Keywords

1 Introduction

It is commonly understood that cryptographic assumptions play a crucial role in the development of secure, efficient protocols with strong functionalities. For instance, upon referring to the rapid development of pairing-based cryptography, X. Boyen [8] says that “it has been supported, in no small part, by a dizzying array of tailor-made cryptographic assumptions”. Although this may be a reasonable price to pay for constructing new primitives or improve their efficiency, one should not lose sight of the ideal of using standard and simple assumptions. This is an important aspect of provable security. Indeed, Goldreich [16], for instance, cites “having clear definitions of one’s assumptions” as one of the three main ingredients of good cryptographic practice.

There are many aspects to this goal. Not only it is important to use clearly defined assumptions, but also to understand the relations between them: to see, for example, if two assumptions are equivalent or one is weaker than the other. Additionally, the definitions should allow to make accurate security claims. For instance, although technically it is correct to say that unforgeability of the Waters’ signature scheme [42] is implied by the \(\mathsf {DDH}\) Assumption, defining the \(\mathsf {CDH}\) Assumption allows to make a much more precise security claim.

A notable effort in reducing the “dizzying array” of cryptographic assumptions is the work of Escala et al. [11]. They put forward a new family of decisional assumptions in a prime order group \(\mathbbm {G}\), the Matrix Diffie-Hellman Assumption (\(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}{}\)). It says that, given some matrix \(\mathbf {{A}} \in \mathbbm {Z}_q^{\ell \times k}\) sampled from some distribution \(\mathcal {D}_{\ell ,k}\), it is hard to decide membership in \(\mathop {\text {Im}}\mathbf {{A}}\), the subspace spanned by the columns of \(\mathbf {{A}}\), in the exponent. Rather than as new assumption, it should be seen as an algebraic framework for decisional assumptions which includes as a special case the widely used \(k\text{- }\mathsf {Lin}\) family.

This framework has some obvious conceptual advantages. For instance, it allows to explain all the members of the \(k\text{- }\mathsf {Lin}\) assumption family (and also others, like the uniform assumption, appeared previously in [13, 14, 41]) as a single assumption and unify different constructions of the same primitive in the literature (e.g., the Naor-Reingold PRF [36] and the Lewko-Waters PRF [29] are special cases of the same construction instantiated with the \(1\text{- }\mathsf {Lin}\) and the \(2\text{- }\mathsf {Lin}\) Assumption, respectively). Another of its advantages is that it avoids arbitrary choices and instead points out to a trade-off between efficiency and security (a scheme based on any \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}{}\) Assumption can be instantiated with many different assumptions, some leading to stronger security guarantees and others leading to more efficient schemes). But follow-up work has also illustrated other possibly less obvious advantages. For instance, Herold et al. [21] have used the Matrix Diffie-Hellman abstraction to extend the model of composite-order to prime-order transformation of Freeman [13] and to derive efficiency improvements which were proven to be impossible in the original model.Footnote 1 We believe this illustrates that the benefits of conceptual clarity can translate into concrete improvements as well.

The security notions for cryptographic protocols can be classified mainly in hiding and unforgeability ones. The former typically appear in encryption schemes and commitments and the latter in signature schemes and soundness in zero-knowledge proofs. Although it is theoretically possible to base the hiding property on computational problems, most of the practical schemes achieve this notion either information theoretically or based on decisional assumptions, at least in the standard model. Likewise, unforgeability naturally comes from computational assumptions (typically implied by stronger, decisional assumptions). Thus, a natural question is if one can find a computational analogue of their \(\mathsf {MDDH}{}\) Assumption which can be used in “unforgeability type” of security notions.

Most computational problems considered in the literature are search problems with a unique solution like the discrete logarithm or \(\mathsf {CDH}\). But unforgeability actually means the inability to produce one among many solutions to a given problem (e.g., in many signature schemes or zero knowledge proofs). Thus, unforgeability is more naturally captured by a flexible computational problem, namely, a problem which admits several solutionsFootnote 2. This maybe explains why several new flexible assumptions have appeared recently when considering “unforgeability-type” security notions in structure-preserving cryptography [2]. Thus a useful computational analogue of the \(\mathsf {MDDH}{}\) Assumption should not only consider problems with a unique solution but also flexible problems which can naturally capture this type of security notions.

1.1 Our Results

In the following \(\mathcal {G}=(\mathbbm {G},q,\mathcal {P})\), being \(\mathbbm {G}\) some group in additive notation of prime order q generated by \(\mathcal {P}\), that is, the elements of \(\mathbbm {G}\) are \(\mathcal {Q}=a\mathcal {P}\) where \(a \in \mathbbm {Z}_q\). They will be denoted as \([a]:=a \mathcal {P}\). This notation naturally extends to vectors and matrices as \([{\varvec{v}}]=(v_1 \mathcal {P},\ldots ,v_n \mathcal {P})\) and \([\mathbf {{A}}]=(A_{ij} \mathcal {P})\).

Computational Matrix Assumptions. In our first attempt to design a computational analogue of the \(\mathsf {MDDH}{}\) Assumption, we introduce the Matrix Computational DH Assumption, (\(\mathsf {MCDH}\)) which says that, given a uniform vector \([{\varvec{v}}] \in \mathbbm {G}^k\) and some matrix \([\mathbf {{A}}]\), \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\) for \(\ell >k\), it is hard to extend \([{\varvec{v}}]\) to a vector in \(\mathbbm {G}^\ell \) in the image of \([\mathbf {{A}}]\), \(\mathop {\text {Im}}[\mathbf {{A}}]\). Although this assumption is natural and is weaker than the \(\mathsf {MDDH}{}\) one, we argue that it is equivalent to \(\mathsf {CDH}\).

We then propose the Kernel Matrix DH Assumption (\(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\)). This new flexible assumption states that, given some matrix \([\mathbf {{A}}]\), \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\) for some \(\ell >k\), it is hard to find a vector \([{\varvec{v}}] \in \mathbbm {G}^{\ell }\) in the kernel of \(\mathbf {{A}}^\top \). We observe that for some special instances of \(\mathcal {D}_{\ell ,k}\), this assumption has appeared in the literature in [2, 18, 19, 27, 32] under different names, like Simultaneous Pairing, Simultaneous Double Pairing (SDP in the following), Simultaneous Triple Pairing, 1-Flexible CDH, 1-Flexible Square CDH. Thus, the new \(\mathsf {KerMDH}\) Assumption allows us to organize and give a unified view on several useful assumptions. This suggests that the \(\mathsf {KerMDH}\) Assumption (and not the \(\mathsf {MCDH}\) one) is the right computational analogue of the \(\mathsf {MDDH}{}\) framework. Indeed, for any matrix distribution the \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) Assumption implies the corresponding \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\) Assumption. As a unifying algebraic framework, it offers the advantages mentioned above: it highlights the algebraic structure of any construction based on it, and it allows writing many instantiations of a given scheme in a compact way.

The Power of Kernel Assumptions. At Eurocrypt 2015, our \(\mathsf {KerMDH}\) Assumptions were applied to design simpler QA-NIZK proofs of membership in linear spaces [26]. They have also been used to give more efficient constructions of structure preserving signatures [25], to generalize and simplify the results on quasi-adaptive aggregation of Groth-Sahai proofs [17] (given originally in [24]) and to construct a tightly secure QA-NIZK argument for linear subspaces with unbounded simulation soundness in [15]. The power of a \(\mathsf {KerMDH}\) Assumption is that it allows to guarantee uniqueness. This has been used by Kiltz and Wee [26], for instance, to compile some secret key primitives to the public key setting. Indeed, Kiltz and Wee [26] modify a hash proof system (which is only designated verifier) to allow public verification (a QA-NIZK proof of membership). In a hash proof system for membership in some linear subspace of \(\mathbbm {G}^{n}\) spanned by the columns of some matrix \([\mathbf {{M}}]\), the public information is \([\mathbf {{M}}^{\top }\mathbf {{K}}]\), for some secret matrix \(\mathbf {{K}}\), and given the proof \([{\varvec{\pi }}]\) that \([{\varvec{y}}]\) is in the subspace, verification tests if \([{\varvec{\pi }}]\mathop {=}\limits ^{?}[\mathbf {{y}}^{\top } \mathbf {{K}}]\).

The core argument to compile this to a public key primitive is that given \(([\mathbf {{A}}],[\mathbf {{K}}\mathbf {{A}}])\), \(\mathbf {{A}} \leftarrow \mathcal {D}_{\ell ,k}\) and any pair \([{\varvec{y}}],[{\varvec{\pi }}]\), the previous test is equivalent to \(e([{\varvec{\pi }}^{\top }], [\mathbf {{A}}]) = e([{\varvec{y}}^{\top }], [\mathbf {{K}}\mathbf {{A}}])\), under the \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) Assumption. Indeed,

$$\begin{aligned}&e([{\varvec{\pi }}^{\top }], [\mathbf {{A}}]) = e([{\varvec{y}}^{\top }], [\mathbf {{K}}\mathbf {{A}}]) \Longleftrightarrow e([{\varvec{\pi }}^{\top }-{\varvec{y}}^{\top }\mathbf {{K}}], [\mathbf {{A}}]) =[ {\varvec{0}}] \mathop {\Longrightarrow }\limits ^{\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}{}}\nonumber \\[.5ex]&\qquad \Longrightarrow [\varvec{\pi }]=[{\varvec{y}}^{\top }\mathbf {{K}}]. \end{aligned}$$
(1)

That is, although potentially there are many possible proofs which satisfy the public verification equation (left hand side of Eq. (1)), the \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\) Assumption guarantees that only one of them is efficiently computable, so verification gives the same guarantees as in the private key setting (right hand side of Eq. (1)). This property is also used in a very similar way in [15] and also in the context of structure preserving signatures in [25]. In Sect. 5 we use it to argue that, of all the possible openings of a commitment, only one is efficiently computable, i.e. to prove computational soundness of a commitment scheme. Moreover, some previous works, notably in the design of structure preserving cryptographic primitives [13, 31], implicitly used this property for one specific \(\mathsf {KerMDH}\) Assumption: the Simultaneous (Double) Pairing Assumption.

