that
|
|
$\mathsf{Chall}_K^\star=H (M^\star, \mathsf{VK}^\star , \mathbf{c}_{\mathbf{v}_i}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
is negligible (at most~$3^{-t}$).
|
|
|
|
It comes that, with probability at least $ \varepsilon' \coloneqq \varepsilon-
|
|
3^{-t} $, $ (M^\star, \mathsf{VK}^\star , \mathbf{c}_{\mathbf{v}_i}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
coincides with the $\kappa^\star$-th random oracle query for some $\kappa^\star
|
|
\leq Q_H$.
|
|
|
|
At this stage, the reduction $\bdv$ runs the
|
|
adversary $\adv$ up to $32 \cdot Q_H / (\varepsilon - 3^{-t})$ times with the \textit{same} random tape and input as in the
|
|
initial run. All queries are answered as previously with
|
|
one difference in the treatment of random oracle queries.
|
|
Namely, the first $\kappa^\star-1$ random oracle queries -- which are
|
|
identical to those of the first execution since $\adv$ is run with the
|
|
same random tape as before -- receive the same answers
|
|
$\mathsf{Chall}_1,\ldots,\mathsf{Chall}_{\kappa^\star-1}$ as in the first run. This implies
|
|
that the $\kappa^\star$-th query will involve exactly the same tuple
|
|
$ (M^\star, \mathsf{VK}^\star , \mathbf{c}_{\mathbf{v}_i}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
as in the first run. However, from the
|
|
$\kappa^\star$-th query onwards, $\adv$ obtains fresh random oracle
|
|
values $\mathsf{Chall}_{\kappa^\star}',\ldots,\mathsf{Chall}_{Q_H}'$ at each new execution. The Improved Forking
|
|
Lemma of Brickell \textit{et al.}~\cite{BPVY00} guarantees that, with probability at least $1/2$, $\bdv$ can obtain a $3$-fork involving the
|
|
same tuple $ (M^\star, \mathsf{VK}^\star , \mathbf{c}_{\mathbf{v}_i}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$ with
|
|
pairwise distinct answers
|
|
$\mathsf{Chall}_{\kappa^\star}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star}^{(2)}, \mathsf{Chall}_{\kappa^\star}^{(3)} \in \{1,2,3\}^t$. With probability $1-(7/9)^t$ it can be shown that there exists an index $j \in \{1,\ldots,t\}$ for which the $j$-th bits
|
|
of $\mathsf{Chall}_{\kappa^\star}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star}^{(2)}, \mathsf{Chall}_{\kappa^\star}^{(3)}$ are $ (\mathsf{Chall}_{\kappa^\star,j}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star,j}^{(2)}, \mathsf{Chall}_{\kappa^\star,j}^{(3)} )=(1,2,3)$. From the corresponding responses $({\mathsf{Resp}_{K,j}^\star}^{(1)},{\mathsf{Resp}_{K,j}^\star}^{(2)},{\mathsf{Resp}_{K,j}^\star}^{(3)})$,
|
|
$\bdv$ is able to extract witnesses $(\mathbf{d}_{1}^\star,\mathbf{d}_{2}^\star) \in \ZZ^m \times \ZZ^m$, $\mathsf{id}^\star \in \{0,1\}^\ell$ and $\mathbf{w}^\star \in \{0,1\}^m$ from the proof of knowledge $\pi_K^\star$
|
|
such that
|
|
\begin{eqnarray*}
|
|
\mathbf{A}_{\mathsf{id}^\star } \cdot \begin{bmatrix} \mathbf{d}_{1}^\star \\ \hline \mathbf{d}_{2}^\star \end{bmatrix} &=& \mathbf{u} + \mathbf{D} \cdot \mathbf{w}^\star \\
|
|
\mathbf{w}^\star &=& \bit \bigl( \mathbf{D}_0 \cdot (\textsf{bin}(\mathbf{v}^\star) + \mathbf{D}_1 \cdot \mathbf{s}^\star \bigr) ,
|
|
\end{eqnarray*}
|
|
At this point, $\bdv$ aborts and
|
|
declares failure in the following situations:
|
|
|
|
\begin{itemize}
|
|
\item[-] $coin=0$ but $\mathsf{id}^\star \in \{0,1\}^\ell$ is recycled from some output of the $\mathcal{Q}_{\ajoin}$ oracle.
|
|
\item[-] $coin=0$ and $\mathsf{id}^\star \neq \mathsf{id}^\dagger$.
|
|
\item[-] $coin=1$ but $\mathsf{id}^\star \in \{0,1\}^\ell$ never appeared in a membership certificate returned by the $\mathcal{Q}_{\ajoin}$ oracle.
|
|
\item[-] $coin=1$ and $\mathsf{id}^\star \in \{0,1\}^{\ell}$ belongs to some user in $U^a$, but this user is not the one introduced at the $i^\star$-th
|
|
$\mathcal{Q}_{\ajoin}$-query (i.e., $i^\star \neq i^\dagger$ and $\mathsf{id}^\star \neq \mathsf{id}^\dagger$).
|
|
\item[-] $coin=1$ and the knowledge extractor revealed vectors $\textsf{bin}(\mathbf{v}^\star) \in \{0,1\}^{2m}$ and $\mathbf{s}^\star \in \ZZ^{2m}$
|
|
satisfying the collision (\ref{collide}),
|
|
where $ \textsf{bin}(\mathbf{v}_{i^\star})$ and $\mathbf{s}_{i^\star}$ are the vectors
|
|
involved in the $i^\star$-th $\mathcal{Q}_{\ajoin}$ query.
|
|
\item[-] $coin=2$ and the knowledge extraction yields vectors $\textsf{bin}(\mathbf{v}^\star) \in \{0,1\}^{2m}$ and $\mathbf{s}^\star \in \ZZ^{2m}$ such that the collision
|
|
(\ref{collide}) does not occur.
|
|
\end{itemize}
|
|
We call $\mathsf{fail}$ the event that one of the above situations occurs. Given that the choices of $coin \sample (\{0,1,2\})$ and $i^\star \sample ([1,Q_a])$ are completely independent of $\adv$'s view,
|
|
the choice of $coin$ is correct with probability $1/3$. If $coin=0$, $\bdv$'s choice of $\mathsf{id}^\dagger \sample (\{0,1\}^\ell)$ is correct with probability $1/(N_{\mathsf{gs}}-Q_a) \geq 1/N_{\mathsf{gs}}$ and, when
|
|
$coin=1$, $\bdv$'s correctly guesses $i^\star \in [1,Q_a]$ with probability $1/Q_a$. We find
|
|
$$\Pr[ \neg \mathsf{fail}] \geq \frac{1}{3 \cdot \max(N_{\mathsf{gs}},Q_a)} =\frac{1}{3 \cdot N_{\mathsf{gs}} } .$$
|
|
|
|
Assuming that $\mathsf{fail}$ does not occur, $\bdv$ can solve the problem instance as follows.
|
|
|
|
|
|
\begin{itemize}
|
|
\item If $coin=0$, we have $\mathsf{id}^\star=\mathsf{id}^\dagger$ and $\bdv$ knows a short vector $\mathbf{e}_u \in \ZZ^m$ such that $\mathbf{u} = \bar{\mathbf{A}}_1 \cdot \mathbf{e}_u \bmod q$. Hence, it can obtain a short integer vector
|
|
\begin{eqnarray*}
|
|
\mathbf{h} = {\mathbf{d}_1^\star} + \big( \mathbf{Q}_0 + \sum_{i=1}^\ell \mathsf{id}^\dagger [i] \mathbf{Q}_i \big) \cdot {\mathbf{d}_2^\star} - \mathbf{Q}_D
|
|
\cdot \textsf{bin}(\mathbf{v}^\star) - \mathbf{e}_u \in \ZZ^m
|
|
\end{eqnarray*}
|
|
such that $ \bar{\mathbf{A}}_1 \cdot \mathbf{h} = \mathbf{0}^m \bmod q$. Moreover,
|
|
we have $\mathbf{h} \neq \mathbf{0}^m$ w.h.p. since the syndrome $\mathbf{u} \in \Zq^n$ statistically hides
|
|
$\mathbf{e}_u \in \ZZ^m$
|
|
in $\Lambda_q^{\mathbf{u}}(\bar{\mathbf{A}}_1)$. Finally, the norm of $\mathbf{h}$ is at most $\| \mathbf{h} \|_2 \leq (\ell+1) \sigma^2 m^{3/2} + \sigma m^{1/2} (m+2)$.
|
|
This implies that $(\mathbf{h}^T \mid \mathbf{0}^m)^T$ is a short non-zero vector of $\Lambda_q^{\perp}(\bar{\mathbf{A}})$ and solves the initial $\mathsf{SIS}$ instance.
|
|
|
|
\item If $coin=1$, the extracted
|
|
witnesses $(\mathbf{d}_{1}^\star,\mathbf{d}_{2}^\star,\mathbf{s}^\star,\mathsf{id}^\star)$ and the decrypted $\textsf{bin}(\mathbf{v}^\star)$
|
|
satisfy $\mathsf{id}^\star=\mathsf{id}^\dagger$, $$\mathbf{w}^\star = \textsf{bin}( \mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}^\star) + \mathbf{D}_1 \cdot \mathbf{s}^\star )
|
|
\neq \textsf{bin}( \mathbf{D}_0 \cdot \textsf{bin}(\mathbf{v}_{i^\star}) + \mathbf{D}_1 \cdot \mathbf{s}_{i^\star} ) = \mathbf{w}_{i^\star} $$
|
|
(since $\neg \mathsf{fail}$ implies that the collision (\ref{collide}) did not occur if $coin=1$)
|
|
and
|
|
\begin{align} \label{rel1}
|
|
\left[
|
|
\begin{array}{c|c|c|c|c|c}
|
|
\mathbf{A} ~&~ \mathbf{A}_0 ~&~ \mathbf{A}_1~ &~ \ldots ~ & ~ \mathbf{A}_{\ell} ~&~ -\mathbf{D}
|
|
\end{array} \right] \cdot
|
|
\begin{bmatrix}
|
|
\mathbf{d}_{1}^\star \\ \hline \mathbf{d}_{2}^\star
|
|
\\ \hline \mathsf{id}^\dagger[1] \mathbf{d}_{2}^\star \\ \hline \vdots \\ \hline ~~ \mathsf{id}^\dagger[\ell] \mathbf{d}_{2}^\star
|
|
\\ \hline \mathbf{w}^\star
|
|
\end{bmatrix}
|
|
= \mathbf{u} \bmod q.
