When random tensors meet random matrices
Abstract
Relying on random matrix theory (RMT), this paper studies asymmetric order- spiked tensor models with Gaussian noise. Using the variational definition of the singular vectors and values of (Lim, 2005) [15], we show that the analysis of the considered model boils down to the analysis of an equivalent spiked symmetric block-wise random matrix, that is constructed from contractions of the studied tensor with the singular vectors associated to its best rank-1 approximation. Our approach allows the exact characterization of the almost sure asymptotic singular value and alignments of the corresponding singular vectors with the true spike components, when with ’s the tensor dimensions. In contrast to other works that rely mostly on tools from statistical physics to study random tensors, our results rely solely on classical RMT tools such as Stein’s lemma. Finally, classical RMT results concerning spiked random matrices are recovered as a particular case.
keywords:
[class=MSC]keywords:
, and
Notations:
denotes the set . The set of rectangular matrices of size is denoted . The set of square matrices of size is denoted . The set of -order tensors of size is denoted . The set of hyper-cubic tensors of size and order is denoted . The notation means that is a random tensor with i.i.d. Gaussian entries. Scalars are denoted by lowercase letters as . Vectors are denoted by bold lowercase letters as . denotes the canonical vector in with . Matrices are denoted by bold uppercase letters as . Tensors are denoted as . denotes the entry of the tensor . denotes the scalar product between and , the -norm of a vector is denoted as . denotes the spectral norm for tensors. denotes the contraction of tensor on the vectors given as arguments. Given some vectors with , the contraction denotes the resulting -th order tensor. denotes the -dimensional unit sphere.
1 Introduction
The extraction of latent and low-dimensional structures from raw data is a key step in various machine learning and signal processing applications. Our present interest is in those modern techniques which rely on the extraction of such structures from a low-rank random tensor model [1] and which extends the ideas from matrix-type data to tensor structured data. We refer the reader to [23, 21, 25] and the references therein which introduce an extensive set of applications of tensor decomposition methods to machine learning, including dimensionality reduction, supervised and unsupervised learning, learning subspaces for feature extraction, low-rank tensor recovery etc. Although random matrix models have been extensively studied and well understood in the literature, the understanding of random tensor models is still in its infancy and the ideas from random matrix analysis do not easily extend to higher-order tensors. Indeed, the resolvent notion (see Definition 1) which is at the heart of random matrices does not generalize to tensors. In our present investigation, we consider the spiked tensor model, which consists in an observed -order tensor of the form
(1) |
where , and . Note that the tensor noise is normalized by with the tensor dimensions, since the spectral norm of is of order from Lemma 4. One aims at retrieving the rank-1 component (or spike) from the noisy tensor , where can be seen as controlling the signal-to-noise ratio (SNR). The identification of the dominant rank-1 component is an important special case of the low-rank tensor approximation problem, a noisy version of the classical canonical polyadic decomposition (CPD) [10, 13]. Extensive efforts have been made to study the performance of low rank tensor approximation methods in the large dimensional regime – when the tensor dimensions , however considering symmetric tensor models where , , and assuming that the noise is symmetric [17, 20, 14, 9, 12, 8].
In particular, in the matrix case (i.e., ), the above spiked tensor model becomes a so-called spiked matrix model. It is well-known that in the large dimensional regime, there exists an order one critical value of the SNR below which it is information-theoretically impossible to detect/recover the spike, while above , it is possible to detect the spike and recover the corresponding components in (at least) polynomial time using singular value decomposition (SVD). This phenomenon is sometimes known as the BBP (Baik, Ben Arous, and Péché) phase transition [2, 5, 7, 19].