On the other hand, we have already discussed the importance of having a precise and clear language when talking about cryptographic assumptions. This justifies the introduction of a framework specific to computational assumptions, because one should properly refer to the assumption on which security is actually based, rather than just saying “security is based on an assumption weaker than \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\)”. A part from being imprecise, a problem with such a statement is that might lead to arbitrary, not optimal choices. For example, the signature scheme of [30] is based on the SDP Assumption but a slight modification of it can be based on the \(\mathcal {L}_{2}\)-\(\mathsf {KerMDH}\) Assumption. If the security guarantee is “the assumption is weaker than \(2\text{- }\mathsf {Lin}\)” then the modified scheme achieves shorter public key and more efficient verification with no loss in security. Further, the claim that security is based on the \(\mathsf {MDDH}\) decisional assumptions when only computational ones are necessary might give the impression that a certain tradeoff is in place when this is not known to be the case. For instance, Jutla and Roy [24] construct constant-size QA-NIZK arguments of membership in linear spaces under what they call the “Switching Lemma”, which is proven under a certain \(\mathcal {D}_{k+1,k}\)-\(\mathsf {MDDH}\) Assumption. However, a close look at the proof reveals that in fact it is based on the corresponding \(\mathcal {D}_{k+1,k}\)-\(\mathsf {KerMDH}\) AssumptionFootnote 3. For these assumptions, prior to our work, it was unclear whether the choice of larger k gives any additional guarantees.

Strictly Increasing Families of Kernel Assumptions. An important problem is that it is not clear whether there are increasingly weaker families of \(\mathsf {KerMDH}\) Assumptions. That is, some decisional assumptions families parameterized by k like the \(k\text{- }\mathsf {Lin}\) Assumption are known to be strictly increasingly weaker. The proof of increasing hardness is more or less immediate and the term strictly follows from the fact that every two \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {MDDH}\) problems with \(\widetilde{k}<k\) are separated by an oracle computing a k-linear map. For the computational case, increasing hardness is also not too difficult, but nothing is known about strictly increasing hardness (see Fig. 1). This means that, as opposed to the decisional case, prior to our work, for protocols based on \(\mathsf {KerMDH}\) Assumptions there was no-known tradeoff between larger k (less efficiency) and security.

In this paper, we prove that the families of matrix distributions in [11], \(\mathcal {U}_{\ell ,k}\), \(\mathcal {L}_k\), \(\mathcal {SC}_k\), \(\mathcal {C}_k\) and \(\mathcal {RL}_k\), as well as a new distribution we propose in Sect. 6, the circulant family \(\mathcal {CI}_{k,d}\), define families of kernel problems with increasing hardness. For this we show a tight reduction from the smaller to the larger problems in each family. Our main result (Theorem 2) is to prove that the hardness of these problems is strictly increasing. For this, we prove that there is no black-box reduction from the larger to the smaller problems in the multilinear generic group model. These new results correspond to the dotted arrows in Fig. 1.

Fig. 1.
figure 1

Implication and separation results between Matrix Assumptions (dotted arrows correspond to the new results).

Having in mind that the computational problems we study in the paper are defined in a generic way, that is without specifying any particular group, the generic group approach arises naturally as the setting for the analysis of their hardness and reducibility relations. Otherwise, we would have to rely on specific properties of the representation of the elements of particular group families, not captured by the generic model.

The proof of Theorem 2 requires dealing with the notion of black-box reduction between flexible problems. A black-box reduction must work for any possible behavior of the oracle, but, contrary to the normal (unique answer) black-box reductions, here the oracle has to choose among the set of valid answers in every call. Ruling out the existence of a reduction implies that for any reduction there is an oracle behavior for which the reduction fails. This is specially subtle when dealing with multiple oracle calls. We think that the proof technique we introduce to deal with these issues can be considered as a contribution in itself and can potentially be used in future work.

Combining the black-box techniques and the generic group model is not new in the literature. For instance Dodis et al. [10] combine the black-box reductions and a generic model for the group \(\mathbbm {Z}_n^*\) to show some uninstantiability results for FDH-RSA signatures.

Theorem 2 supports the intuition that there is a tradeoff between the size of the matrix—which typically results in less efficiency—and the hardness of the \(\mathsf {KerMDH}\) Problems, and justifies the generalization of several protocols to different choices of k given in [17, 2426].

Applications. The discussion of our results given so far should already highlight some of the advantages of using the new Kernel family of assumptions and the power of these new assumptions, which have already been used in compelling applications in follow-up work in [17, 25, 26]. To further illustrate the usefulness of the new framework, we apply it to the study of trapdoor commitments. First, we revisit the Pedersen commitment [38] to vectors of scalars and its extension to vectors of group elements of Abe et al. [2] in bilinear groups. We unify these two constructions and we generalize to commit vectors of elements at each level \(\mathbbm {G}_r\), for any \(0 \le r \le m\) under the extension of \(\mathsf {KerMDH}\) Assumptions to the ideal m-graded encodings setting. In particular, when \(m=2\) we recover in a single construction as a special case both the original Pedersen and Abe et al. commitments.

The (generalized) Pedersen commitment maps vectors in \(\mathbbm {G}_r\) to vectors in \(\mathbbm {G}_{r+1}\), is perfectly hiding and computationally binding under any Kernel Assumption. In Sect. 5.2 we use it as a building block to construct a “group-to-group” commitment, which maps vectors in \(\mathbbm {G}_r\) to vectors in the same group \(\mathbbm {G}_{r}\). These commitments were defined in [3] because they are a good match to Groth-Sahai proofs. In [3], two constructions were given, one in asymmetric and the other in symmetric bilinear groups. Both are optimal in terms of commitment size and number of verification equations. Rather surprisingly, we show that both constructions in [3] are special instances of our group-to-group commitment for some specific matrix distributions.

A New Family of MDDH Assumptions of Optimal Representation Size. We also propose a new interesting family of Matrix distributions, the circulant matrix distribution, \(\mathcal {CI}_{k,d}\), which defines new \(\mathsf {MDDH}\) and \(\mathsf {KerMDH}\) assumptions. This family generalizes the Symmetric Cascade Distribution (\(\mathcal {SC}_{k}\)) defined in [11] to matrices of size \(\ell \times k\), \(\ell =k+d >k+1\). We prove that it has optimal representation size d independent of k among all matrix distributions of the same size. The case \(\ell > k+1\) typically arises when one considers commitments/encryption in which the message is a vector of group elements instead of a single group element and the representation size typically affects the size of the public parameters.

We prove the hardness of the \(\mathcal {CI}_{k,d}\)-\(\mathsf {KerMDH}\) Problem, by proving that the \(\mathcal {CI}_{k,d}\)-\(\mathsf {MDDH}\) Problem is generically hard in k-linear groups. Analyzing the hardness of a family of decisional problems (depending on a parameter k) can be rather involved, specially when an efficient k-linear map is supposed to exist. This is why in [11], the authors gave a practical criterion for generic hardness when \(\ell =k+1\) in terms of irreducibility of some polynomials involved in the description of the problem. This criterion was used then to prove the generic hardness of several families of \(\mathsf {MDDH}\) Problems. To analyze the generic hardness of the \(\mathcal {CI}_{k,d}\)-\(\mathsf {MDDH}\) Problem for any d, the techniques in [11] are not practical enough, and we need some extensions of these techniques for the case \(\ell >k+1\), recently introduced in [20]. However, we could not avoid the explicit computation of a large (but well-structured) Gröbner basis of an ideal associated to the matrix distribution. The new assumption can be used to instantiate the commitment schemes of Sect. 5 with shorter public parameters and improved efficiency.

2 Preliminaries

For \(\lambda \in \mathbbm {N}\), we write \(1^\lambda \) for the string of \(\lambda \) ones. For a set S, \(s \leftarrow S\) denotes the process of sampling an element s from S uniformly at random. For an algorithm \(\mathcal {A}\), we write \(z \leftarrow \mathcal {A}(x,y,\ldots )\) to indicate that \(\mathcal {A}\) is a (probabilistic) algorithm that outputs z on input \((x,y,\ldots )\). For any two computational problems \(\mathbb {P}_1\) and \(\mathbb {P}_2\) we recall that \(\mathbb {P}_1\Rightarrow \mathbb {P}_2\) denotes the fact that \(\mathbb {P}_1\) reduces to \(\mathbb {P}_2\), and then ‘\(\mathbb {P}_1\) is hard’ \(\Rightarrow \)\(\mathbb {P}_2\) is hard’. Thus, we will use ‘\(\Rightarrow \)’ both for computational problems and for the corresponding hardness assumptions.

Let \(\mathsf {Gen}\) denote a cyclic group instance generator, that is a probabilistic polynomial time (PPT) algorithm that on input \(1^\lambda \) returns a description \(\mathcal {G}=(\mathbbm {G},q,\mathcal {P})\) of a cyclic group \(\mathbbm {G}\) of order q for a \(\lambda \)-bit prime q and a generator \(\mathcal {P}\) of \(\mathbbm {G}\). We use additive notation for \(\mathbbm {G}\) and its elements are \(a\mathcal {P}\), for \(a\in \mathbbm {Z}_q\) and will be denoted as \([a]:=a\mathcal {P}\). The notation extends to vectors and matrices in the natural way as \([{\varvec{v}}]=(v_1 \mathcal {P},\ldots ,v_n \mathcal {P})\) and \([\mathbf {{A}}]=(A_{ij} \mathcal {P})\). For a matrix \(\mathbf {{A}}\in \mathbbm {Z}_q^{\ell \times k}\), \(\mathop {\text {Im}}\mathbf {{A}}\) denotes the subspace of \(\mathbbm {Z}_q^\ell \) spanned by the columns of \(\mathbf {{A}}\). Thus, \(\mathop {\text {Im}}[\mathbf {{A}}]\) is the corresponding subspace of \(\mathbbm {G}^\ell \).