|
|
\end{align}
|
|
Since $\bdv$ already knew short vectors $(\mathbf{d}_{i^\star,1},\mathbf{d}_{i^\star,2}, \mathbf{w}_{i^\star}) \in \ZZ^m \times \ZZ^m \times \ZZ^m $ such that
|
|
\begin{align} \label{rel2}
|
|
\left[
|
|
\begin{array}{c|c|c|c|c|c}
|
|
\mathbf{A} ~&~ \mathbf{A}_0 ~&~ \mathbf{A}_1~ &~ \ldots ~ & ~ \mathbf{A}_{\ell} ~&~ -\mathbf{D}
|
|
\end{array} \right] \cdot
|
|
\begin{bmatrix}
|
|
\mathbf{d}_{i^\star,1}^\star \\ \hline \mathbf{d}_{i^\star,2}^\star
|
|
\\ \hline \mathsf{id}^\dagger[1] \mathbf{d}_{i^\star,2}^\star \\ \hline \vdots \\ \hline ~~ \mathsf{id}^\dagger[\ell] \mathbf{d}_{i^\star,2}^\star
|
|
\\ \hline \mathbf{w}_{i^\star}
|
|
\end{bmatrix}
|
|
= \mathbf{u} \bmod q,
|
|
\end{align}
|
|
by subtracting (\ref{rel2}) from (\ref{rel1}), we find that
|
|
\begin{align} \label{the-vec}
|
|
\mathbf{h} &= \mathbf{S} \cdot (\mathbf{d}_1^\star - \mathbf{d}_{i^\star,1}) + (\mathbf{S}_0 + \sum_{j=1}^\ell {\mathsf{id}^\dagger} [j] \mathbf{S}_j ) \cdot (\mathbf{d}_2^\star - \mathbf{d}_{i^\star,2} )
|
|
\ + ( \mathbf{w}^\star - \mathbf{w}_{i^\star} ) \quad
|
|
\end{align}
|
|
is a small-norm vector $\mathbf{h} \in \ZZ^m$ satisfying $ \bar{\mathbf{A}}_1 \cdot \mathbf{h}=\mathbf{0} \bmod q$. We claim that $\mathbf{h} \neq \mathbf{0}$ with high probability.
|
|
Indeed, we know that $\mathbf{w}^\star \neq \mathbf{w}_{i^\star}$ if $\neg \mathsf{fail}$ occurs.
|
|
This implies that the last term of (\ref{the-vec}) is non-zero, which rules out that $(\mathbf{d}_1^\star,\mathbf{d}_2^\star)=(\mathbf{d}_{i^\star,1},\mathbf{d}_{i^\star,2})$.
|
|
Since the columns of $\mathbf{S}$ and $\{\mathbf{S}_j\}_{j=0}^\ell$ have a lot of entropy conditionally on $\mathcal{Y}$, this implies that we can only have $\mathbf{h}=\mathbf{0}^m$ with negligible probability. Furthermore, the norm of $\mathbf{h}$ can be bounded by $\| \mathbf{h} \|_2 \leq 4 \sigma^2 m^{3/2} (\ell+2) + 2 m^{1/2} $,
|
|
so that $(\mathbf{h}^T \mid \mathbf{0}^m)^T$ solves the original $\mathsf{SIS}$ instance.
|
|
|
|
\item If $coin=2$, $\bdv$ is done as well since the collision (\ref{collide}) directly provides a vector
|
|
$$\mathbf{h}=\textsf{bin}(\mathbf{v}^\star) - \textsf{bin}(\mathbf{v}_i^\star) + \mathbf{Q} \cdot (\mathbf{s}^\star - \mathbf{s}_i^\star) ~ \in \ZZ^{2m}$$ of $\Lambda_q^{\perp}(\mathbf{D}_0)$ (which is also in
|
|
the lattice $\Lambda_q^{\perp}(\bar{\mathbf{A}})$ by construction) and has
|
|
norm $\| \mathbf{h} \|_2 \leq 2(\sigma^2 (2m)^{3/2} + (2m)^{1/2}) $. Moreover, $\mathbf{h} \in \ZZ^{2m}$ is non-zero with overwhelming probability
|
|
given that $\textsf{bin}(\mathbf{v}^\star) \neq \textsf{bin}(\mathbf{v}_i^\star)$ and the large amount of entropy retained by the columns $\mathbf{Q} \in \ZZ^{2m \times 2m}$ given $\mathbf{D}_1= \mathbf{D}_0 \cdot \mathbf{Q}$.
|
|
\end{itemize}
|
|
\end{proof}
|
|
|
|
|
|
\begin{theorem} \label{non-frame}
|
|
The scheme is secure against framing attacks under the $\mathsf{SIS}_{4n,4m,q,\beta''}$ assumption, where $\beta'' = 4\sigma \sqrt{m}$.
|
|
\end{theorem}
|
|
|
|
\begin{proof}
|
|
Let us assume that a PPT adversary $\adv$ can create a
|
|
forgery $(M^\star,\Sigma^\star)$ that opens to some honest user
|
|
$i\in U^b$ who did not sign $M^\star$. In the random oracle model, we give a reduction $\bdv$ that uses $\adv$ to solve an instance of the ~$\SIS_{4n,4m,q,\beta''}$ problem:
|
|
$\bdv$ takes as input~$\bar{\mathbf{A}} \in
|
|
\Zq^{4n \times 4m}$ and finds a non-zero short vector $\mathbf{w} \in
|
|
\Lambda_q^{\perp}(\bar{\mathbf{A}})$. % with~$0 < \|\mathbf{w}\| \leq \beta$.
|
|
\\
|
|
\indent Algorithm $\bdv$ generates the group public key $\mathcal{Y}$ by faithfully running the real setup algorithm with the sole difference that, at step 2 of $\mathsf{Setup}$,
|
|
$\bdv$ defines $\mathbf{F}=\bar{\mathbf{A}} \in \Zq^{4n \times 4m}$. However, the distribution of $\mathcal{Y}$ is as in the real scheme.
|
|
As a result of having generated $\mathcal{Y}$ itself, $\bdv$ knows
|
|
$\mathcal{S}_{\GM}=\mathbf{T}_{\mathbf{A}}$ and $\mathcal{S}_{\OA}= \mathbf{T}_{\mathbf{B}}$. The adversary $\bdv$ is run on input of the
|
|
group public key
|
|
$$ \mathcal{Y}\coloneqq \Bigl(\mathbf{A}, \{\mathbf{A}_j\}_{j=0}^\ell,~\mathbf{B},~\mathbf{D},~\mathbf{D}_0,~\mathbf{D}_1,~\mathbf{F}=\bar{\mathbf{A}},~\mathbf{u},~\Pi^{\mathsf{OTS}},~H,~H_0 ) \Bigr). $$
|
|
|
|
If $\adv$ chooses
|
|
to corrupt the group manager or the opening authority during the
|
|
game, $\bdv$ is able to reveal
|
|
$\mathcal{S}_{\GM}=\mathbf{T}_{\mathbf{A}}$ and
|
|
$\mathcal{S}_{\OA}= \mathbf{T}_{\mathbf{B}}$. % At the very beginning of the game, $\bdv$ draws a random index $j^\star \sample \{1,\ldots,Q_b\}$ and
|
|
Then, $\bdv$ starts interacting with $\adv$ as follows.
|
|
\begin{itemize}
|
|
\item[-] $Q_{\mathsf{keyGM}}$-queries: If $\adv$ decides to corrupt the group manager, $\bdv$
|
|
hands the secret key $\mathcal{S}_{\GM}=\mathbf{T}_{\mathbf{A}}$ to $\adv$.
|
|
\item[-] $Q_{\bjoin}$-queries: At any time $\adv$ can act as a corrupted group manager and introduce a new honest user $i$ in the group by invoking the $Q_{\bjoin}$ oracle.
|
|
At each $Q_{\bjoin}$-query, $\bdv$ faithfully
|
|
runs $\mathsf{J}_{\mathsf{user}}$ on behalf of the honest user in an execution of $\mathsf{Join}$ protocol.
|
|
|
|
\item[-] $Q_{\mathsf{pub}}$-queries: These
|
|
can be answered as in the real game, by having the simulator return
|
|
$\mathcal{Y}$.
|
|
\item[-] $Q_{\mathsf{sig}}$-queries: When the adversary $\adv$ requests user $ i \in
|
|
U^b$ to sign a message $M$, $\bdv$ first generates a one-time key pair $(\mathsf{VK},\mathsf{SK}) \leftarrow \mathcal{G}(n)$ to
|
|
compute $\mathbf{G}_0=H_0(\mathsf{VK}) \in \Zq^{n \times 2m}$. Next,
|
|
$\bdv$ recalls the vector $\mathbf{z}_i \in \ZZ^{4m}$ that was chosen to define the syndrome $\mathbf{v}_i = \mathbf{F} \cdot \mathbf{z}_i$ at step 1 of the $\mathsf{Join}$ protocol as well as
|
|
the identifier $\mathsf{id}_i \in \{0,1\}^\ell$ and the short vectors $(\mathbf{d}_{i,1},\mathbf{d}_{i,2},\mathbf{s}_i) $
|
|
that were supplied by $\adv$ in an earlier $Q_{\bjoin}$-query. It faithfully computes a signature by IBE-encrypting
|
|
$\textsf{bin}(\mathbf{v}_i) \in \{0,1\}^{2m}$ and using $(\mathbf{d}_{i,1},\mathbf{d}_{i,2},\mathbf{s}_i,\mathbf{z}_i,\mathbf{s}_i,\mathsf{id}_i)$ to compute a witness indistinguishable proof $\pi_K=(
|
|
\{\mathsf{Comm}_{K,j}\}_{j=1}^t,\mathsf{Chall}_K,\{\mathsf{Resp}_{K,j}\}_{j=1}^t)$.