In the (symmetric) spiked tensor model for , there also exists an order one critical value111Depending on the tensor order . We will sometimes omit the dependence on if there is no ambiguity. (in the high-dimensional asymptotic) below which it is information-theoretically impossible to detect/recover the spike, while above recovery is theoretically possible with the maximum likelihood (ML) estimator. Computing the maximum likelihood in the matrix case corresponds to the computation of the largest singular vectors of the considered matrix which has a polynomial time complexity, while for , computing ML is NP-hard [17, 6]. As such, a more practical phase transition for tensors is to characterize the algorithmic critical value (which might depend on the tensor dimension ) above which the recovery of the spike is possible in polynomial time. Richard and Montanari [17] first introduced the symmetric spiked tensor model (of the form with symmetric ) and also considered the related algorithmic aspects. In particular, they used heuristics to highlight that spike recovery is possible, with Approximate Message Passing (AMP) or the tensor power iteration method, in polynomial time222Using tensor power iteration or AMP with random initialization. provided . This phase transition was later proven rigorously for AMP by [14, 12] and recently for tensor power iteration by [11].
Richard and Montanari [17] further introduced a method for tensor decomposition based on tensor unfolding, which consists in unfolding to an matrix for , to which a SVD is then performed. Setting , they predicted that their proposed method recovers successively the spike if . In a very recent paper by Ben Arous et al. [3], a study of spiked long rectangular random matrices333Number of rows are allowed to grow polynomially in the number of columns , i.e., . has been proposed under fairly general (bounded fourth-order moment) noise distribution assumptions. They particularly proved the existence of a critical SNR for which the extreme singular value and singular vectors exhibit a BBP-type phase transition. They applied their result for the asymmetric rank-one spiked model in Eq. (1) (with equal dimensions) using the tensor unfolding method, and found the exact threshold obtained by [17], i.e., for tensor unfolding to succeed in signal recovery.
For the asymmetric spiked tensor model in Eq. (1), few results are available in the literature (to the best of our knowledge only [3] considered this setting by applying the tensor unfolding method proposed by [17]). This is precisely the model we consider in the present work and our more general result is derived as follows. Given the asymmetric model from Eq. (1), the maximum likelihood (ML) estimator of the best rank-one approximation of is given by
(2) |
In the above equation, and the can be respectively interpreted as the generalization to the tensor case of the concepts of dominant singular value and associated singular vectors [15]. Following variational arguments therein, Eq. (2) can be reformulated using contractions of as
(3) |
the Lagrangian of which writes as with . Hence, the stationary points , with the ’s being unitary vectors, must satisfy the Karush-Kuhn-Tucker conditions, for
(4) |
An interesting question concerns the computation of the expected number of stationary points (local optima or saddle points) satisfying the identities in Eq. (4). [4] studied the landscape of a symmetric spiked tensor model and found that for the values of the objective function of all local maxima (including the global one) tend to concentrate on a small interval, while for the value achieved by the global maximum exits that interval and increases with . In contrast, very recently Goulart et al. [8] have studied an order symmetric spiked random tensor using a RMT approach, where it was stated that there exists a threshold such that for there exists a local optimum of the ML problem that correlates with the spike and such local optimum coincides with the global one for . We conjecture that such observations extend to asymmetric spiked tensors, namely that there exists an order one critical value above which the ML problem in Eq. (3) admits a global maximum. As for [8], our present findings do not allow to express such and its exact characterization is left for future investigation. However, for asymmetric spiked random tensors, we also exhibit a threshold such that for there exists a local optimum of the ML objective that correlates with the true spike. Figure 1 provides an illustration of the different thresholds of and see the last part of Subsection 3.2 for a more detailed discussion.

Main Contributions
Starting form the conditions in Eq. (4), we provide an exact expression of the asymptotic singular value and alignments , when the tensor dimensions with , where the tuple is associated to a local optimum of the ML problem verifying some technical conditions (detailed in Assumption 4). We conjecture that when the SNR is large enough, there is a unique local optimum verifying Assumption 4 and for which our results characterize the corresponding alignments. We further conjecture that coincides with the global maximum above some444Such a critical value has been characterized by [12] for symmetric tensors. See [8] for a detailed discussion about this aspect in the case of symmetric tensors. – that needs to be characterized.