2.1 Multilinear Maps

In the case of groups with a bilinear map, or more generally with a k-linear map for \(k\ge 2\), we consider a generator producing the tuple \((e_k,\mathbbm {G}_1,\mathbbm {G}_{k}, q, \mathcal {P}_1,\mathcal {P}_k)\), where \(\mathbbm {G}_1,\mathbbm {G}_{k}\) are cyclic groups of prime-order q, \(\mathcal {P}_i\) is a generator of \(\mathbbm {G}_i\) and \(e_k\) is a non-degenerate efficiently computable k-linear map \(e_k:\mathbbm {G}_1^k\rightarrow \mathbbm {G}_{k}\), such that \(e_k(\mathcal {P}_1,\ldots ,\mathcal {P}_1)=\mathcal {P}_{k}\). We actually consider graded encodings which offer a richer structure. For any fixed \(k \ge 1\), let \(\mathsf {MGen}_{k}\) be a PPT algorithm that on input \(1^\lambda \) returns a description of a graded encoding \(\mathcal {MG}_{k}=(e,\mathbbm {G}_1,\ldots , \mathbbm {G}_{k}, q, \mathcal {P}_1,\ldots ,\mathcal {P}_k)\), where \(\mathbbm {G}_1,\ldots ,\mathbbm {G}_{k}\) are cyclic groups of prime-order q, \(\mathcal {P}_i\) is a generator of \(\mathbbm {G}_i\) and e is a collection of non-degenerate efficiently computable bilinear maps \(e_{i,j}:\mathbbm {G}_i\times \mathbbm {G}_j\rightarrow \mathbbm {G}_{i+j}\), for \(i+j\le k\), such that \(e(\mathcal {P}_i,\mathcal {P}_j)=\mathcal {P}_{i+j}\). For simplicity we will omit the subindexes of e when they become clear from the context. Sometimes \(\mathbbm {G}_0\) is used to refer to \(\mathbbm {Z}_q\). For group elements we use the following implicit notation: for all \(i=1,\ldots ,k\), \([a]_i:=a\mathcal {P}_i\). The notation extends in a natural way to vectors and matrices and to linear algebra operations. We sometimes drop the index when referring to elements in \(\mathbbm {G}_1\), i.e., \([a]:=[a]_1=a\mathcal {P}_1\). In particular, it holds that \(e([a]_i,[b]_j)=[ab]_{i+j}\).

Additionally, for the asymmetric case, let \(\mathsf {AGen}_2\) be a PPT algorithm that on input \(1^\lambda \) returns a description of an asymmetric bilinear group \(\mathcal {AG}_2=(e,\mathbbm {G},\mathbbm {H},\mathbbm {T}, q, \mathcal {P},\mathcal {Q})\), where \(\mathbbm {G},\mathbbm {H},\mathbbm {T}\) are cyclic groups of prime-order q, \(\mathcal {P}\) is a generator of \(\mathbbm {G}\), \(\mathcal {Q}\) is a generator of \(\mathbbm {H}\) and \(e:\mathbbm {G}\times \mathbbm {H}\rightarrow \mathbbm {T}\) is a non-degenerate, efficiently computable bilinear map. In this case we refer to group elements as: \([a]_G:=a\mathcal {P}\), \([a]_H:=a\mathcal {Q}\) and \([a]_{T}:=a e(\mathcal {P},\mathcal {Q})\).

2.2 A Generic Model for Groups with Graded Encodings

In this section we describe a (purely algebraic) generic model for the graded encodings functionality, in order to obtain meaningful results about the hardness and separations of computational problems. The model is an adaptation of Maurer’s generic group model [33, 34] including the k-graded encodings, but in a completely algebraic formulation that follows the ideas in [5, 12, 20]. Since the k-graded encodings functionality implies the k-linear group functionality, the former gives more power to the adversaries or reductions working within the corresponding generic model. This in particular means that non-existential results proven in the richer k-graded encodings generic model also imply the same results in the k-linear group generic model. Therefore, in this paper we consider the former model. Due to the space limitations, we can only give a very succinct description of the model. See the full version of the paper [35] for a detailed and more formal description.

In a first approach we consider Maurer’s model adapted to the graded encodings functionality, but still not phrased in a purely algebraic language. In this model, an algorithm \(\mathcal {A}\) does not deal with proper group elements in \([y]_a\in \mathbbm {G}_a\), but only with labels (Ya), and it has access to an additional oracle internally performing the group operations, so that \(\mathcal {A}\) cannot benefit from the particular way the group elements are represented. Namely, on start all the group elements \([x_1]_{a_1},...,[x_\alpha ]_{a_\alpha }\) in the input intended for \(\mathcal {A}\) are replaced by the labels \((X_1,a_1),\ldots ,(X_\alpha ,a_\alpha )\). Then, \(\mathcal {A}\) actually receives as input the set of labels, and possibly some other non-group elements (i.e., that do not belong to any of the groups \(\mathbbm {G}_1,\ldots ,\mathbbm {G}_k\)), denoted as \(\widetilde{x}\), and considered as a bit string. For each group \(\mathbbm {G}_a\) two additional labels (0, a), (1, a), corresponding to the neutral element and the generator, are implicitly given to \(\mathcal {A}\). Then \(\mathcal {A}\) can adaptively make the following queries to an oracle implementing the k-graded encodings:

  • \(\mathsf {GroupOp}((Y_1,a),(Y_2,a))\): group operation in \(\mathbbm {G}_a\) for two previously issued labels in \(\mathbbm {G}_a\) resulting in a new label \((Y_3,a)\) in \(\mathbbm {G}_a\).

  • \(\mathsf {GroupInv}((Y,a))\): similarly for group inversion in \(\mathbbm {G}_a\).

  • \(\mathsf {GroupPair}((Y_1,a),(Y_2,b))\): bilinear map for two previously issued labels in \(\mathbbm {G}_a\) and \(\mathbbm {G}_b\), \(a+b\le k\), resulting in a new label \((Y_3,a+b)\) in \(\mathbbm {G}_{a+b}\).

  • \(\mathsf {GroupEqTest}((Y_1,a),(Y_2,a))\): test two previously issued labels in \(\mathbbm {G}_a\) for equality of the corresponding group elements, resulting in a bit (\(1=\) equality).

In addition, the oracle performs the actual computations with the group elements, and it uses them to answer the \(\mathsf {GroupEqTest}\) queries. Every badly formed query (for instance, containing a label not previously issued by the oracle or as an input to \(\mathcal {A}\)) is answered with a special rejection symbol \(\perp \). Following the usual step in generic group model proofs (see for instance [5, 11, 20]), we use polynomials as labels to group elements. Namely, labels in \(\mathbbm {G}_a\) are polynomials in \(\mathbbm {Z}_q[{\varvec{X}}]\), where the algebraic variables \({\varvec{X}}=(X_1,\ldots ,X_\alpha )\) are just formal representations of the group elements in the input of \(\mathcal {A}\). Now the oracle computes the new labels using the natural polynomial operations: \(\mathsf {GroupOp}((Y_1,a),(Y_2,a))=(Y_1+Y_2,a)\), \(\mathsf {GroupInv}((Y,a))=(-Y,a)\) and \(\mathsf {GroupPair}((Y_1,a),(Y_2,b))=(Y_1Y_2,a+b)\). It is easy to see that for any valid label (Ya), \(\deg Y\le a\).Footnote 4

The output of \(\mathcal {A}\) consists only of some labels \((Y_1,b_1),\ldots ,(Y_\beta ,b_\beta )\) (given at some time by the oracle) corresponding to group elements \([y_1]_{b_1},...,[y_\beta ]_{b_\beta }\), along with some non-group elements, denoted as \(\widetilde{y}\). Therefore, for any fixed random tape of \(\mathcal {A}\) and any choice of the non-group elements \(\widetilde{x}\), there exist polynomials \(Y_1,\ldots ,Y_\beta \in \mathbbm {Z}_q[{\varvec{X}}]\) of degrees upper bounded by \(b_1,\ldots ,b_\beta \) respectively, with coefficients known to \(\mathcal {A}\). Notice that \(\mathcal {A}\) itself can predict all answers given by the oracle except for some \(\mathsf {GroupEqTest}\) queries. In particular, some \(\mathsf {GroupEqTest}\) queries trivially result in 1, due to the group structure (e.g., \(\mathsf {GroupOp}((Y,a),\mathsf {GroupInv}((Y,a)))\) is the same as (0, a)), or due to the (known) a priori constraints in the input group elements (i.e., the definition of the problem instance given to \(\mathcal {A}\)). The answers to nontrivial \(\mathsf {GroupEqTest}\) queries (i.e., queries that cannot be trivially predicted by \(\mathcal {A}\)) are the only effective information \(\mathcal {A}\) can receive from the generic group oracle.

We now introduce a “purely algebraic” version of the generic model. For that, we need to assume that the distribution of \({\varvec{x}}\) can be sampled by evaluating a polynomial map f of constant degree at a random point.Footnote 5 This is not an actual restriction in our context since all Matrix Diffie-Hellman problems fulfil this requirement. In the “purely algebraic” model we redefine the oracle \(\mathsf {GroupEqTest}\) to answer 1 if and only if \(\mathcal {A}\) can itself predict the positive answer. Namely \(\mathsf {GroupEqTest}((Y_1,a),(Y_2,a))=1\) if and only if \(Y_1\circ f=Y_2\circ f\) as polynomials over \(\mathbbm {Z}_q\). With this change the behavior of \(\mathcal {A}\) can only differ negligibly from the original,Footnote 6 meaning that generic algorithms perform almost equally in Maurer’s model and its purely algebraic version. But now, any generic algorithm is just modelled by a set of polynomials. As we need to handle elements in different groups, we will use the shorter vector notation \([{\varvec{x}}]_{{\varvec{a}}} = ([x_1]_{a_1},\ldots ,[x_\alpha ]_{a_\alpha }) = (x_1\mathcal {P}_{a_1},\ldots ,x_\alpha \mathcal {P}_{a_\alpha })\in \mathbbm {G}_{a_1}\!\times \cdots \times \mathbbm {G}_{a_\alpha }\). Note that the length of a vector of indices \({\varvec{a}}\) is denoted by a corresponding Greek letter \(\alpha \). We will also use a tilde to denote variables containing only non-group elements (i.e., elements not in any of \(\mathbbm {G}_1,\ldots ,\mathbbm {G}_k\)).

Lemma 1

Let \(\mathcal {A}\) be an algorithm in the (purely algebraic) generic multilinear group model. Let \(([{\varvec{x}}]_{{\varvec{a}}},\widetilde{x})\) and \(([{\varvec{y}}]_{{\varvec{b}}},\widetilde{y})\) respectively be the input and output of \(\mathcal {A}\). Then, for every choice of \(\widetilde{x}\) and any choice of the random tape of \(\mathcal {A}\), there exist polynomials \(Y_1,\ldots ,Y_\beta \in \mathbbm {Z}_q[{\varvec{X}}]\) of degree upper bounded by \(b_1,\ldots ,b_\beta \) such that \({\varvec{y}}={\varvec{Y}}({\varvec{x}})\), for all possible \({\varvec{x}}\in \mathbbm {Z}_q^n\), where \({\varvec{Y}}=(Y_1,\ldots ,Y_\beta )\). Moreover, \(\widetilde{y}\) does not depend on \({\varvec{x}}\).