|
|
Finally, $\bdv$ computes a one-time signature
|
|
$sig=\mathcal{S}(\mathsf{SK},(\mathbf{c}_{\mathbf{v}_i},\pi_K))$ and returns the signature
|
|
$\Sigma=\big( \mathsf{VK} ,\mathbf{c}_{\mathbf{v}_i}, \pi_K,sig \big)$ to $\adv$.
|
|
\end{itemize}
|
|
When $\adv$ halts, it outputs a signature
|
|
$ \Sigma^\star = \big( \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \pi_K^\star,sig^\star \big)$
|
|
for
|
|
some message $M^\star$, which opens to ${i^\star} \in
|
|
U^b$ although user $i^\star$ did not sign the message $M^\star$ at any time. Since $(M^\star,\Sigma^\star)$ supposedly frames user $i^\star$, the opening of
|
|
$\Sigma^\star$ must reveal the $m$-bit string $\textsf{bin}(\mathbf{v}_{i^\star}) \in \{0,1\}^m$. We note that the reduction $\bdv$ has
|
|
recollection of a short vector $\mathbf{z}_{i^\star} \in \ZZ^{4m}$ (of norm $\| \mathbf{z}_{i^\star} \| < 2\sigma \sqrt{m}$)
|
|
such that $\mathbf{v}_{i^\star} = \mathbf{F} \cdot \mathbf{z}_{i^\star} \bmod q$ which it
|
|
chose when running $\mathsf{J}_{\mathsf{user}}$ on behalf of user $i^\star$ when this user was introduced in the group. Hence,
|
|
$\bdv$
|
|
would be able to solve its given $\mathsf{SIS}$ instance if it had another short vector $\mathbf{z}' \in \ZZ^{4m}$ satisfying $\mathbf{v}_{i^\star} = \mathbf{F} \cdot {\mathbf{z}'} \bmod q $.
|
|
To compute such a
|
|
vector, $\bdv$ proceeds by replaying the adversary $\adv$ sufficiently many times and applying the Improved Forking
|
|
Lemma of Brickell \textit{et al.}~\cite{BPVY00}. \\
|
|
\indent
|
|
If we parse $\pi_K^\star$ as
|
|
$(\{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t,\mathsf{Chall}_K^\star,\{\mathsf{Resp}_{K,j}^\star \}_{j=1}^t)$, with high
|
|
probability, $\adv$ must have queried~$H$ on the
|
|
input~$ (M^\star, \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$.
|
|
Otherwise, we would only have
|
|
$\mathsf{Chall}_K^\star=H (M^\star, \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
with negligible probability~$3^{-t}$. It comes that, with probability at least $ \varepsilon' \coloneqq \varepsilon-
|
|
3^{-t} $, the tuple $ (M^\star, \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
was the input of the $\kappa^\star$-th random oracle query for some index $\kappa^\star
|
|
\leq Q_H$. \\
|
|
\indent
|
|
At this point, the reduction $\bdv$ runs the
|
|
adversary $\adv$ up to $32 \cdot Q_H / (\varepsilon - 3^{-t})$ times with the \textit{same} random tape and input as in the
|
|
first run. All queries are answered as previously with
|
|
one difference in the way to handle $H$-queries.
|
|
Namely, the first $\kappa^\star-1$ $H$-queries -- which are
|
|
the same as in the first execution since $\adv$ is run with the
|
|
same random tape -- obtain the same answers
|
|
$\mathsf{Chall}_1,\ldots,\mathsf{Chall}_{\kappa^\star-1}$ as in the original run. This implies
|
|
that the $\kappa^\star$-th query will also involve exactly the same tuple
|
|
$ (M^\star, \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$
|
|
as in the original run. From the
|
|
$\kappa^\star$-th query forward, however, the adversary $\adv$ obtains fresh random oracle
|
|
outputs $\mathsf{Chall}_{\kappa^\star}',\ldots,\mathsf{Chall}_{Q_H}'$ at each new execution. The Improved Forking
|
|
Lemma of~\cite{BPVY00} ensures that, with probability $>1/2$, $\bdv$ obtains a $3$-fork involving the
|
|
tuple $ (M^\star, \mathsf{VK}^\star ,\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$ of the initial run and with
|
|
pairwise distinct answers
|
|
$\mathsf{Chall}_{\kappa^\star}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star}^{(2)}, \mathsf{Chall}_{\kappa^\star}^{(3)} \in \{1,2,3\}^t$. Since the forgeries of the $3$-fork all correspond to the tuple $ (M^\star, \mathsf{VK}^\star ,
|
|
\mathbf{c}_{\mathbf{v}}^\star, \{\mathsf{Comm}_{K,j}^\star\}_{j=1}^t)$, they open to the same $m$-bit string $\textsf{bin}(\mathbf{v}_{i^\star}) \in \{0,1\}^m$ and
|
|
which is uniquely determined
|
|
by $\mathbf{c}_{\mathbf{v}}^\star$. In turn, this implies that the three forgeries all reveal the same $\textsf{bin}(\mathbf{v}_{i^\star})$
|
|
at the second step of $\mathsf{Open}$.
|
|
With probability $1-(7/9)^t$ it can be shown that there exists $j \in \{1,\ldots,t\}$ such that the $j$-th bits
|
|
of $\mathsf{Chall}_{\kappa^\star}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star }^{(2)}, \mathsf{Chall}_{\kappa^\star }^{(3)}$ are $ (\mathsf{Chall}_{\kappa^\star,j}^{(1)} ,
|
|
\mathsf{Chall}_{\kappa^\star,j}^{(2)}, \mathsf{Chall}_{\kappa^\star,j}^{(3)} )=(1,2,3)$. From the corresponding responses $({\mathsf{Resp}_{K,j}^\star}^{(1)},{\mathsf{Resp}_{K,j}^\star}^{(2)},{\mathsf{Resp}_{K,j}^\star}^{(3)})$,
|
|
$\bdv$ is able to extract a short vector $ \mathbf{z}' \in \ZZ^{4m} $ such that $\mathbf{v}_{i^\star} = \mathbf{F} \cdot {\mathbf{z}'} \bmod q $. \\ \indent Due to the statistical witness indistinguishability of
|
|
the Stern-like proof of knowledge which is used to generate signature, with overwhelming
|
|
probability, we have $\mathbf{z}' \neq \mathbf{z}_{i^\star}$. Indeed, from the adversary's view, the distribution of
|
|
$\mathbf{z}_{i^\star}$ is $D_{\Lambda_q^{\mathbf{v}_{i^\star}}(\mathbf{F}),\sigma}$, which means that it has at least $n$ bits of min-entropy.
|
|
Hence, the difference $\mathbf{h} = \mathbf{z}' - \mathbf{z}_{i^\star} \in \ZZ^{4m}$ is a suitably short non-zero vector of $ \Lambda_q^{\perp}( \bar{\mathbf{A}} ) $.
|
|
\end{proof}
|
|
|
|
\begin{theorem} \label{anonymity-thm}
|
|
In the random oracle model, the scheme provides \textsf{CCA}-anonymity if
|
|
the $\LWE_{n,q,\chi}$ assumption holds and if $\Pi^\mathrm{OTS}$ is a strongly unforgeable one-time signature.
|
|
\end{theorem}
|
|
|
|
\begin{proof}
|
|
We proceed as in~\cite{LNW15} and prove the result via a sequence of games which are computationally indistinguishable.
|
|
The first game consists of the real anonymity experiment which is parameterized by a bit $d \in \{0,1\}$ that determines the challenger's choice in the challenge phase.
|
|
The last game is the same regardless of whether $d=0$ or $d=1$. It follows that, under the stated assumptions, no PPT adversary can distinguish $\Expt^\textrm{anon$-0$}_\adv$ from $\Expt^\textrm{anon$-1$}_\adv$ with noticeable advantage.
|
|
\medskip
|
|
|
|
\begin{description}
|
|
\item[$\textsf{Game}^{(d)}$~0:] This is the real anonymity experiment $\Expt^\textrm{anon$-d$}_\adv(\lambda)$ as described in Definition~\ref{def:anon}.
|
|
More precisely, the challenger starts by running the~$\mathsf{Setup}(1^\lambda, 1^{\Ngs})$ algorithm to obtain $(\gspk, \mathcal{S}_\GM = \mathbf{T_A} \in \ZZ^{m \times m}, \mathcal{S}_\OA = \mathbf{T_B} \in \ZZ^{m \times m})$ along with state information $St$. The challenger next hands the public parameters $\gspk$ and the group manager key $\mathcal{S}_\GM$ to the adversary $\adv$.
|
|
On the following adversary signature opening queries on signatures $\Sigma = (\vk, \mathbf{c}_{\mathbf{v}_d}, \pi_K, sig)$, the challenger uses the opening authority key $\mathbf{T_A} \in \ZZ^{m \times m}$ he possesses to decrypt the GPV encryption of the signer identity $\mathbf{c}_{\mathbf{v}_d} \in \Zq^m \times \Zq^{2m}$.
|
|
At some point, the adversary $\adv$ requests a challenge by outputting a target message $M^\star \in \bit^*$ and two user key pairs
|
|
\[ \bigl(\scr_i^\star = \mathbf{z}^\star_i \in \ZZ^{4m}, \crt_i^\star \in (\mathsf{id}^\star_i, \mathbf{d}^\star_i, \mathbf{s}^\star_i) \in \bit^\ell \times \ZZ^{2m} \times \ZZ^{2m} \bigr)_{i \in \bit} \]
|
|
which must be valid and distinct (otherwise, the challenger aborts the experiment).
|
|
This challenge query is answered by having the challenger return a signature of the target message under the identity $id_d$: namely, this challenge signature is computed as $\Sigma^\star = (\vk^\star, \mathbf{c}_{\mathbf{v}_d}^\star, \pi_K^\star, sig^\star) \gets \Sign(\mathcal{Y}, \crt_d^\star, \scr_d^\star, M^\star)$ for the given parameter $d$
|
|
of the \textsf{Game}.
|
|
Finally, $\adv$ outputs a bit $d' \in \bit$ which is also the experiment's output. % and the experiment outputs $1$ if $b = b'$ or $0$ otherwise. By assumption, $\adv$ has advantage $\varepsilon$ in this game.