Technically, we first show that the considered random tensor can be mapped to an equivalent symmetric random matrix , constructed through contractions of with directions among . Then, leveraging on random matrix theory, we first characterize the limiting spectral measure of and then provide estimates of the asymptotic alignments . We precisely show (see Theorem 8) that under Assumption 4, for , there exists such that for
(5) |
where satisfies with and
being the solution to the fixed point equation (for large enough); see Section B.11 for the existence of . Besides, for with555For arbitrary values of ’s, the upper bound of can be computed numerically as the minimum non-negative real number for which Algorithm 1 converges. for all
(6) |
that is (asymptotically) ceases to correlate with .
Remark 1.
Note that can be equivalently expressed as with . Note that such expression is defined for with . See details in Section B.13.
We highlight that in our formulation the threshold corresponds to the minimal value of the SNR above which the derived asymptotic formulas are algebraically defined, which may differ from the true phase transition of the ML problem: In the case of symmetric tensors, the results form [8] seem to indicate that is slightly below obtained by [12] where the ML problem was studied.

Figure 2 notably depicts the asymptotic alignments as in Eq. (5) when all the tensor dimensions are equal. Since our result characterizes the alignments for , it is not possible to recover the matrix case by simply setting since the equations are not defined for (see Remark 1). However, the matrix case is recovered by considering an order tensor () and then taking the limit (equivalent to the degenerate case which results in a spiked random matrix model), see Section 4 for more details. From Figure 2-(b), unlike the matrix case, i.e., , the predicted asymptotic alignments are not continuous for orders , this phenomenon has already been observed in the case of symmetric random tensors [12]. In particular, the predicted theoretical threshold in the matrix case coincides with the classical so-called BBP (Baik, Ben Arous, and Péché) phase transition [2, 5, 7, 19]. Moreover, our result for the matrix case characterizes the asymptotic alignments for the long rectangular matrices studied by [3] and we also recover the threshold for the case of tensor unfolding method (See remark 9). From a methodological view point, our results are derived based solely on Stein’s lemma without the use of complex contour integrals as classically performed in RMT. In essence, we follow the approach in [8] and further introduce a new object (the mapping defined subsequently in Eq. (31)) that simplifies drastically the use of RMT tools.
The remainder of the paper is organized as follows. Section 2 recalls some fundamental random matrix theory tools. In Section 3, we study asymmetric spiked tensors of order where we provide the main steps of our approach. Section 4 characterizes the behavior of spiked random matrices given our result on spiked -order tensor models from Section 3. The generalization of our results to arbitrary -order tensors is then presented in Section 5. Section 6 discusses the application of our findings to arbitrary rank- tensors with mutually orthogonal components. Further discussions are presented in Section 7. In Appendix A, we provide some simulations to support our findings and discuss some algorithmic aspects. Finally, Appendix B provides the proofs of the main developed results.
2 Random matrix theory tools
Before digging into our main findings, we briefly recall some random matrix theory results that are at the heart of our analysis. Specifically, we recall the resolvent formalism which allows to assess the spectral behavior of large symmetric random matrices. In particular, given a symmetric matrix and denoting its eigenvalues with corresponding eigenvectors for , its spectral decomposition writes as
In the sequel we omit the dependence on by simply writing and if there is no ambiguity.
2.1 The resolvent matrix
We start by defining the resolvent of a symmetric matrix and present its main properties.
Definition 1.
Given a symmetric matrix , the resolvent of is defined as
where is the spectrum of .
The resolvent is a fundamental object since it retrieves the spectral characteristics (spectrum and eigenvectors) of . It particularly verifies the following property which we will use extensively to derive our main results,
(7) |
The above identity, coupled with Stein’s Lemma (Lemma 1 in Section 2.3 below), is a fundamental tool used to derive fixed point equations that allow the evaluation of functionals of interests involving . Another interesting property of the resolvent concerns its spectral norm, which we denote by . Indeed, if the spectrum of has a bounded support, then the spectral norm of is bounded. This is a consequence of the inequality
(8) |
where denotes the distance of to a set. The resolvent encodes rich information about the behavior of the eigenvalues of through the so-called Stieltjes transform which we describe subsequently.