The proof of the lemma comes from the above discussion.

As usually, the proposed generic model reduces the analysis of the hardness of some problems to solving a merely algebraic problem related to polynomials. As an example, consider a computational problem \(\mathcal {P}\) which instances are entirely described by some group elements in the base group \(\mathbbm {G}_1\), \([{\varvec{x}}]\leftarrow \mathcal {P}\textsf {.InstGen}(1^\lambda )\), and its solutions are also described by some group elements \([{\varvec{y}}]_{{\varvec{b}}}\in \mathcal {P}\textsf {.Sol}([{\varvec{x}}])\). We also assume that \(\mathcal {P}\textsf {.InstGen}\) just samples \({\varvec{x}}\) by evaluating polynomial functions of constant degree at a random point. Then, \(\mathcal {P}\) is hard in the purely algebraic generic multilinear group model if and only if for all (randomized) polynomials \(Y_1,\ldots ,Y_\beta \in \mathbbm {Z}_q[{\varvec{X}}]\) of degrees upper bounded by \(b_1,\ldots ,b_\beta \) respectively,

$$ \Pr ([{\varvec{y}}]_{{\varvec{b}}}\in \mathcal {P}\textsf {.Sol}([{\varvec{x}}]):\; [{\varvec{x}}]\leftarrow \mathcal {P}\textsf {.InstGen}(1^\lambda ),\, {\varvec{y}}={\varvec{Y}}({\varvec{x}}))\in negl (\lambda ) $$

where \({\varvec{Y}}=(Y_1,\ldots ,Y_m)\) and the probability is computed with respect the random coins of the instance generator and the randomized polynomials.Footnote 7 In a few words, this means that the set \(\mathcal {P}\textsf {.Sol}([{\varvec{x}}])\) cannot be hit by polynomials of the given degree evaluated at \({\varvec{x}}\).

This model extends naturally to algorithms with oracle access (e.g., black-box reductions) but only when the oracles fit well into the generic model. Let us consider the algorithm \(\mathcal {A}^{\mathcal {O}}\), with oracle access to \(\mathcal {O}\). A completely arbitrary oracle (specified in the plain model) could have access to the internal representation of the group elements, and then it could leak some information about the group elements that is outside the generic group model. Thus, we will impose the very limiting constraint that the oracles are also “algebraic”, meaning that the oracle’s input/output behavior respects the one-wayness of the graded encodings, and it only performs polynomial operations on the input labels.

Definition 1

Let \(([{\varvec{u}}]_{{\varvec{d}}},\widetilde{u})\) and \(([{\varvec{v}}]_{{\varvec{e}}},\widetilde{v})\) respectively be a query to an oracle \(\mathcal {O}\) and its corresponding answer, where \(\widetilde{u}\) and \(\widetilde{v}\) contain the respective non-group elements. The oracle \(\mathcal {O}\) is called algebraic if for any choice of \(\widetilde{u}\) there exist polynomials \(V_1,\ldots ,V_\epsilon \in \mathbbm {Z}_q[{\varvec{U}},{\varvec{R}}]\), \({\varvec{R}}=(R_1,\ldots ,R_\rho )\), of constant degree (in the security parameter) such that

  • for the specific choice of \(\widetilde{u}\), \(v_i=V_i({\varvec{u}},{\varvec{r}})\), \(i=1,\ldots ,\epsilon \), for all \({\varvec{u}}\in \mathbbm {Z}_q^\epsilon \) and \({\varvec{r}}\in \mathbbm {Z}_q^\rho \), where \({\varvec{r}}=(r_1,\ldots ,r_\rho )\) are random parameters defined and uniformly sampled by the oracle,

  • \(\widetilde{v}\) does not depend on \({\varvec{u}},{\varvec{r}}\) (thus, \({\varvec{r}}\) can only have influence in the group elements in the answer),

  • \(V_j\) does not depend on any \(U_i\) such that \(e_j<d_i\) (in order to preserve the one-wayness of the graded encodings).

The parameters \({\varvec{r}}\) capture the behavior of an oracle solving a problem with many solutions (called here a “flexible” problem). They could be independent or not across different oracle calls, depending on whether the oracle is stateless or stateful. For technical reasons we consider only the stateless case with uniform sampling. Observe that the first two requirements in the definition mean that \({\varvec{v}}\) depends algebraically on \({\varvec{u}},{\varvec{r}}\) and no extra information about \({\varvec{u}},{\varvec{r}}\) can be leaked through \(\widetilde{v}\). Removing any of these requirements from the definition results in that a generic algorithm using such an oracle will no longer be algebraically generic. Also notice that after a call to an algebraic oracle, there is no guarantee that labels (Ya) fulfil the bound \(\deg Y \le a\).

Although the notion of algebraic oracle looks very limiting (e.g., it excludes a Discrete Logarithm oracle, as it destroys the one-wayness property of the graded encodings, but oracles solving \(\mathsf {CDH}\) or the Bilinear Computational Diffie-Hellman problem fit well in the definition), it is general enough for our purposes. We will need the following generalization of Lemma 1:

Lemma 2

Let \(\mathcal {A}^\mathcal {O}\) be an oracle algorithm in the (purely algebraic) generic multilinear group model, making a constant number of calls Q to an algebraic oracle \(\mathcal {O}\). Let \(([{\varvec{x}}]_{{\varvec{a}}},\widetilde{x})\) and \(([{\varvec{y}}]_{{\varvec{b}}},\widetilde{y})\) respectively be the input and output of \(\mathcal {A}\). Then, for every choice of \(\widetilde{x}\) and the random tape, there exist polynomials of constant degree \(Y_1,\ldots ,Y_\beta \in \mathbbm {Z}_q[{\varvec{X}},{\varvec{R}}_1,\ldots ,{\varvec{R}}_Q]\), such that \({\varvec{y}}={\varvec{Y}}({\varvec{x}},{\varvec{r}}_1,\ldots ,{\varvec{r}}_Q)\), for all possible inputs, where \({\varvec{Y}}=(Y_1,\ldots ,Y_\beta )\), and \({\varvec{r}}_1,\ldots ,{\varvec{r}}_Q\) are the parameters introduced in Definition 1 for the Q queries. Moreover, \(\widetilde{y}\) does not depend on \({\varvec{x}}\) or \({\varvec{r}}_1,\ldots ,{\varvec{r}}_Q\).

The proof of this lemma is given in Appendix A.

2.3 The Matrix Decisional Diffie-Hellman Assumption

We recall here the definition of the decisional assumptions introduced in [11], which are the starting point of our flexible computational matrix problems.

Definition 2

[11], Let \(\ell ,k \in \mathbbm {N}\) with \(\ell > k\). We call \(\mathcal {D}_{\ell ,k}\) a matrix distribution if it outputs (in polynomial time, with overwhelming probability) matrices in \(\mathbbm {Z}_q^{\ell \times k}\) of full rank k. We denote \(\mathcal {D}_k := \mathcal {D}_{k+1,k}\).

Definition 3

( \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) Assumption). [11] Let \(\mathcal {D}_{\ell ,k}\) be a matrix distribution. The \(\mathcal {D}_{\ell ,k}\)-Matrix Diffie-Hellman (\(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\)) Problem is telling apart the two probability distributions \((\mathbbm {G},q,\mathcal {P},[\mathbf {{A}}],[\mathbf {{A}} {\varvec{w}}])\) and \((\mathbbm {G},q,\mathcal {P},[\mathbf {{A}}],[{\varvec{z}}])\), where \(\mathbf {{A}} \leftarrow \mathcal {D}_{\ell ,k}, {\varvec{w}} \leftarrow \mathbbm {Z}_q^k, {\varvec{z}} \leftarrow \mathbbm {Z}_q^{\ell }\).

We say that the \(\mathcal {D}_{\ell ,k}\)-Matrix Diffie-Hellman (\(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\)) Assumption holds relative to \(\mathsf {Gen}\) if the corresponding problem is hard, that is, if for all PPT adversaries \(\mathcal {A}\), the advantage

$$\mathbf {Adv}_{\mathcal {D}_{\ell ,k},\mathsf {Gen}}(\mathcal {A}) = \Pr [\mathcal {A}(\mathcal {G},[\mathbf {{A}}], [\mathbf {{A}} {\varvec{w}}])=1]-\Pr [\mathcal {A}(\mathcal {G},[\mathbf {{A}}], [{\varvec{z}}])=1] \in negl (\lambda ),$$

where the probability is taken over \(\mathcal {G}=(\mathbbm {G},q,\mathcal {P}) \leftarrow \mathsf {Gen}(1^\lambda )\), \(\mathbf {{A}} \leftarrow \mathcal {D}_{\ell ,k}, {\varvec{w}} \leftarrow \mathbbm {Z}_q^k, {\varvec{z}} \leftarrow \mathbbm {Z}_q^{\ell }\) and the coin tosses of adversary \(\mathcal {A}\).

In the case of asymmetric bilinear groups or symmetric k-linear groups, we similarly say that the \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) Assumption holds relative to \(\mathsf {AGen}_2\) or \(\mathsf {MGen}_{k}\), respectively. In the former we specify if the assumption holds in the left (\(\mathcal {A}\) receives \([\mathbf {{A}}]_G\), \([\mathbf {{A}} {\varvec{w}}]_G\) or \([{\varvec{z}}]_G\)), or in the right (\(\mathcal {A}\) receives \([\mathbf {{A}}]_H\), \([\mathbf {{A}} {\varvec{w}}]_H\) or \([{\varvec{z}}]_H\)).

Definition 4

A matrix distribution \(\mathcal {D}_{\ell ,k}\) is hard if the corresponding \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) problem is hard in the generic k-linear group model.

Many different matrix distributions appear in the literature. Namely, the cascade \(\mathcal {C}_{k}\) and symmetric cascade \(\mathcal {SC}_{k}\) distributions were presented in [11], while the uniform \(\mathcal {U}_{\ell ,k}\), the linear \(\mathcal {L}_{k}\), the randomized linear \(\mathcal {RL}_{k}\) and the square polynomial \(\mathcal {P}_{\ell ,2}\) distributions were implicitly used in some previous works. We give their explicit definitions in Appendix B.