|
|
\smallskip
|
|
|
|
\item[$\textsf{Game}^{(d)}$~1:] In this experiment, we slightly change $\mathsf{Game}^{(d)}~0$ as follows. At the outset of the game, the challenger generates the one-time signature key pair $(\vk^\star, \sk^\star)$ that will be used in the challenge phase.
|
|
During the game, if the adversary $\adv$ requests the opening of a valid signature $\Sigma = (\vk, \mathbf{c}_{\mathbf{v}_i}, \pi_K, sig)$ where $\vk = \vk^\star$, the challenger returns a random bit and aborts.
|
|
However, this event $F_1$ would contradict the strong unforgeability of the one-time signature $\Pi^{\mathrm{OTS}}$.
|
|
Indeed, before the challenge phase $\vk^\star$ is independent of $\adv$'s view and the probability that $\vk^\star$ shows up in $\adv$'s queries is negligible.
|
|
After seeing the challenge signature $\Sigma^\star$, if $\adv$ comes up with a valid signature $\Sigma = (\vk, \mathbf{c}_{\mathbf{v}_i}, \pi_K, sig)$ such that $\vk = \vk^\star$, then $sig$ is a forged one-time signature, which defeats the strong unforgeability of $\Pi^{\mathrm{OTS}}$.
|
|
Therefore the probability $\Pr[F_1]$ that the challenger aborts in this experiment is negligible.
|
|
From here on, we thus assume that $\adv$'s opening queries for valid signatures do not include $\vk^\star$.
|
|
\smallskip
|
|
|
|
\item[$\textsf{Game}^{(d)}$~2:] In this game, we program the random oracle $H_0$ in the following way: at the beginning of the game, we choose
|
|
a uniformly random matrix $\mathbf{G}_0^\star \sample (\Zq^{n \times 2m})$ and set $H_0(\vk^\star) = \mathbf{G}^\star_0$. From the adversary's view, the distribution of
|
|
$\mathbf{G}_0^\star$ is statistically close to the one in the real attack game, as in \cite{GPV08}.
|
|
As for other queries, for each fresh $H_0$-queries on $\vk$,
|
|
the challenger samples small-norm matrices $\mathbf{E}_{0,\vk} \sample D_{\ZZ^m, \sigma}^{2m}$ and programs the oracle such that
|
|
$H_0(\vk) = \mathbf{B} \cdot \mathbf{E}_{0,\vk} \bmod q$. The chosen matrices $\mathbf{E}_{0,\vk}$
|
|
are retained for later use.
|
|
Note that the values of $H_0(\vk)$ are statistically close to the uniform.
|
|
For any query involving a previously queried $\vk$, the challenger consistently returns the previously stored images.
|
|
The adversary's view remains the same as in $\mathsf{Game}^{(d)}~1$, analogously to the security proof of the GPV IBE~\cite{GPV08}.
|
|
\smallskip
|
|
|
|
\item[$\textsf{Game}^{(d)}$~3:] Here, we will change the behavior of the opening algorithm.
|
|
Namely, at each fresh oracle query, we still store the matrices $\mathbf{E}_{0,\vk} \in \Zq^{m \times 2m}$ and, at the beginning of the game, the challenger
|
|
samples an uniformly random $\mathbf{B^\star} \in \Zq^{n \times m}$ that is later used in place of $\mathbf{B}$ to answer $H_0$-queries.
|
|
To answer the adversary's queries of the opening of a signature
|
|
$\Sigma = (\vk, \mathbf{c}_{\mathbf{v}_i}, \ \pi_K, sig)$,
|
|
the challenger recalls the small-norm matrices $\mathbf{E}_{0,\vk}$ which were defined when $\adv$ first queried $H_0(\vk)$.
|
|
These matrices are used as ``decryption matrices'' to open $\Sigma$ for the corresponding $\mathbf{G}_0 = H_0(\vk) \in \Zq^{n \times 2m}$.
|
|
For similar reasons as in the security proof of~\cite{GPV08}, the distribution of $\mathbf{G}_0$ is statistically close to the uniform,
|
|
which implies that $\mathsf{Game}^{(d)}~2$ and $\mathsf{Game}^{(d)}~3$ are statistically indistinguishable.
|
|
\smallskip
|
|
|
|
|
|
\item[$\textsf{Game}^{(d)}$~4:] Instead of faithfully generating the
|
|
NIZKPoK $\pi_K$ of Section~\ref{subsection:zk-for-group-signature}, the challenger simulates the proof without using the witness (note that this is possible since the HVZK property of the underlying proof system is preserved
|
|
under parallel repetitions). This
|
|
is done by running the simulator for the underlying interactive protocol for
|
|
each $j \in \{1,\ldots, t\}$, and then programming the random oracle $H$
|
|
accordingly. The challenge signature
|
|
$\Sigma^\star = (\vk^\star, \mathbf{c}_{\mathbf{v}_d}^\star , \pi_K^\star, sig^\star)$
|
|
is statistically close to the challenge signature of the previous game, because the
|
|
proof system is statistically zero-knowledge as stated in Lemma~\ref{le:zk-ktx}.
|
|
Consequently, $\mathsf{Game}^{(d)}~3$ and $\mathsf{Game}^{(d)}~4$ are indistinguishable.
|
|
\smallskip
|
|
|
|
\item[$\textsf{Game}^{(d)}$~5:] In this game, we modify the generation of the challenge ciphertext $\mathbf{c}_{\mathbf{v}_d}^\star$.
|
|
Instead of using the real encryption algorithm of the GPV IBE to compute $\mathbf{c}_{\mathbf{v}_d}^\star$ as the encryption of $\mathbf{v}_d^\star = \mathbf{F} \cdot \mathbf{z}_d \in \Zq^{4n}$, we return truly random
|
|
ciphertexts. In other words, we let
|
|
\[ \mathbf{c}_{\mathbf{v}_d}^\star = \begin{pmatrix}
|
|
\mathbf{r}_1 \\ \mathbf{r}_2 + \textsf{bin}(\mathbf{v}_{d}^\star) \lfloor q/2 \rfloor
|
|
\end{pmatrix}, \]
|
|
%where $\mathbf{v}_{i_b}^\star= \mathbf{F} \cdot \mathbf{z}_{i_b}^\star $, and
|
|
where $\mathbf{r}_1 \sample (\Zq^{m})$, $\mathbf{r}_2 \sample (\Zq^{2m})$ are uniformly random.
|
|
The hardness of the decisional $\LWE_{n, q, \chi}$ problem implies that $\mathbf{c}^\star_{\mathbf{v}_d}$ in \ extsf{Game} $4$ and \ extsf{Game} $5$ are computationally indistinguishable.
|
|
If $\adv$ can distinguish between these two games, it can furthermore distinguish
|
|
\[ \begin{pmatrix}
|
|
\mathbf{B}^T \\ \hline {\mathbf{G}_0^\star }^T
|
|
\end{pmatrix} \mathbf{e}_0 + \begin{pmatrix} \mathbf{x}_1 \\\hline \mathbf{x}_2 \end{pmatrix} \mbox{ from } \begin{pmatrix}
|
|
\mathbf{r}_1 \\ \hline \mathbf{r}_2
|
|
\end{pmatrix},\]
|
|
which would break the decisional $\LWE_{n,q,\chi}$ assumption.
|
|
|
|
Therefore, $\mathsf{Game}^{(d)}~4$ and $\mathsf{Game}^{(d)}~5$ are computationally indistinguishable.
|
|
\smallskip
|
|
|
|
\item[\textsf{Game}~6:] We finally make a conceptual modification on the previous game. Namely we sample uniformly random $\mathbf{r}_1^\prime
|
|
\sample (\Zq^{m})$, $\mathbf{r}_2^\prime \sample (\Zq^{2m})$ and assign
|
|
\[ \mathbf{c}_{\mathbf{v}_d}^\star = \begin{pmatrix}
|
|
\mathbf{r}_1^\prime \\ \mathbf{r}_2^\prime
|
|
\end{pmatrix} .\]
|
|
\end{description}
|
|
|
|
Clearly, the distribution of $\mathbf{c}_{\mathbf{v}_i}^\star $ has not changed since $\mathsf{Game}^{(d)}~5$. Since \textsf{Game} $6$ does no longer depend on the
|
|
challenger's bit $d\in \{0,1\}$, the result follows.
|
|
\end{proof}
|
|
|
|
\section{Subprotocols for Stern-like Argument}
|
|
\addcontentsline{tof}{section}{\protect\numberline{\thesection} Protocoles pour les preuves à la Stern}
|
|
\label{se:gs-lwe-stern}
|
|
|
|
\subsection{Proving the Consistency of Commitments}\label{subsection:zk-for-commitments}
|
|
The argument system used in our protocol for signing a committed value in Section~\ref{commit-sig} can be summarized as follows.
|
|
\begin{description}
|
|
\item[Common Input:] Matrices $\{\mathbf{D}_k\in \ZZ_q^{2n \times 2m}\}_{k=0}^N$; $\mathbf{B}\in \ZZ_q^{n \times m}$; $\mathbf{G}_1 \in \mathbb{Z}_q^{n \times 2m}$;
|
|
|
|
\smallskip
|
|
\hspace*{-7.5pt}vectors $\mathbf{c}_{\mathfrak{m}} \in \mathbb{Z}_q^{2n}$; $\{\mathbf{c}_{k,1} \in \ZZ_q^{m}\}_{k=1}^N$; $\{\mathbf{c}_{k,2} \in \ZZ_q^{2m}\}_{k=1}^N$; $\mathbf{c}_{\mathbf{s}', 1} \in \ZZ_q^{m}$; $\mathbf{c}_{\mathbf{s}',2} \in \mathbb{Z}_q^{2m}$. \medskip
|
|
\item[Prover's Input:] $\mathfrak{m} = (\mathfrak{m}_1^T \| \ldots \| \mathfrak{m}_N^T)^T \in \mathsf{CorEnc}(mN)$;
|
|
|
|
$\{\mathbf{s}_{k} \in [-B,B]^n, \hspace*{2.5pt} \mathbf{e}_{k,1}\in [-B,B]^m; \hspace*{2.5pt} \mathbf{e}_{k,2}\in [-B,B]^{2m}\}_{k=1}^N$; \hspace*{5pt} $\mathbf{s}_0\in [-B,B]^n$;
|
|
|
|
$\mathbf{e}_{0,1}\in [-B,B]^m; \hspace*{5pt} \mathbf{e}_{0,2}\in [-B,B]^{2m}$; \hspace*{5pt} $\mathbf{s}' \in [-(p-1), (p-1)]^{2m}$ \smallskip
|
|
\item[Prover's Goal:] Convince the verifier in \textsf{ZK} that:
|
|
\end{description}
|
|
|
|
\begin{eqnarray}\label{equation:R-commit-statement}
|
|
\hspace*{-5pt}
|
|
\begin{cases}
|
|
\mathbf{c}_{\mathfrak{m}}= \mathbf{D}_0 \cdot \mathbf{s}' + \sum_{k=1}^N \mathbf{D}_k \cdot \mathfrak{m}_k \bmod q; \\[2.5pt]
|
|
|
|
\mathbf{c}_{\mathbf{s}', 1}= \mathbf{B}^T\cdot \mathbf{s}_0 + \mathbf{e}_{0,1} \bmod q; \hspace*{5pt}\mathbf{c}_{\mathbf{s}',2}= \mathbf{G}_1^T\cdot \mathbf{s}_0 + \mathbf{e}_{0,2} + \lfloor q/p \rfloor\cdot \mathbf{s}'\bmod q; \\[2.5pt]
|
|
\forall k\in [N]: \mathbf{c}_{k,1}= \mathbf{B}^T\cdot\mathbf{s}_{k} + \mathbf{e}_{k,1}; \hspace*{5pt}\mathbf{c}_{k,2}= \mathbf{G}_1^T\cdot \mathbf{s}_{k} + \mathbf{e}_{k,2} + \lfloor q/2 \rfloor\cdot \mathfrak{m}_k.
|
|
\end{cases}
|
|
\end{eqnarray}
|
|
We will show that the above argument system can be obtained from the one in \cref{sse:stern-abstraction}. We proceed in two steps.
|
|
|
|
\smallskip \smallskip
|
|
|
|
\textbf{Step 1:} \emph{Transforming the equations in~(\ref{equation:R-commit-statement}) into a unified one of the form $\mathbf{P}\cdot \mathbf{x} = \mathbf{v} \bmod q$, where $\|\mathbf{x}\|_\infty =1$ and $\mathbf{x} \in \mathsf{VALID}$ - a ``specially-designed'' set.}
|
|
|
|
To do so, we first form the following vectors and matrices:
|
|
\[
|
|
\scriptsize
|
|
\begin{cases}
|
|
\mathbf{x}_1 \hspace*{-1pt}= \hspace*{-1pt}\big(\mathbf{s}_0^T \| \mathbf{e}_{0,1}^T \| \mathbf{e}_{0,2}^T \| \mathbf{s}_{1}^T \| \mathbf{e}_{1,1}^T \| \mathbf{e}_{1,2}^T \| \ldots \| \mathbf{s}_{N}^T \| \mathbf{e}_{N,1}^T \| \mathbf{e}_{N,2}^T \big)^T\hspace*{-3.5pt} \in \hspace*{-1.5pt}[-B,B]^{(n+3m)(N+1)}; \\[2.5pt]
|
|
\mathbf{v} = \big(\mathbf{c}_{\mathfrak{m}}^T \| \mathbf{c}_{\mathbf{s}',1}^T\| \mathbf{c}_{\mathbf{s}',2}^T\| \mathbf{c}_{1,1}^T \|\mathbf{c}_{1,2}^T \| \ldots \|\mathbf{c}_{N,1}^T \|\mathbf{c}_{N,2}^T \big)^T \in \mathbb{Z}_q^{2n + 3m(N+1)};\\[5pt]
|
|
\mathbf{P}_1 = \left(
|
|
\begin{array}{ccc}
|
|
\begin{array}{c}
|
|
\mathbf{B}^T \\
|
|
\hline
|
|
\rule{0pt}{3ex}\mathbf{G}_1^T
|
|
\end{array}
|
|
& \vline
|
|
& \mathbf{I}_{3m}
|
|
\end{array}
|
|
\right); \hspace*{10pt}
|
|
\mathbf{Q}_2 = \left(
|
|
\begin{array}{c}
|
|
\mathbf{0} \\
|
|
\hline
|
|
\rule{0pt}{3ex}\lfloor\frac{q}{2}\rfloor\mathbf{I}_{2m}
|
|
\end{array}
|
|
\right); \hspace*{10pt}
|
|
\mathbf{Q}_p = \left(
|
|
\begin{array}{c}
|
|
\mathbf{0} \\
|
|
\hline
|
|
\rule{0pt}{3ex}\lfloor\frac{q}{p}\rfloor\mathbf{I}_{2m}
|
|
\end{array}
|
|
\right)\\[5pt]
|
|
|
|
\mathbf{M}_1 = \left(
|
|
\begin{array}{c}
|
|
\mathbf{0} \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
|
|
\begin{array}{cccc}
|
|
\mathbf{P}_1 & & & \\
|
|
& \mathbf{P}_1 & & \\
|
|
& & \xddots & \\
|
|
& & & \mathbf{P}_1 \\
|
|
\end{array}
|
|
|
|
\\
|
|
\end{array}
|
|
\right); \hspace*{15pt}
|
|
\mathbf{M}_2 = \left(
|
|
\begin{array}{c}
|
|
\mathbf{D}_1 | \ldots | \mathbf{D}_N \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
\mathbf{0} \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
\begin{array}{ccc}
|
|
\mathbf{Q}_2 & & \\
|
|
& \xddots & \\
|
|
& & \mathbf{Q}_2 \\
|
|
\end{array}
|
|
|
|
\\
|
|
\end{array}
|
|
\right); \hspace*{15pt}
|
|
\mathbf{M}_3 = \left(
|
|
\begin{array}{c}
|
|
\mathbf{D}_0 \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
\mathbf{Q}_p \\
|
|
\hline
|
|
\rule{0pt}{3ex}
|
|
\\
|
|
\mathbf{0} \\
|
|
\\
|
|
\end{array}
|
|
\right).
|
|
\end{cases}
|
|
\]
|
|
|
|
We then observe that (\ref{equation:R-commit-statement}) can be rewritten as:
|
|
\begin{eqnarray}\label{equation:R-commit-unified}
|
|
\mathbf{M}_1 \cdot \mathbf{x}_1 + \mathbf{M}_2 \cdot \mathfrak{m} + \mathbf{M}_3 \cdot \mathbf{s}' = \mathbf{v} \in \mathbb{Z}_q^D,
|
|
\end{eqnarray}
|
|
where $D = 2n + 3m(N+1)$.
|
|
Now we employ the techniques from \cref{sse:stern-abstraction} to convert~\eqref{equation:R-commit-unified} into the form $\mathbf{P}\cdot \mathbf{x} = \mathbf{v} \bmod q$. Specifically, if we let:
|
|
\[
|
|
\begin{cases}
|
|
\mathsf{DecExt}_{(n+3m)(N+1),B}(\mathbf{x}_1) \rightarrow \hat{\mathbf{x}}_1 \in \mathsf{B}^3_{(n+3m)(N+1)\delta_B}; \\[2.5pt]
|
|
{\mathbf{M}}'_1 = \mathbf{M}_1 \cdot \widehat{\mathbf{K}}_{(n+3m)(N+1),B} \in \ZZ_q^{D \times 3(n+3m)(N+1)\delta_B}; \\[2.5pt]
|
|
\mathsf{DecExt}_{2m, p-1}(\mathbf{s}') \rightarrow \hat{\mathbf{s}} \in \mathsf{B}^3_{2m\delta_{p-1}}; \hspace*{5pt}
|
|
{\mathbf{M}}'_3 = \mathbf{M}_3 \cdot \widehat{\mathbf{K}}_{2m,p-1} \in \mathbb{Z}_q^{D \times 6m\delta_{p-1}},
|
|
\end{cases}
|
|
\]
|
|
$L = 3(n+3m)(N+1)\delta_B + 2mN + 6m\delta_{p-1}$, and $\mathbf{P} \hspace*{-1pt}= \hspace*{-1pt}\big[\mathbf{M}'_1 | \mathbf{M}_2 | \mathbf{M}'_3\big] \hspace*{-2pt}\in \hspace*{-1pt}\mathbb{Z}_q^{D \times L}$, and $\mathbf{x} = \big(\hat{\mathbf{x}}_1^T \| \mathfrak{m}^T \| \hat{\mathbf{s}}^T\big)^T$, then we will obtain the desired equation:
|
|
\[
|
|
\mathbf{P}\cdot \mathbf{x} = \mathbf{v} \bmod q.
|
|
\]
|
|
Having performed the above unification, we now define $\mathsf{VALID}$ as the set of all vectors $\mathbf{t} \hspace*{-1pt}\in\hspace*{-1pt} \{-1,0,1\}^L$ of the form $\mathbf{t}\hspace*{-1pt} =\hspace*{-1pt} \big(\mathbf{t}_1^T \| \mathbf{t}_2^T \| \mathbf{t}_3^T\big)^T$\hspace*{-2.5pt}, where $\mathbf{t}_1 \in \mathsf{B}^3_{(n+3m)(N+1)\delta_B}$, $\mathbf{t}_2 \in \mathsf{CorEnc}(mN)$, and $\mathbf{t}_3 \in \mathsf{B}^3_{2m\delta_{p-1}}$. Note that $\mathbf{x} \in \mathsf{VALID}$. \\
|
|
|
|
\smallskip
|
|
|
|
\textbf{Step 2:} \emph{Specifying the set $\mathcal{S}$ and permutations of $L$ elements $\{T_\pi: \pi \in \mathcal{S}\}$ for which the conditions in~(\ref{eq:zk-equivalence}) hold.}
|
|
|
|
\begin{itemize}
|
|
\item Define $\mathcal{S}: = \mathcal{S}_{3(n+3m)(N+1)\delta_B} \times \{0,1\}^{mN} \times \mathcal{S}_{6m\delta_{p-1}}$. \smallskip
|
|
\item For $\pi = (\pi_1, \mathbf{b}, \pi_3) \in \mathcal{S}$, and for vector $\mathbf{w} = \big(\mathbf{w}_1^T \| \mathbf{w}_2^T \| \mathbf{w}_3^T\big)^T \in \mathbb{Z}_q^L$, where $\mathbf{w}_1 \in \ZZ_q^{3(n+3m)(N+1)\delta_B}$, $\mathbf{w}_2 \in \ZZ_q^{2mN}$, $\mathbf{w}_3 \in \ZZ_q^{6m\delta_{p-1}}$, we define:
|
|
\[
|
|
T_{\pi} = \big(\pi_1(\mathbf{w}_1)^T \| E_{\mathbf{b}}(\mathbf{w}_2)^T \| \pi_3(\mathbf{w}_3)^T\big)^T.