3 The Matrix Diffie-Hellman Computational Problems

In this section we introduce two families of search problems naturally related to the Matrix Decisional Diffie-Hellman problems. In the first family, given a matrix \([\mathbf {{A}}]\), where \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\), and the first k components of a vector \([{\varvec{z}}]\), the problem is completing it so that \({\varvec{z}}\in \mathop {\text {Im}}\mathbf {{A}}\).

Definition 5

( \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MCDH}\) ). Given a matrix distribution \(\mathcal {D}_{\ell ,k}\), such that the upper \(k\times k\) submatrix of \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\) has full rank with overwhelming probability, the computational matrix Diffie-Hellman Problem is given \(([\mathbf {{A}}],[{\varvec{z}}_0])\), with \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\), \({\varvec{z}}_0\leftarrow \mathbbm {Z}_q^k\), compute \([{\varvec{z}}_1]\in \mathbbm {G}^{\ell -k}\) such that \(({\varvec{z}}_0\Vert {\varvec{z}}_1)\in \mathop {\text {Im}}\mathbf {{A}}\).

The full-rank condition ensures the existence of solutions to the \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MCDH}\) problem instance. Thus, we tolerate the existence of a negligible fraction of unsolvable problem instances. Indeed, all known interesting matrix distributions fulfil this requirement. Notice that \(\mathsf {CDH}\) and the computational \(k\text{- }\mathsf {Lin}\) problems are particular examples of \(\mathsf {MCDH}\) problems. Namely, \(\mathsf {CDH}\) is exactly \(\mathcal {L}_{1}\)-\(\mathsf {MCDH}\) and the computational \(k\text{- }\mathsf {Lin}\) problem is \(\mathcal {L}_{k}\)-\(\mathsf {MCDH}\). Indeed, the \(\mathcal {L}_{1}\)-\(\mathsf {MCDH}\) problem is given \([1],[a],[z_1]\), compute \([z_2]\) such that \((z_1,z_2)\) is collinear with (1, a), or equivalently, \(z_2=z_1 a\), which is solving the \(\mathsf {CDH}\) problem. All \(\mathsf {MCDH}\) problems have a unique solution and they appear naturally in some scenarios using \(\mathsf {MDDH}\) problems. For instance, the one-wayness of the encryption scheme in [11] is equivalent to the corresponding \(\mathsf {MCDH}\) assumption.

There is an immediate relation between any \(\mathsf {MCDH}\) problem and its decisional counterpart. Not surprisingly, for any matrix distribution \(\mathcal {D}_{\ell ,k}\), \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) \(\Rightarrow \) \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MCDH}\).

We are not going to study the possible reductions between \(\mathsf {MCDH}\) problems, due to the fact that, essentially, any \(\mathsf {MCDH}\) problem amounts to computing some polynomial on the elements of \(\mathbf {{A}}\), and it is then equivalent to \(\mathsf {CDH}\) ([4, 23]), although the tightness of the reduction depends on the degree of the polynomial.

In the second family of computational problems, given a matrix \([\mathbf {{A}}]\), where \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\), the problem is finding \([{\varvec{x}}]\) such that \({\varvec{x}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\). It is notable that some computational problems in the literature are particular cases of this second family.

Definition 6

( \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\) ). Given a matrix distribution \(\mathcal {D}_{\ell ,k}\), the Kernel Diffie-Hellman Problem is given \([\mathbf {{A}}]\), with \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\), find a nonzero vector \([{\varvec{x}}]\in \mathbbm {G}^{\ell }\) such that \({\varvec{x}}\) is orthogonal to \(\mathop {\text {Im}}\mathbf {{A}}\), that is, \({\varvec{x}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\).

Definition 6 naturally extends to asymmetric bilinear groups. There, given \([\mathbf {{A}}]_H\), the problem is to find \([{\varvec{x}}]_G\) such that \({\varvec{x}} \in \ker \mathbf {{A}}^{\top } {\setminus } \{{\varvec{0}}\}\). A solution can be obviously verified by checking if \(e([{\varvec{x}}^{\top }]_G,[\mathbf {{A}}]_H)=[{\varvec{0}}]_T\). We can also consider an extension of this problem in which the goal is to solve the same problem but giving the solution in a different group \(\mathbbm {G}_r\), in some ideal graded encoding \(\mathcal {MG}_{m}\), for some \(0 \le r \le \min (m,k-1)\). The case \(r=1\) corresponds to the previous problem defined in a m-linear group.

Definition 7

( \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) ). Given a matrix distribution \(\mathcal {D}_{\ell ,k}\) over a m-linear group \(\mathcal {MG}_{m}\) and r an integer \(0 \le r \le \min (m,k-1)\), the \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) Problem is to find \([{\varvec{x}}]_r \in \mathbbm {G}_r^{\ell }\) such that \({\varvec{x}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\).

When the precise degree of multilinearity m is not an issue, we will write \((r,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) instead of \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\), for any \(m\ge r\). We excluded the case \(r\ge k\) because the problem is easy.

Lemma 3

For all integers \(k\le r\le m\) and for all matrix distributions \(\mathcal {D}_{\ell ,k}\), the \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) Problem is easy.

The kernel problem is also harder than the corresponding decisional problem, in multilinear groups.

Lemma 4

In a m-linear group, \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) \(\Rightarrow \) \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) for any matrix distribution \(\mathcal {D}_{\ell ,k}\) and for any \(0 \le r \le m-1\). In particular, for \(m\ge 2\), \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {MDDH}\) \(\Rightarrow \) \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\).

The proofs of Lemmas 3, and 4 can be found in the full version of this paper [35].

3.1 The Kernel DH Assumptions in the Multilinear Maps Candidates

We have shown that for any hard matrix distribution \(\mathcal {D}_{\ell ,k}\) the \(\mathcal {D}_{\ell ,k}\)-\(\mathsf {KerMDH}\) problem is generically hard in m-linear groups. We emphasize that all our results refer to generic, ideal multilinear maps (in fact, to graded encodings, which have more functionality). Our aim is only to give necessary condition for the assumptions to hold in candidate multilinear maps. The status of current candidate multilinear maps is rather uncertain, e.g. it is described in [28] as “break-and-repair mode”. Thus, it is hard to argue if our assumptions hold in any concrete instantiation and we leave this as an open question for further investigation.

3.2 A Unifying View on Computational Matrix Problems

In this section we show how some computational problems in the cryptographic literature are unified as particular instances of \(\mathsf {KerMDH}\) problems. Their explicit definitions are given in Appendix C. It is straightforward to see that Find-Rep [9] Assumption is just \((0,\mathcal {U}_{\ell ,1})\)-\(\mathsf {KerMDH}\), the Simultaneous Double Pairing Assumption (SDP) [2] is \(\mathcal {RL}_{2}\)-\(\mathsf {KerMDH}\), the Simultaneous Triple Pairing [18] Assumption is \(\mathcal {U}_{2}\)-\(\mathsf {KerMDH}\), the Simultaneous Pairing [19] Assumption is \(\mathcal {P}_{\ell ,2}\)-\(\mathsf {KerMDH}\). The Double Pairing (DP) [18] Assumption corresponds to \(\mathcal {U}_{1}\)-\(\mathsf {KerMDH}\) in an asymmetric bilinear setting. On the other hand, the 1-Flexible Diffie-Hellman (1-FlexDH) [32] Assumption is \(\mathcal {C}_{2}\)-\(\mathsf {KerMDH}\), the 1-Flexible Square Diffie-Hellman (1-FlexSDH) [27] Assumption is \(\mathcal {SC}_{2}\)-\(\mathsf {KerMDH}\), and the \(\ell \)-Flexible Diffie-Hellman (\(\ell \)-FlexDH) [32] Assumption for \(\ell >1\) is the only one which is not in the \(\mathsf {KerMDH}\) family. However, \(\ell \)-FlexDH \(\Rightarrow \mathcal {C}_{\ell +1}\)-\(\mathsf {KerMDH}\). Getting the last three results requires a bit more work, and they are proven in the full version [35].

4 Reduction and Separation of Kernel Diffie-Hellman Problems

In this section we prove that the most important matrix distribution families \(\mathcal {U}_{\ell ,k}\), \(\mathcal {L}_k\), \(\mathcal {SC}_k\), \(\mathcal {C}_k\) and \(\mathcal {RL}_k\) (see Appendix B) define families of \(\mathsf {KerMDH}\) problems with strictly increasing hardness, as we precisely state in Theorem 2, at the end of the section. By ‘strictly increasing’ we mean that (1) there are known reductions of the smaller problems to the larger problems (in terms of k) within each family, and (2) there are no black-box reductions in the other way in the multilinear generic group model. This result shows the necessity of using \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) Assumptions for \(k>2\). A similar result is known for the corresponding \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) problems. Indeed, one can easily prove a separation between large and small decisional problems. Observe that any efficient m-linear map can efficiently solve any \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) problem with \(k\le m-1\), and therefore every two \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {MDDH}\) problems with \(\widetilde{k}<k\) are separated by an oracle computing a k-linear map. However, when dealing with the computational \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) family, no such a trivial argument is known to exist. Actually, a m-linear map does not seem to help to solve any \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) problem with \(k>1\). Furthermore, the m-linear map seems to be useless for any (reasonable) reduction between \(\mathsf {KerMDH}\) problems defined over the same group. Indeed, all group elements involved in the problem instances and their solutions belong to the base group \(\mathbbm {G}\), and the result of computing any m-linear map is an element in \(\mathbbm {G}_m\), where no efficient map from \(\mathbbm {G}_m\) back to \(\mathbbm {G}\) is supposed to exist.

4.1 Separation

In this section we firstly show the non-existential part of Theorem 2. Namely, we show that there is no black-box reduction in the generic group model (described in Sect. 2.2) from \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\) for \(k > \widetilde{k}\), assuming that the two matrix distributions \(\mathcal {D}_{\ell ,k}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\) are hard (see Definition 4). Before proving the main result we need some technical lemmas and also a new geometrical notion defined on a family of subspaces of a vector space, named t-Elusiveness.

In the first lemma we show that the natural (black-box, algebraic) reductions between \(\mathsf {KerMDH}\) problems have a very special form. Observe that a black-box reduction to a flexible problem must work for any adversary solving it. In particular, the reduction should work for any solution given by this adversary, or for any probability distribution of the solutions given by it. Informally, the lemma states that the output of a successful reduction can always be computed in essentially two ways: (1) By just applying a (randomized) linear map to the answer given by the adversary in the last call. Therefore, all possibly existing previous calls to the adversary are just used to prepare the last one. (2) By just ignoring the last call to the adversary and using only the information gathered in the previous ones.