|
|
\]
|
|
\end{itemize}
|
|
|
|
By inspection, it can be seen that the properties in~(\ref{eq:zk-equivalence}) are satisfied, as desired. As a result, we can obtain the required argument system by running the protocol in \cref{sse:stern-abstraction} with common input $(\mathbf{P}, \mathbf{v})$ and prover's input $\mathbf{x}$.
|
|
|
|
%--------------------------------------------------
|
|
\subsection{Proving the Possession of a Signature on a Committed Value}\label{subsection:zk-for-signature}
|
|
We now describe how to derive the protocol for proving the possession of a signature on a committed value, that is used in Section~\ref{commit-sig}.
|
|
\begin{description}
|
|
\item[Common Input:] Matrices $\mathbf{A}, \{\mathbf{A}_j\}_{j=0}^\ell, \mathbf{D} \in \ZZ_q^{n \times m}$; $\{\mathbf{D}_k\in \ZZ_q^{2n \times 2m}\}_{k=0}^N$;$\mathbf{B}\in \ZZ_q^{n \times m}$; $\mathbf{G}_1 \in \mathbb{Z}_q^{n \times 2m}$;
|
|
$\mathbf{G}_0 \in \mathbb{Z}_q^{n \times \ell}$;\\
|
|
vectors
|
|
$ \{\mathbf{c}_{k,1}\}_{k=1}^N, \mathbf{c}_{\tau,1}, \mathbf{c}_{\mathbf{v}, 1}, \mathbf{c}_{s, 1} \in \ZZ_q^m$; $\{\mathbf{c}_{k,2}\}_{k=1}^N,\mathbf{c}_{\mathbf{v}, 2}, \mathbf{c}_{s,2} \in \ZZ_q^{2m}$; $\mathbf{c}_{\tau,2} \in \ZZ_q^\ell$; $\mathbf{u} \in \mathbb{Z}_q^n$.
|
|
|
|
\smallskip
|
|
|
|
\item[Prover's Input:] $\mathbf{v} = \left(
|
|
\begin{array}{c}
|
|
\mathbf{v}_1 \\
|
|
\mathbf{v}_2 \\
|
|
\end{array}
|
|
\right)
|
|
$, where $\mathbf{v}_1, \mathbf{v}_2\in [-\beta, \beta]^m$ and $\beta = \sigma\cdot \omega(\log m)$ - the infinity norm bound of signatures; $\tau \in \{0,1\}^\ell$; $\mathbf{s} \in [-(p-1), (p-1)]^{2m}$;
|
|
|
|
\smallskip
|
|
$\mathfrak{m} = (\mathfrak{m}_1^T \| \ldots \| \mathfrak{m}_N^T)^T \in \mathsf{CorEnc}(mN)$; $\{\mathbf{s}_{k}\}_{k=1}^N$, $\mathbf{s}_{\mathbf{v}}$, $\mathbf{s}_0$, $\mathbf{s}_\tau \in [-B,B]^n$;
|
|
|
|
\smallskip
|
|
$\{\mathbf{e}_{k,1}\}_{k=1}^N$, $\mathbf{e}_{\mathbf{v}, 1}$, $\mathbf{e}_{0,1}$, $\mathbf{e}_{\tau,1} \in [-B,B]^m$;
|
|
$\{\mathbf{e}_{k,2}\}_{k=1}^N, \mathbf{e}_{0,2},\mathbf{e}_{\mathbf{v},2} \in [-B,B]^{2m}$;
|
|
|
|
\smallskip
|
|
$\mathbf{e}_{\tau,2} \in [-B,B]^\ell$.
|
|
\end{description}
|
|
|
|
\textbf{Prover's Goal:} Convince the verifier in \textsf{ZK} that:
|
|
\begin{eqnarray}\label{equation:R-sign-signature}
|
|
\hspace*{-5pt}
|
|
\mathbf{A}\hspace*{-1.5pt}\cdot\hspace*{-1.5pt}\mathbf{v}_1 + \mathbf{A}_0\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{v}_2 + \sum_{i=1}^\ell \mathbf{A}_i\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \tau[i] \mathbf{v}_2 - \mathbf{D}\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathsf{bin}(\mathbf{D}_0\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{s} + \sum_{k=1}^N \mathbf{D}_i\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathfrak{m}_k) = \mathbf{u} \bmod q,
|
|
\end{eqnarray}
|
|
and that (modulo $q$)
|
|
\begin{eqnarray}\label{equation:R-sign-ciphertext}
|
|
\begin{cases}
|
|
\forall k\in [N]: \mathbf{c}_{k,1}= \mathbf{B}^T\cdot\mathbf{s}_{k} + \mathbf{e}_{k,1} ; \mathbf{c}_{k,2}= \mathbf{G}_1^T\cdot \mathbf{s}_{k} + \mathbf{e}_{k,2} + \lfloor q/2 \rfloor\cdot \mathfrak{m}_k ; \\
|
|
|
|
\mathbf{c}_{\mathbf{v}, 1} = \mathbf{B}^T\cdot \mathbf{s}_{\mathbf{v}} + \mathbf{e}_{\mathbf{v},1} ; \\
|
|
\mathbf{c}_{\mathbf{v},2} = \mathbf{G}_1^T \cdot \mathbf{s}_{\mathbf{v}} +\mathbf{e}_{\mathbf{v},2}+ \lfloor\frac{q}{p}\rfloor \cdot \mathbf{v} \\
|
|
\hphantom{\mathbf{c}_{\mathbf{v},2}} = \mathbf{G}_1^T \cdot \mathbf{s}_{\mathbf{v}} + \mathbf{e}_{\mathbf{v},2} + \left(
|
|
\begin{array}{c}
|
|
\lfloor\frac{q}{p}\rfloor \mathbf{I}_m \\
|
|
\mathbf{0}\\
|
|
\end{array}
|
|
\right)\cdot \mathbf{v}_1
|
|
+ \left(
|
|
\begin{array}{c}
|
|
\mathbf{0}\\
|
|
\lfloor\frac{q}{p}\rfloor \mathbf{I}_m \\
|
|
\end{array}
|
|
\right)\cdot \mathbf{v}_2
|
|
; \\
|
|
|
|
|
|
\mathbf{c}_{\mathbf{s}, 1}= \mathbf{B}^T\cdot \mathbf{s}_0 + \mathbf{e}_{0,1} ; \hspace*{5pt}\mathbf{c}_{\mathbf{s},2}= \mathbf{G}_1^T\cdot \mathbf{s}_0 + \mathbf{e}_{0,2} + \lfloor q/p \rfloor\cdot \mathbf{s} ; \\
|
|
|
|
\mathbf{c}_{\tau,1} = \mathbf{B}^T\cdot \mathbf{s}_\tau + \mathbf{e}_{\tau,1} ; \hspace*{2.5pt} \mathbf{c}_{\tau,2}= \mathbf{G}_0^T\cdot \mathbf{s}_\tau + \mathbf{e}_{\tau,2} + \lfloor q/2 \rfloor\cdot \tau .
|
|
\end{cases}
|
|
\end{eqnarray}
|
|
$~$ \\
|
|
We proceed in two steps.
|
|
\medskip \smallskip
|
|
|
|
\textbf{Step 1:} \emph{Transforming the equations in~(\ref{equation:R-sign-signature}) and~(\ref{equation:R-sign-ciphertext}) into a unified one of the form $\mathbf{P}\cdot \mathbf{x} = \mathbf{c} \bmod q$, where $\|\mathbf{x}\|_\infty =1$ and $\mathbf{x} \in \mathsf{VALID}$ - a ``specially-designed'' set.}
|
|
|
|
Note that, if we let $\mathbf{y} = \mathsf{bin}(\mathbf{D}_0\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{s} + \sum_{k=1}^N \mathbf{D}_i\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathfrak{m}_k) \in \{0,1\}^{m}$, then we have $\mathbf{H}_{2n \times m}\cdot \mathbf{y} = \mathbf{D}_0\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{s} + \sum_{k=1}^N \mathbf{D}_i\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathfrak{m}_k \bmod q$, and~(\ref{equation:R-sign-signature}) can be equivalently written as:
|
|
\begin{eqnarray*}\label{equation:R-sign-signature-2}
|
|
\hspace*{-10pt}
|
|
\left(
|
|
\begin{array}{c}
|
|
\mathbf{A} \\
|
|
\mathbf{0} \\
|
|
\end{array}
|
|
\right)\hspace*{-1.5pt}\cdot\hspace*{-1.5pt}
|
|
\mathbf{v}_1 +
|
|
\left(
|
|
\begin{array}{c}
|
|
\mathbf{A}_0 \\
|
|
\mathbf{0} \\
|
|
\end{array}
|
|
\right)\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{v}_2 &+&
|
|
\sum_{i=1}^\ell \left(
|
|
\begin{array}{c}
|
|
\mathbf{A}_i \\
|
|
\mathbf{0} \\
|
|
\end{array}
|
|
\right)\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \tau[i] \mathbf{v}_2 +
|
|
\left(
|
|
\begin{array}{c}
|
|
\mathbf{0} \\
|
|
\mathbf{D}_0 \\
|
|
\end{array}
|
|
\right)\cdot \mathbf{s} + \left(
|
|
\begin{array}{c}
|
|
-\mathbf{D} \\
|
|
-\mathbf{H}_{2n \times m} \\
|
|
\end{array}
|
|
\right)\cdot \mathbf{y} \\
|
|
&+&\left(
|
|
\begin{array}{c}
|
|
\mathbf{0} \\
|
|
\mathbf{D}_1 | \ldots | \mathbf{D}_N \\
|
|
\end{array}
|
|
\right)\cdot \mathfrak{m} = \left(
|
|
\begin{array}{c}
|
|
\mathbf{u} \\
|
|
\mathbf{0}^{2n} \\
|
|
\end{array}
|
|
\right) ~\bmod q.