Let \(\mathcal {R}^{\mathcal {O}}\) be a black-box reduction of \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\), in the purely algebraic generic multilinear group model, discussed in Sect. 2.2, for some matrix distributions \(\mathcal {D}_{\ell ,k}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\). Namely, \(\mathcal {R}^{\mathcal {O}}\) solves \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) with a non-negligible probability by making \(Q\ge 1\) queries to an oracle \(\mathcal {O}\) solving \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\) with probability one. As we aim at ruling out the existence of some reductions, we just consider the best possible case any black-box reduction must be able to handle. Now we split the reduction as \(\mathcal {R}^{\mathcal {O}}=(\mathcal {R}^{\mathcal {O}}_0,\mathcal {R}_1)\), where the splitting point is the last oracle call, as shown in Fig. 2. We actually use the same splitting in the proof of Lemma 2 in Appendix D. More formally, on the input of \([\mathbf {{A}}]\), for \(\mathbf {{A}}\leftarrow \mathcal {D}_{\ell ,k}\), and after making \(Q-1\) oracle calls, \(\mathcal {R}^{\mathcal {O}}_0\) stops by outputting the last query to \(\mathcal {O}\), that is a matrix \([\mathbf {{\widetilde{A}}}]\), where \(\mathbf {{\widetilde{A}}}\in \mathcal {D}_{\widetilde{\ell },\widetilde{k}}\), together with some state information s for \(\mathcal {R}_1\). Next, \(\mathcal {R}_1\) resumes the execution from s and the answer \([{\varvec{w}}]\in \mathbbm {G}^{\widetilde{\ell }}\) given by the oracle, and finally outputs \([{\varvec{v}}]\in \mathbbm {G}^\ell \). Without loss of generality, we assume that both stages \(\mathcal {R}^{\mathcal {O}}_0\) and \(\mathcal {R}_1\) receive the same random tape, $ (\(\mathcal {R}_1\) can redo the computations performed by \(\mathcal {R}^{\mathcal {O}}_0\)).

Fig. 2.
figure 2

Splitting of the black-box reduction.

Lemma 5

There exists an algebraic oracle \(\mathcal {O}\) (in the sense of Definition 1), that solves the \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) Problem with probability one.

All the proofs in Sect. 4 are given in Appendix D.

Lemma 2 applied to \(\mathcal {R}^{\mathcal {O}}_0\) (and using also Lemma 5) implies that only the group elements in s can depend on \(\mathbf {{A}}\). Indeed, the non-group elements in s can only depend on the random tape $. Now, from Lemma 1 applied to \(\mathcal {R}_1\), we know that its output \([{\varvec{v}}]\) is determined by a polynomial map of total degree at most one in the input group elements (i.e., \(\mathbf {{\widetilde{A}}}\) and the group elements in s), and the coefficients of this polynomial can only depend on $, and the non-group elements in s, which in turn only depend on $. Therefore, splitting the polynomial map into two parts, for every fixed $ and every fixed oracle behavior in the first \(Q-1\) oracle calls there exists a vector \(u\in \mathbbm {Z}_q^\ell \) and a linear map \(\eta :\mathbbm {Z}_q^{\widetilde{\ell }}\rightarrow \mathbbm {Z}_q^\ell \) such that we can write \({\varvec{v}}={\varvec{u}}+\eta ({\varvec{w}})\), where \({\varvec{u}}\) actually depends on the group elements in s. The important fact here is that \(\eta \) can only depend on $, but not on \(\mathbf {{A}}\).

Lemma 6

Let \(\mathcal {R}^{\mathcal {O}}=(\mathcal {R}^{\mathcal {O}}_0,\mathcal {R}_1)\) be a black-box reduction from \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\), in the purely algebraic generic multilinear group model, making \(Q\ge 1\) calls to an oracle \(\mathcal {O}\) solving the latter with probability one. If \(\mathcal {R}^{\mathcal {O}}\) succeeds with a non negligible probability \(\varepsilon \) then, for every possible behavior of the oracle, either \(\Pr (\eta ({\varvec{w}})\in S')> negl \) or \(\Pr ({\varvec{u}}\in S')> negl \), where \(S'=\ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\), \([\mathbf {{A}}]\) is the input of \(\mathcal {R}^{\mathcal {O}}\), and its output is written as \([{\varvec{u}}+\eta ({\varvec{w}})]\), for some \({\varvec{u}}\) only depending on the state output by \(\mathcal {R}^{\mathcal {O}}_0\), \([{\varvec{w}}]\) is the answer to the Q-th oracle query, and \(\eta :\mathbbm {Z}_q^{\widetilde{l}}\rightarrow \mathbbm {Z}_q^l\) is a (randomized) linear map that only depends on the random tape of \(\mathcal {R}^{\mathcal {O}}\).

The following property of the hard matrix distributions allows us to prove that indeed in the last lemma \(\Pr (\eta ({\varvec{w}})\in S{\setminus }\{{\varvec{0}}\})\in negl \).

Definition 8

( t -Elusiveness). A family of subspaces \(\mathcal {S}\) of a vector space X over the finite field \(\mathbbm {Z}_q\) is called t-elusive for some \(t<\dim X\) if for all t-dimensional subspaces \(F\subset X\), \(\Pr (F\cap S \ne \{{\varvec{0}}\}) \in negl \), where the probability is computed with respect to the choice of \(S\in \mathcal {S}\). A matrix distribution \(\mathcal {D}_{\ell ,k}\) is called t-elusive if the family \(\{\ker \mathbf {{A}}^\top \}_{\mathbf {{A}}\in \mathcal {D}_{\ell ,k}}\) is t-elusive.

Lemma 7

If a matrix distribution \(\mathcal {D}_{\ell ,k}\) is hard (as given in Definition 4) then \(\mathcal {D}_{\ell ,k}\) is k-elusive.

In the next theorem we use the k-elusiveness to prove that \(\Pr ({\varvec{u}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\})> negl \) for all possible behaviors of the oracle in the first \(Q-1\) calls. This actually implies that the reduction can directly output \({\varvec{u}}\), and only \(Q-1\) oracle calls are actually needed. Therefore, by the descent method we show that no successful reduction exists unless \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) is easy.

Theorem 1

Let \(\mathcal {D}_{\ell ,k}\) be k-elusive. If there exists a black-box reduction in the purely algebraic generic multilinear group model from \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to another problem \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\) with \(\widetilde{k}< k\), then \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) is easy.

Now we consider the contrapositive statement, that directly applies to the known families of hard matrix distributions.

Corollary 1

If a matrix distribution family \(\{\mathcal {D}_{\ell ,k}\}\) is hard then for any \(\mathcal {D}_{\ell ,k}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\) in the family with \(k>\widetilde{k}\) there is no black-box reduction in the generic group model from \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\).

Proof

Since all \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\) problems in the family are generically hard on a k-linear group, we know that \(\mathcal {D}_{\ell ,k}\) is k-elusive by Lemma 7, and also \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) is hard in that group (otherwise, any solution to \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) can be used to solve \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {MDDH}\)). By the above theorem, no black-box reduction in the generic group model from \(\mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\) to \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\) can exist for \(k>\widetilde{k}\).

4.2 Increasing Families of \(\mathsf {KerMDH}\) Problems

Most matrix distributions, like \(\mathcal {U}_{\ell ,k}\), \(\mathcal {L}_k\), \(\mathcal {SC}_k\), \(\mathcal {C}_k\) and \(\mathcal {RL}_k\), are indeed families parameterized by their size k. The negative results in Corollary 1 prevent us from finding reductions from larger to smaller \(\mathsf {KerMDH}\) problems. Nevertheless, we provide here some examples of (tight) reductions going in the other way, within each of the previous families.

Lemma 8

\(\mathcal {U}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {U}_{\ell ,k}\text {-}\mathsf {KerMDH}\) for any \(\widetilde{k}\le k\), \(\widetilde{\ell }>\widetilde{k}\) and \(\ell >k\).

Proof

We divide the proof into two steps: Firstly, assume that \(\widetilde{\ell }=\widetilde{k}+1\), \(k\ge \widetilde{k}\), \(\ell \ge k+1\). Given an instance \([\mathbf {{\widetilde{A}}}]\), with \(\mathbf {{\widetilde{A}}}\leftarrow \mathcal {U}_{\widetilde{k}+1,\widetilde{k}}\), we choose a full-rank matrix \(\mathbf {{L}}\in \mathbbm {Z}_q^{\ell \times (k+1)}\) and compute \([\mathbf {{A}}]=\mathbf {{L}}([\mathbf {{\widetilde{A}}}]\oplus [\mathbf {{I}}])\), where \(\mathbf {{I}}\) is the identity matrix of size \((k-\widetilde{k})\times (k-\widetilde{k})\) and \(\oplus \) operation denotes diagonal block matrix concatenation. That is

$$ U\oplus V=\begin{pmatrix}U&{}0\\ 0&{} V\end{pmatrix}. $$

Clearly, the probability distribution of the new matrix is statistically close to the uniform distribution in \(\mathbbm {Z}_q^{\ell \times k}\). Any vector \([{\varvec{x}}]\), obtained from a solver of \(\mathcal {U}_{\ell ,k}\text {-}\mathsf {KerMDH}\), such that \({\varvec{x}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\) can be transformed into \([{\varvec{\widetilde{x}}}]\) such that \({\varvec{\widetilde{x}}}\in \ker \mathbf {{\widetilde{A}}}^\top {\setminus }\{{\varvec{0}}\}\) with overwhelming probability,Footnote 8 by just letting \([{\varvec{\widetilde{x}}}]\) to be the first \(\widetilde{k}+1\) components of \(\mathbf {{L}}^\top [{\varvec{x}}]\). Thus, we have built a tight reduction \(\mathcal {U}_{\widetilde{k}+1,\widetilde{k}}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {U}_{\ell ,k}\text {-}\mathsf {KerMDH}\).