|
|
\end{eqnarray*}
|
|
Next, we use linear algebra to combine this equation and~(\ref{equation:R-sign-ciphertext}) into (modulo $q$):
|
|
\begin{align}\label{equation:R-sign-almost}
|
|
\hspace*{-10pt}
|
|
\mathbf{F}\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{v}_1 \hspace*{-1.5pt}+\hspace*{-1.5pt} \mathbf{F}_0\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{v}_2 \hspace*{-1.5pt}+\hspace*{-1.5pt} \sum_{i=1}^\ell \mathbf{F}_i \hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \tau[i]\mathbf{v}_2 + \mathbf{M}_1 \hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \tau \hspace*{-1.5pt}+\hspace*{-1.5pt} \mathbf{M}_2\hspace*{-1.5pt}\cdot\hspace*{-1.5pt}\mathbf{y} + \mathbf{M}_3\hspace*{-1.5pt}\cdot\hspace*{-1.5pt}\mathfrak{m}
|
|
\hspace*{-1.5pt}+ \hspace*{-1.5pt}\mathbf{M}_4 \hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{s} \hspace*{-2pt}+\hspace*{-2pt} \mathbf{M}_5\hspace*{-1.5pt}\cdot\hspace*{-1.5pt} \mathbf{e} \hspace*{-2pt}=\hspace*{-2pt} \mathbf{c},
|
|
\end{align}
|
|
where, for dimensions $D = \ell + 3n + 7m + 3mN$ and $L_0 = D + nN$,
|
|
\begin{itemize}
|
|
\item Matrices $\mathbf{F}, \mathbf{F}_0, \mathbf{F}_1, \ldots, \mathbf{F}_\ell \in \mathbb{Z}_q^{D \times m}$, $\mathbf{M}_1 \in \mathbb{Z}_q^{D \times \ell}$, $\mathbf{M}_2 \in \mathbb{Z}_q^{D \times m}$, $\mathbf{M}_3 \in \mathbb{Z}_q^{D \times 2mN}$, $\mathbf{M}_4 \in \mathbb{Z}_q^{D \times 2m}$, $\mathbf{M}_5 \in \mathbb{Z}_q^{D \times L_0}$ and vector $\mathbf{c} \in \mathbb{Z}_q^D$ are built from the public input.
|
|
\item Vector $\mathbf{e} = \big(\hspace*{1pt}\mathbf{s}_1^T \hspace*{1pt}\|\hspace*{1pt} \ldots \hspace*{1pt}\|\hspace*{1pt} \mathbf{s}_N^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{s}_{\mathbf{v}}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{s}_0^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{s}_\tau^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{1,1}^T \hspace*{1pt}\|\hspace*{1pt} \ldots \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{N,1}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{\mathbf{v},1}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{0,1}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{\tau, 1}^T \hspace*{1pt}\| \\
|
|
~~~~~~~~~~~~~~~~~\|\hspace*{1pt}\mathbf{e}_{1,2}^T \hspace*{1pt}\|\hspace*{1pt} \ldots \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{N,2}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{0,2}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{\mathbf{v},2}^T \hspace*{1pt}\|\hspace*{1pt} \mathbf{e}_{\tau,2}^T\hspace*{1pt}\big)^T \in [-B,B]^{L_0}$.
|
|
\end{itemize}
|
|
|
|
Now we further transform~\eqref{equation:R-sign-almost} using the techniques from \cref{sse:stern-abstraction}. Specifically, we form the following:
|
|
\[
|
|
\begin{cases}
|
|
\mathsf{DecExt}_{m, \beta}(\mathbf{v}_1) \rightarrow \hat{\mathbf{v}}_1 \in \mathsf{B}^3_{m\delta_\beta}; \hspace*{5pt}\mathsf{DecExt}_{m, \beta}(\mathbf{v}_2)\rightarrow \hat{\mathbf{v}}_2 \in \mathsf{B}^3_{m\delta_\beta}; \\[2.5pt]
|
|
{\mathbf{F}}' = \big[\mathbf{F} \cdot \widehat{\mathbf{K}}_{m, \beta} | \mathbf{F}_0 \cdot \widehat{\mathbf{K}}_{m, \beta} | \mathbf{F}_1 \cdot \widehat{\mathbf{K}}_{m, \beta} | \ldots | \mathbf{F}_\ell \cdot \widehat{\mathbf{K}}_{m, \beta} | \mathbf{0}^{D \times 3m\delta_\beta \ell}\big] \in \mathbb{Z}_q^{D \times 3m\delta_\beta(2\ell+2)}; \\[2.5pt]
|
|
\mathsf{Ext}_{2\ell}(\tau) \rightarrow \hat{\tau} = (\tau[1], \ldots, \tau[\ell], \ldots, \tau[2\ell])^T \in \mathsf{B}^2_{\ell}; \hspace*{2.5pt}\mathbf{M}'_1 = [\mathbf{M}_1 | \mathbf{0}^{D \times \ell}] \in \mathbb{Z}_q^{D \times 2\ell};\\[2.5pt]
|
|
\mathsf{Ext}_{2m}(\mathbf{y})\rightarrow \hat{\mathbf{y}} \in \mathsf{B}^2_{m}; \hspace*{2.5pt}\mathbf{M}'_2 = [\mathbf{M}_2 | \mathbf{0}^{D \times m}] \in \ZZ_q^{D \times 2m }; \\[2.5pt]
|
|
\mathsf{DecExt}_{2m, p-1}(\mathbf{s}) \rightarrow \hat{\mathbf{s}} \in \mathsf{B}^3_{2m\delta_{p-1}}; \hspace*{2.5pt} \mathbf{M}'_4 = \mathbf{M}_4 \cdot \widehat{\mathbf{K}}_{2m, p-1} \in \mathbb{Z}_q^{D \times 6m\delta_{p-1}}; \\[2.5pt]
|
|
\mathsf{DecExt}_{L_0, B}(\mathbf{e}) \rightarrow \hat{\mathbf{e}} \in \mathsf{B}^3_{L_0\delta_{B}}; \hspace*{2.5pt}
|
|
\mathbf{M}'_5 = \mathbf{M}_5 \cdot \widehat{\mathbf{K}}_{L_0, B} \in \mathbb{Z}_q^{D \times 3L_0\delta_B}.
|
|
\end{cases}
|
|
\]
|
|
Now, let $L = 3m\delta_\beta(2\ell+2) + 2\ell + 2m + 2mN + 6m\delta_{p-1} + 3L_0\delta_B$, and construct matrix
|
|
$\mathbf{P} = \big[\hspace*{1pt}\mathbf{F}' \hspace*{1pt}|\hspace*{1pt} \mathbf{M}'_1 \hspace*{1pt}| \hspace*{1pt}\mathbf{M}'_2 \hspace*{1pt}|\hspace*{1pt} \mathbf{M}_3\hspace*{1pt}|\hspace*{1pt} \mathbf{M}'_4\hspace*{1pt}| \hspace*{1pt} \mathbf{M}'_5 \hspace*{1pt}\big] \in \ZZ_q^{D \times L}$ and vector
|
|
\[
|
|
\mathbf{x} = \big(\hspace*{1.5pt}\hat{\mathbf{v}}_1^T\hspace*{1.5pt} \|\hspace*{1.5pt} \hat{\mathbf{v}}_2^T \hspace*{1.5pt}\| \hspace*{1.5pt}\tau[1]\hat{\mathbf{v}}_2^T\hspace*{1.5pt} \| \ldots \|\hspace*{1.5pt} \tau[\ell]\hat{\mathbf{v}}_2^T\hspace*{1.5pt}\| \ldots \| \hspace*{1.5pt}\tau[2\ell]\hat{\mathbf{v}}_2^T\hspace*{1.5pt}\| \hspace*{1.5pt} \hat{\tau}^T \hspace*{1.5pt}\| \hspace*{1.5pt}\hat{\mathbf{y}}^T\hspace*{1.5pt} \|\hspace*{1.5pt} \mathfrak{m}^T \hspace*{1.5pt}\|\hspace*{1.5pt} \hat{\mathbf{s}}^T\hspace*{1.5pt}\| \hspace*{1.5pt} \hat{\mathbf{e}}^T\hspace*{1.5pt}\big)^T,
|
|
\]
|
|
then we will obtain the equation $\mathbf{P}\cdot \mathbf{x} = \mathbf{c} \bmod q$.
|
|
|
|
Before going on, we define $\mathsf{VALID}$ as the set of
|
|
$\mathbf{w} \in \{-1,0,1\}^L$ of the form:
|
|
\[
|
|
\mathbf{w} = \big(\mathbf{w}_{1}^T \| \mathbf{w}_2^T \| g_1 \mathbf{w}_2^T\| \ldots \| g_{2\ell}\mathbf{w}_2^T \| \mathbf{g}^T\| \mathbf{w}_3^T\| \mathbf{w}_4^T \| \mathbf{w}_5^T \| \mathbf{w}_6^T\big)^T
|
|
\]
|
|
for some $\mathbf{w}_1, \mathbf{w}_2 \in \mathsf{B}^3_{m\delta_\beta}$, $\mathbf{g} = (g_1, \ldots, g_{2\ell}) \in \mathsf{B}_{2\ell}$, $\mathbf{w}_3 \in \mathsf{B}^2_{m}$, $\mathbf{w}_4 \in \mathsf{CorEnc}(mN)$, $\mathbf{w}_5 \in \mathsf{B}^3_{2m\delta_{p-1}}$, and $\mathbf{w}_6 \in \mathsf{B}^3_{L_0\delta_B}$.