The second step, \(k=\widetilde{k}\), \(\widetilde{\ell }>\ell =\widetilde{k}+1\), is simpler. Given an instance \([\mathbf {{\widetilde{A}}}]\), with \(\mathbf {{\widetilde{A}}}\leftarrow \mathcal {U}_{\widetilde{\ell },\widetilde{k}}\), define the matrix \([\mathbf {{A}}]\) to be the upper \(\widetilde{k}+1\) rows of \([\mathbf {{\widetilde{A}}}]\). Clearly \(\mathbf {{A}}\) follows the uniform distribution in \(\mathbbm {Z}_q^{(\widetilde{k}+1)\times \widetilde{k}}\). Now, any vector \([{\varvec{x}}]\) such that \({\varvec{x}}\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\) can be transformed into \([{\varvec{\widetilde{x}}}]\) such that \({\varvec{\widetilde{x}}}\in \ker \mathbf {{\widetilde{A}}}^\top {\setminus }\{{\varvec{0}}\}\), by just padding \({\varvec{x}}\) with \(\widetilde{\ell }-\widetilde{k}-1\) zeros. Thus, \(\mathcal {U}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {U}_{\widetilde{k}+1,\widetilde{k}}\text {-}\mathsf {KerMDH}\). By concatenating the two tight reductions we obtain the general case.

Lemma 9

For \(\mathcal {D}_k=\mathcal {L}_k\), \(\mathcal {SC}_k\), \(\mathcal {C}_k\) and \(\mathcal {RL}_k\),   \(\mathcal {D}_k\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {D}_{k+1}\text {-}\mathsf {KerMDH}\).

Proof

We start with the case \(\mathcal {D}_k=\mathcal {L}_{k}\). Observe that given a matrix \(\mathbf {{\widetilde{A}}}\leftarrow \mathcal {L}_k\), with parameters \(a_1,\ldots ,a_k\), we can build a matrix \(\mathbf {{A}}\) following the distribution \(\mathcal {L}_{k+1}\), by adding an extra row and column to \(\mathbf {{\widetilde{A}}}\) corresponding to new random parameter \(a_{k+1}\in \mathbbm {Z}_q\). Moreover, given \({\varvec{x}}=(x_1,\ldots ,x_{k+2})\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\), the vector \({\varvec{\widetilde{x}}}=(x_1,\ldots ,x_k,x_{k+2})\) is in \(\ker \mathbf {{\widetilde{A}}}^\top {\setminus }\{{\varvec{0}}\}\) (except for a negligible probability due to the possibility that \(a_{k+1}=0\) and \({\varvec{\widetilde{x}}}={\varvec{0}}\), while \({\varvec{x}}\ne {\varvec{0}}\)). The reduction consists of choosing a random \(a_{k+1}\), then building \([\mathbf {{A}}]\) from \([\mathbf {{\widetilde{A}}}]\) as above, and finally obtaining \([{\varvec{\widetilde{x}}}]\) from \([{\varvec{x}}]\) by deleting the \((k+1)\)-th coordinate.

Similarly, from a matrix \(\mathbf {{\widetilde{A}}}\leftarrow \mathcal {SC}_k\), with parameter a, we can obtain a matrix \(\mathbf {{A}}\) following \(\mathcal {SC}_{k+1}\) by adding a new row and column to \(\mathbf {{\widetilde{A}}}\). Now given \({\varvec{x}}=(x_1,\ldots ,x_{k+2})\in \ker \mathbf {{A}}^\top {\setminus }\{{\varvec{0}}\}\), it is easy to see that the vector \({\varvec{\widetilde{x}}}=(x_1,\ldots ,x_{k+1})\) is always in \(\ker \mathbf {{\widetilde{A}}}^\top {\setminus }\{{\varvec{0}}\}\).

\(\mathcal {C}_{k}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {C}_{k+1}\text {-}\mathsf {KerMDH}\) and \(\mathcal {RL}_{k}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {RL}_{k+1}\text {-}\mathsf {KerMDH}\) are proven using the same ideas.

By combining Corollary 1 with the explicit reductions given above, we can now state our main result in this section.

Theorem 2

The matrix distribution families \(\{\mathcal {U}_{\ell ,k}\}\), \(\{\mathcal {L}_k\}\), \(\{\mathcal {SC}_k\}\), \(\{\mathcal {C}_k\}\) and \(\{\mathcal {RL}_k\}\) define families of \(\mathsf {KerMDH}\) problems with strictly increasing hardness. Namely, for any \(\mathcal {D}_{\ell ,k}\) and \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\) belonging to one of the previous families, such that \(\widetilde{k}< k\),

  1. 1.

    there exists a tight reduction, \(\mathcal {D}_{\widetilde{\ell },\widetilde{k}}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {D}_{\ell ,k}\text {-}\mathsf {KerMDH}\),

  2. 2.

    there is no black-box reduction in the generic group model in the opposite direction.

5 Applications

We have already mentioned that the Kernel Matrix Diffie-Hellman Assumptions have already found applications in follow-up work, more concretely: (a) to generalize and improve previous constructions of QA-NIZK proofs for linear spaces [26], (b) to construct more efficient structure preserving signatures starting from affine algebraic MACS [25], (c) to improve and generalize aggregation of Groth-Sahai proofs [17] or (d) to construct a tightly secure QA-NIZK argument for linear subspaces with unbounded simulation soundness [15].

As a new application, we use our new framework to abstract two constructions of trapdoor commitments. See for instance [3] for the formal definition of a trapdoor commitment scheme \(C=(\mathsf {K},\mathsf {Comm},\mathsf {Vrfy},\mathsf {TrapdoorEquiv})\) and Sect. 6 for a discussion on the advantages of instantiating these commitments with the new circulant matrix distribution.

5.1 Generalized Pedersen Commments in Multilinear Groups

In a group \((\mathbbm {G},q,\mathcal {P})\) where the discrete logarithm is hard, the Pedersen commitment is a statistically hiding and computationally binding commitment to a scalar. It can be naturally generalized to several scalars. Abe et al. [2] show how to do similar Pedersen type commitments to vectors of group elements. With our new assumption family we can write both the Pedersen commitment and the commitment of [2] as a single construction and generalize it to (ideal) graded encodings.

  • \(\mathsf {K}(1^{\lambda },d,m)\): Let \(\mathcal {MG}_{m}=(e,\mathbbm {G}_1,\mathbbm {G}_2,\ldots , \mathbbm {G}_{m}, q, \mathcal {P}_1,\ldots ,\mathcal {P}_m) \leftarrow \mathsf {MGen}_{m}(1^{\lambda })\). Sample \(\mathbf {{A}}\leftarrow \mathcal {D}_{k+d,k}\). Let \(\overline{\mathbf {{A}}}\) be the first k rows of \(\mathbf {{A}}\) and \(\underline{\mathbf {{A}}}\) the remaining d rows and \(\mathbf {{T}}:=\underline{\mathbf {{A}}} \overline{\mathbf {{A}}}^{-1}\) (w.l.o.g. we can assume \(\overline{\mathbf {{A}}}\) is invertible). Output \(ck:=(\mathcal {MG}_{m},[\mathbf {{A}}]_1)\), \(tk:=(\mathbf {{T}})\).

  • \(\mathsf {Comm}(ck,[{\varvec{v}}]_r)\): To commit to a vector \([{\varvec{v}}]_r \in \mathbbm {G}^{d}_r\), for any \(r < m\), pick \({\varvec{s}} \leftarrow \mathbbm {Z}_q^k\), and output \([{\varvec{c}}]_{r+1}:=e( [\begin{pmatrix} {\varvec{s}}^{\top }&||&{\varvec{v}}^{\top } \end{pmatrix}]_r, [\mathbf {{A}}]_1)=[\begin{pmatrix} {\varvec{s}}^{\top }&||&{\varvec{v}}^{\top } \end{pmatrix} \mathbf {{A}}]_{r+1} \in \mathbbm {G}_{r+1}^{k},\) and the opening \(Op=([{\varvec{s}}]_r)\).

  • \(\mathsf {Vrfy}(ck,[{\varvec{v}}]_r,Op)\): Given a message \([{\varvec{v}}]_r\) and opening \(Op=([{\varvec{s}}]_r)\), this algorithm outputs 1 if \([{\varvec{c}}]_{r+1}=e([\begin{pmatrix} {\varvec{s}}^{\top }&||&{\varvec{v}}^{\top } \end{pmatrix}]_r, [\mathbf {{A}}]_1).\)

  • \(\mathsf {TrapdoorEquiv}(ck,tk,[{\varvec{c}}]_{r+1}, [{\varvec{v}}]_r, Op, [{\varvec{v}}']_r )\): On a commitment \([{\varvec{c}}]_{r+1} \in \mathbbm {G}^{k}_{r+1}\) to message \([{\varvec{v}}]_r\) with opening \(Op=([{\varvec{s}}]_r)\), compute: \([{\varvec{s}}']_r:= [{\varvec{s}}]_r + \mathbf {{T}}^{\top } [({\varvec{v}}-{\varvec{v}}')]_r \in \mathbbm {G}_r^{k}.\) Output \(Op'=([{\varvec{s}}']_r)\) as the opening of \([{\varvec{c}}]_{r+1}\) to \([{\varvec{v}}']_r\).

The analysis is almost identical to [2]. The correctness of the trapdoor opening is straightforward. The hiding property of the commitment is unconditional, while the soundness (at level r) is based on the \((r,m,\mathcal {D}_{\ell ,k})\)-\(\mathsf {KerMDH}\) Assumption. Indeed, given two messages \([{\varvec{v}}]_r,[{\varvec{v}}']_r\) with respective openings \([{\varvec{s}}]_r,[{\varvec{s}}']_r\), it obviously follows that \([{\varvec{w}}]:=[\begin{pmatrix} ({\varvec{s}}-{\varvec{s}}')^{\top }&||&({\varvec{v}}-{\varvec{v}}')^{\top } \end{pmatrix}]_r\) is a nonzero element in the kernel (in \(\mathbbm {G}_r\)) of \(\mathbf {{A}}^{\top }\), i.e. \(e([{\varvec{w}}^{\top }]_r,[\mathbf {{A}}]_1)=[{\varvec{0}}]_{r+1}\).

Notice that the Pedersen commitment (to multiple elements) is for messages in \(\mathbbm {G}_0\) and \(\mathbf {{A}}\leftarrow \mathcal {U}_{d+1,1}\) and soundness is based on the \((0,m,\mathcal {U}_{d+1,1})\)-\(\mathsf {KerMDH}\). The construction proposed in [2] is for an asymmetric bilinear group \(\mathcal {AG}_2\), and in this case messages are vectors in the group \(\mathbbm {H}\) and the commitment key consists of elements in \(\mathbbm {G}\), i.e. \(ck=(\mathcal {AG}_2,[\mathbf {{A}}]_{G})\), \(\mathbf {{A}}\leftarrow \mathcal {U}_{d+1,1}\). Further, a previous version of the commitment scheme of [2] in symmetric bilinear groups (in [18]) corresponds to our construction with \(\mathbf {{A}}\leftarrow \mathcal {U}_{2+d,2}\).