|
|
It can be checked that the constructed vector $\mathbf{x}$ belongs to this tailored set $\mathsf{VALID}$.\\
|
|
|
|
|
|
|
|
|
|
|
|
\textbf{Step 2:} \emph{Specifying the set $\mathcal{S}$ and permutations of $L$ elements $\{T_\pi: \pi \in \mathcal{S}\}$ for which the conditions in~(\ref{eq:zk-equivalence}) hold.}
|
|
|
|
\begin{itemize}
|
|
\item Define $\mathcal{S} = \mathcal{S}_{3m\delta_\beta} \times \mathcal{S}_{3m\delta_\beta} \times \mathcal{S}_{2\ell} \times\mathcal{S}_{2m} \times \{0,1\}^{mN}\times \mathcal{S}_{6m\delta_{p-1}} \times \mathcal{S}_{3L_0\delta_B}$. \medskip
|
|
\item For $\pi = (\phi, \psi, \gamma, \rho, \mathbf{b}, \eta, \xi) \in \mathcal{S}$ and $\mathbf{z} = \big(\mathbf{z}_0^1 \| \mathbf{z}_0^2 \| \mathbf{z}_1 \| \ldots \| \mathbf{z}_{2\ell} \| \mathbf{g} \| \mathbf{t}_1 \| \mathbf{t}_2 \|\mathbf{t}_3 \| \mathbf{t}_4\big) \in \mathbb{Z}_q^L$,
|
|
where ${\mathbf{z}_0^1}, {\mathbf{z}_0^2}, \mathbf{z}_1, \ldots, \mathbf{z}_{2\ell} \in \mathbb{Z}_q^{3m\delta_\beta}$, $\mathbf{g} \in \mathbb{Z}_q^{2\ell}$, $\mathbf{t}_1\in \mathbb{Z}_q^{2m}$, $\mathbf{t}_2 \in \mathbb{Z}_q^{2mN}$, $\mathbf{t}_3 \in \mathbb{Z}_q^{6m\delta_{p-1}}$, and $\mathbf{t}_4 \in \mathbb{Z}_q^{3L_0\delta_B}$, we define:
|
|
\begin{eqnarray*}
|
|
\hspace*{-15pt}
|
|
T_{\pi}(\mathbf{z}) = \big(\phi(\mathbf{z}_0^1)^T\hspace*{1pt} \| \psi(\mathbf{z}_0^2)^T \hspace*{1pt}\| \psi(\mathbf{z}_{\gamma(1)})^T \hspace*{1pt}\| \ldots \| \psi(\mathbf{z}_{\gamma(2\ell)})^T \hspace*{1pt}\| \gamma(\mathbf{g})^T\hspace*{1pt} \| \\
|
|
~~~~~~~~\|\rho(\mathbf{t}_1)^T \| E_{\mathbf{b}}(\mathbf{t}_2)^T \hspace*{1pt}\| \eta(\mathbf{t}_3)^T \| \xi(\mathbf{t}_4)^T\hspace*{1pt}\big)^T
|
|
\end{eqnarray*}
|
|
as the permutation that transforms $\mathbf{z}$ as follows:
|
|
\begin{enumerate}
|
|
\item It rearranges the order of the $2\ell$ blocks $\mathbf{z}_1, \ldots, \mathbf{z}_{2\ell}$ according to $\gamma$.
|
|
\item It then {permutes} block $\mathbf{z}_0^1$ according to $\phi$, blocks $\mathbf{z}_0^2$, $\{\mathbf{z}_i\}_{i=1}^{2\ell}$ according to~$\psi$, block $\mathbf{g}$ according to $\gamma$, block $\mathbf{t}_1$ according to $\rho$, block $\mathbf{t}_2$ according to $E_{\mathbf{b}}$, block $\mathbf{t}_3$ according to~$\eta$, and block $\mathbf{t}_4$ according to $\xi$.
|
|
\end{enumerate}
|
|
\end{itemize}
|
|
It can be check that~(\ref{eq:zk-equivalence}) holds. Therefore, we can obtain a statistical \textsf{ZKAoK} for the given relation by running the protocol in \cref{sse:stern-abstraction}.
|
|
|
|
\subsection{The Underlying ZKAoK for the Group Signature Scheme}\label{subsection:zk-for-group-signature}
|
|
The argument system upon which our group signature scheme is built can be summarized as follows.
|
|
\begin{description}
|
|
\item[Common Input:] Matrices $\mathbf{A}, \{\mathbf{A}_j\}_{j=0}^\ell, \mathbf{B} \in \mathbb{Z}_q^{n \times m}$, $\mathbf{D}_0, \mathbf{D}_1 \in \mathbb{Z}_q^{2n \times 2m}$, $\mathbf{F} \in \mathbb{Z}_q^{4n \times 4m}$, $\mathbf{H}_{2n \times m} \hspace*{-1.5pt}\in\hspace*{-1.5pt} \ZZ_q^{2n \times m}$, $\mathbf{H}_{4n \times 2m} \hspace*{-1.5pt}\in\hspace*{-1.5pt} \ZZ_q^{4n \times 2m}$, $\mathbf{G}_0 \hspace*{-1.5pt}\in\hspace*{-1.5pt} \mathbb{Z}_q^{n \times 2m}$; vectors $\mathbf{u} \hspace*{-1.5pt}\in\hspace*{-1.5pt} \mathbb{Z}_q^n$, $\mathbf{c}_1 \hspace*{-1.5pt}\in\hspace*{-1.5pt} \mathbb{Z}_q^m$, $\mathbf{c}_2 \hspace*{-1.5pt}\in \hspace*{-1.5pt}\mathbb{Z}_q^{2m}$. \smallskip
|
|
\item [Prover's Input:] $\mathbf{z} \in [-\beta,\beta]^{4m}$, $\mathbf{y} \in \{0,1\}^{2m}$, $\mathbf{w} \in \{0,1\}^m$, $\mathbf{d}_1, \mathbf{d}_2 \in [-\beta, \beta]^m$, $\mathbf{s} \in [-\beta,\beta]^{2m}$, $\mathrm{id} = (\mathrm{id}[1], \ldots, \mathrm{id}[\ell])^T \in \{0,1\}^\ell$,
|
|
|
|
$\mathbf{e}_0 \in [-B,B]^n$, $\mathbf{e}_1 \in [-B,B]^m$, $\mathbf{e}_2 \in [-B,B]^{2m}$. \smallskip
|
|
\item[Prover's Goal:] Convince the verifier in \textsf{ZK} that
|
|
\end{description}
|
|
\[
|
|
\begin{cases}
|
|
\mathbf{F}\cdot \mathbf{z} = \mathbf{H}_{4n\times 2m}\cdot \mathbf{y}\bmod q; \hspace*{5pt} \mathbf{H}_{2n \times m}\cdot \mathbf{w} = \mathbf{D}_0 \cdot \mathbf{y} + \mathbf{D}_1 \cdot \mathbf{s} \bmod q; \\
|
|
\mathbf{A}\cdot \mathbf{d}_1 + \mathbf{A}_0 \cdot \mathbf{d}_2 + \sum_{j=1}^\ell \mathbf{A}_j \cdot (\mathrm{id}[j]\cdot \mathbf{d}_2) - \mathbf{D} \cdot \mathbf{w} = \mathbf{u} \bmod q;\\
|
|
\mathbf{c}_1 = \mathbf{B}^T\cdot \mathbf{e}_0 + \mathbf{e}_1 \bmod q; \hspace*{5pt} \mathbf{c}_2 = \mathbf{G}_0^T\cdot \mathbf{e}_0 + \mathbf{e}_2 + \lfloor q/2\rfloor\cdot \mathbf{y} \bmod q.
|
|
\end{cases}
|
|
\]
|
|
|
|
Using the same strategy as in Sections~\ref{subsection:zk-for-commitments} and~\ref{subsection:zk-for-signature}, we can derive a statistical \textsf{ZKAoK} for the above relation from the protocol in Section~\ref{sse:stern-abstraction}. As the transformations are similar to those in Section~\ref{subsection:zk-for-signature}, we only sketch main points.
|
|
|
|
In the first step, we combine the given equations to an equation of the form:
|
|
\[
|
|
\mathbf{M}\cdot \left(
|
|
\begin{array}{c}
|
|
\mathbf{d}_1 \\
|
|
\mathbf{s} \\
|
|
\mathbf{z} \\
|
|
\end{array}
|
|
\right) + \mathbf{M}_0 \cdot \mathbf{d}_2 + \sum_{j=1}^\ell \mathbf{M}_j(\mathrm{id}[j]\mathbf{d}_2) + \mathbf{M}' \cdot \left(
|
|
\begin{array}{c}
|
|
\mathbf{w} \\
|
|
\mathbf{y} \\
|
|
\end{array}
|
|
\right) + \mathbf{M}'' \cdot \left(
|
|
\begin{array}{c}
|
|
\mathbf{e}_0 \\
|
|
\mathbf{e}_1 \\
|
|
\mathbf{e}_2 \\
|
|
\end{array}
|
|
\right) = \mathbf{v} \bmod q,
|
|
\]
|
|
where matrices $\mathbf{M}, \mathbf{M}_0, \ldots, \mathbf{M}_\ell, \mathbf{M}', \mathbf{M}''$ and vector $\mathbf{v}$ are built from the input.
|
|
|
|
We then apply the techniques of \cref{sse:stern-abstraction} for %the vectors
|
|
$\mathbf{x}_0 = (\mathbf{d}_1^T \| \mathbf{s}^T \| \mathbf{z}^T)^T \in [-\beta, \beta]^{7m}$, $\mathbf{d}_2 \in [-\beta,\beta]^m$; $\mathbf{x}_1 = (\mathbf{w}^T \| \mathbf{y}^T)^T\in \{0,1\}^{3m}$; and $\mathbf{x}_2 = (\mathbf{e}_0^T \| \mathbf{e}_1^T \| \mathbf{e}_2^T)^T \in [-B,B]^{n + 3m}$. This allows us to obtain a unified equation $\mathbf{P}\cdot \mathbf{x} = \mathbf{v} \bmod q$, and to define the sets $\mathsf{VALID}$, $\mathcal{S}$, and permutations $\{T_\pi: \pi \in \mathcal{S}\}$ so that the conditions in~(\ref{eq:zk-equivalence}) hold, in a similar manner as in Section~\ref{subsection:zk-for-signature}.
|