5.2 Group-to-Group Commitments

The commitments of the previous section are “shrinking” because they map a vector of length d in the group \(\mathbbm {G}_r\) to a vector of length k, for some k independent of and typically smaller than d. Abe et al. [3] noted that in some applications it is useful to have “group-to-group” commitments, i.e. commitments which are defined in the same group as the vector message. The motivation for doing so in the bilinear case is that these commitments are better compatible with Groth-Sahai proofs.

There is a natural construction of group-to-group commitments which uses the generalized Pedersen commitment of Sect. 5.1, which is denoted as \(\mathsf {Ped.C}=(\widetilde{\mathsf {K}},\widetilde{\mathsf {Comm}},\widetilde{\mathsf {Vrfy}},\widetilde{\mathsf {TrapdoorEquiv}})\) in the following.

  • \(\mathsf {K}(1^{\lambda },d,m)\): Run \((\widetilde{ck},\widetilde{tk}) \leftarrow \widetilde{\mathsf {K}}(1^{\lambda },m,d)\), output \(ck=\widetilde{ck}\) and \(tk=\widetilde{tk}\).

  • \(\mathsf {Comm}(ck,[{\varvec{v}}]_r)\): To commit to a vector \([{\varvec{v}}]_r \in \mathbbm {G}_r^{d}\), for any \(0< r < m\), pick \([{\varvec{t}}]_{r-1} \leftarrow [\mathbbm {G}]_{r-1}^{k}\). Let \(([\tilde{{\varvec{c}}}]_{r},\widetilde{Op}=([{\varvec{s}}]_{r-1})) \leftarrow \widetilde{\mathsf {Comm}}(ck,[{\varvec{t}}]_{r-1})\) and output \(c:=([{\varvec{t}}+{\varvec{v}}]_r, [\tilde{{\varvec{c}}}]_r)\) and the opening \(Op=([{\varvec{s}}]_{r})\).

  • \(\mathsf {Vrfy}(ck,c,[{\varvec{v}}]_r,Op)\): On input \(c=([{\varvec{y}}]_r,[\widetilde{{\varvec{c}}}]_r)\), this algorithm computes \([\widetilde{{\varvec{c}}}]_{r+1}\) and outputs 1 if \([{\varvec{t}}]_r:=[{\varvec{y}}-{\varvec{v}}]_r\) satisfies that \(1 \leftarrow \widetilde{\mathsf {Vrfy}}(ck,[\widetilde{{\varvec{c}}}]_{r+1},[{\varvec{t}}]_{r},[{\varvec{s}}]_r)\), else it outputs 0.

  • \(\mathsf {TrapdoorEquiv}(ck,tk,c, [{\varvec{v}}]_r, Op, [{\varvec{v}}']_r )\): On a commitment \(c=([{\varvec{y}}]_r,[\widetilde{{\varvec{c}}}]_r)\) with opening \(Op=([{\varvec{s}}]_{r})\), if \([{\varvec{t}}]_r:=[{\varvec{y}}-{\varvec{v}}]_r\) and \([{\varvec{t}}']_r:=[{\varvec{y}}-{\varvec{v}}']_r\), this algorithm computes \([\widetilde{{\varvec{c}}}]_{r+1}\) and runs the algorithm \(\widetilde{Op} \leftarrow \widetilde{\mathsf {TrapdoorEquiv}}(ck,tk,[\widetilde{{\varvec{c}}}]_{r+1},[{\varvec{t}}]_{r},[{\varvec{s}}]_{r},[{\varvec{t}}']_{r})\), and outputs \(\widetilde{Op}\).

A commitment is a vector of size \(k+d\) and an opening is of size k. The required security properties follow easily from the properties of the generalized Pedersen commitment.

Theorem 3

C is a perfectly hiding, computationally binding commitment.

Proof

Since the generalized Pedersen commitment is perfectly hiding, then \(([{\varvec{t}}+{\varvec{v}}]_r, \widetilde{\mathsf {Comm}}(\widetilde{ck},[{\varvec{t}}]_{r-1}))\) perfectly hides \([{\varvec{v}}]_r\) because \([{\varvec{t}}]_r\) acts as a one-time pad. Similarly, it is straightforward to see that the computationally binding property of C follows from the computationally binding property of the generalized Pedersen commitment.

Interestingly, this construction explains the two instantiations of “group-to-group” commitments given in [3] (see the full version [35] for more details).

6 A New Matrix Distribution and Its Applications

Both of our commitment schemes of Sect. 5 base security on some \(\mathcal {D}_{k+d,k}\)-\(\mathsf {KerMDH}\) assumptions, where d is the length of the committed vector. When \(d>1\), the only example of \(\mathcal {D}_{k+d,k}\)-\(\mathsf {MDDH}\) Assumption considered in [11] is the one corresponding to the uniform matrix distribution \(\mathcal {U}_{k+d,k}\), which is the weakest \(\mathsf {MDDH}\) Assumption of size \((k+d) \times k\). Another natural assumption for \(d>1\) is the one associated to the matrix distribution resulting from sampling from an arbitrary hard distribution \(\mathcal {D}_{k+1,k}\) (e.g., \(\mathcal {L}_{k}\)) and adding \(d-1\) new random rows. Following the same ideas in the proof of Lemma 8, it is easy to see that the resulting \(\mathcal {D}_{k+d,k}\)-\(\mathsf {MDDH}\) assumption is equivalent to the original \(\mathcal {D}_{k+1,k}\)-\(\mathsf {MDDH}\) assumption. However, for efficiency reasons, we would like to have a matrix distributions with an even smaller representation size. This motivates us to introduce a new family of matrix distributions, the \(\mathcal {CI}_{k,d}\) family.

Definition 9

(Circulant Matrix Distribution). We define \(\mathcal {CI}_{k,d}\) as

$$ \mathbf {{A}} = \left( \begin{matrix} a_1 &{}\;&{} &{}\;\;&{} &{}\;\;&{} 0 \\ \vdots &{}&{} a_1 &{}&{} &{}&{} \\ a_d &{}&{} \vdots &{}&{} \ddots &{}&{} \\ 1 &{}&{} a_d &{}&{} &{}&{} a_1 \\ &{}&{} 1 &{}&{} \ddots &{}&{} \vdots \\ &{}&{} &{}&{} \ddots &{}&{} a_d \\ 0 &{}&{} &{}&{} &{}&{} 1 \end{matrix}\right) \in \mathbbm {Z}_q^{(k+d)\times k}, \qquad \text {where }a_i \leftarrow \mathbbm {Z}_q $$

Matrix \(\mathbf {{A}}\) is such that each column can be obtained by rotating one position the previous column, which explains the name. Notice that when \(d=1\), \(\mathcal {CI}_{k,d}\) is exactly the symmetric cascade distribution \(\mathcal {SC}_k\), introduced in [11]. It can be shown that the representation size of \(\mathcal {CI}_{k,d}\), which is the number of parameters d, is the optimal among all hard matrix distributions \(\mathcal {D}_{k+d,k}\) defined by linear polynomials in the parameters. A similar argument shows that the circulant assumption is also optimal in the sense that it has a minimal number of nonzero entries among all hard matrix distributions \(\mathcal {D}_{k+d,k}\). It can also be proven that \(\mathcal {CI}_{k,d}\)-\(\mathsf {MDDH}\) holds generically in k-linear groups, which implies the hardness of the corresponding \(\mathsf {KerMDH}\) problem. To prove the generic hardness of the assumption, we turn to a result of Herold [20, Theorem 5.15 and corollaries]. It states that if all matrices produced by the matrix distribution are full-rank, \(\mathcal {CI}_{k,d}\) is a hard matrix distribution. Indeed, an algorithm solving the \(\mathcal {CI}_{k,d}\)-\(\mathsf {MDDH}\) problem in the generic k-linear group model must be able to compute a polynomial in the ideal \(\mathfrak {H}\subset \mathbbm {Z}_q[a_1,\ldots ,a_d,z_1,\ldots ,z_{k+d}]\) generated by all the \((k+1)\)-minors of \(\mathbf {{A}}\Vert {\varvec{z}}\) as polynomials in \(a_1,\ldots ,a_d,z_1,\ldots ,z_{k+d}\). Although this ideal can actually be generated using only a few of the minors, we need to build a Gröbner basis of \(\mathfrak {H}\) to reason about the minimum degree a nonzero polynomial in \(\mathfrak {H}\) can have. We show that, carefully selecting a monomial order, the set of all \((k+1)\)-minors of \(\mathbf {{A}}\Vert {\varvec{z}}\) form a Gröbner basis, and all these minors have total degree exactly \(k+1\). Therefore, all nonzero polynomials in \(\mathfrak {H}\) have degree at least \(k+1\), and then they cannot be evaluated by any algorithm in the generic k-linear group model. The full proof of both properties of \(\mathcal {CI}_{k,d}\) can be found in the full version [35].

As for other matrix distribution families, we can combine Corollary 1 and the techniques used in Lemma 9 to show that for any fixed \(d\ge 1\) the \(\mathcal {CI}_{k,d}\)-\(\mathsf {KerMDH}\) problem family has strictly increasing hardness.

Theorem 4

For any \(d\ge 1\) and for any \(k,\widetilde{k}\) such that \(\widetilde{k}< k\)

  1. 1.

    there exists a tight reduction, \(\mathcal {CI}_{\widetilde{k},d}\text {-}\mathsf {KerMDH}\Rightarrow \mathcal {CI}_{k,d}\text {-}\mathsf {KerMDH}\),

  2. 2.

    there is no black-box reduction in the generic group model in the opposite direction.

The new assumption gives new instantiations of the commitment schemes of Sect. 5 with public parameters of size d, independent of k. Further, because the matrix \(\mathbf {{A}}\leftarrow \mathcal {CI}_{k,d}\) has a many zero entries, the number of exponentiations computed by the \(\mathsf {Commit}\) algorithm, and the number of pairings of the verification algorithm is kd—as opposed to \(k(k+d)\) for the uniform assumption. This seems to be optimal—but we do not prove this formally.