2.2 The Stieltjes transform
Random matrix theory, originally, aims at describing the limiting spectral measure of random matrices when their dimensions grow large. Typically, under certain technical conditions on , the empirical spectral measure of defined as
(9) |
where is a Dirac mass on , converges to a deterministic probability measure . To characterize such asymptotic measure, the Stieltjes transform (defined below) approach is a widely used tool.
Definition 2 (Stieltjes transform).
Given a probability measure , the Stieltjes transform of is defined by
Particularly, the Stieltjes transform of is closely related to its associated resolvent through the algebraic identity
The Stieltjes transform has several interesting analytical properties, among which, we have
-
1.
is complex analytic on its definition domain and if ;
-
2.
is bounded for if is bounded, since ;
-
3.
Since , is monotonously increasing for .
The Stieltjes transform admits an inverse formula which provides access to the evaluation of the underlying probability measure , as per the following theorem.
Theorem 1 (Inverse formula of the Stieltjes transform).
Let be some continuity points of the probability measure , then the segment is measurable with , precisely
Moreover, if admits a density function at some point , i.e., is differentiable in a neighborhood of with , then we have the inverse formula
When it comes to large random matrices, the Stieltjes transform admits a continuity property, in the sense that if a sequence of random probability measures converges to a deterministic measure then the corresponding Stieltjes transforms converge almost surely to a deterministic one, and vice-versa. The following theorem from [24] states precisely this continuity property.
Theorem 2 (Stieltjes transform continuity).
A sequence of random probability measures , supported on with corresponding Stieltjes transforms , converges almost surely weakly to a deterministic measure , with corresponding Stieltjes transform if and only if almost surely, for all .
2.3 Gaussian calculations
The following lemma by Stein (also called Stein’s identity or Gaussian integration by parts) allows to replace the expectation of the product of a Gaussian variable with a differentiable function by the variance of that variable times the expectation of .
Lemma 1 (Stein’s Lemma [22]).
Let and a continuously differentiable function having at most polynomial growth, then
when the above expectations exist.
We will further need the Poincaré inequality which allows to control the variance of a functional of Gaussian random variables.
Lemma 2 (Poincaré inequality).
Let a continuously differentiable function having at most polynomial growth and a collection of i.i.d. standard Gaussian variables. Then,
3 The asymmetric rank-one spiked tensor model of order
To best illustrate our approach, we start by considering the asymmetric rank-one tensor model of order of the form
(10) |
where and both have dimensions and . We assume that and are on the unit spheres , and respectively, is a Gaussian noise tensor with i.i.d. entries independent from .
3.1 Tensor singular value and vectors
According to Eq. (4), the -singular value and vectors, corresponding to the best rank-one approximation of , satisfy the identities
(11) |
where we denoted and . Furthermore, the singular value can be characterized through the contraction of along all its singular vectors , i.e.
(12) |
3.2 Associated random matrix model
We follow the approach developed in [8] which consists in studying random matrices that are obtained through contractions of a random tensor model, and extend it to the asymmetric spiked tensor model of Eq. (10). Indeed, it has been shown in [8] that the study of a rank-one symmetric spiked tensor model boils down to the analysis of the symmetric random matrix666The contraction of the tensor with its eigenvector . where stands for the eigenvector of corresponding to the best symmetric rank-one approximation of , i.e.,
(13) |
In the asymmetric case of Eq. (10), the choice of a “relevant” random matrix to study is not trivial since the corresponding contractions and yield asymmetric random matrices which present more technical difficulties from the random matrix theory perspective, therefore the extension of the approach in [8] to asymmetric tensors is not straightforward. We will see in the following that such a choice of the relevant matrix to study an asymmetric is naturally obtained through the use of the Pastur’s Stein approach [16].
As described in [8] and since the singular vectors and depend statistically on , the first technical challenge consists in expressing the derivatives of the singular vectors and w.r.t. the entries of the Gaussian noise tensor . Indeed, one can show that there exists a differentiable mapping that maps to singular-value and vectors of , since the components of and are bounded and has polynomial growth. Indeed, we have the following Lemma which is analog to [8, Lemma 8] and which justifies the application of Stein’s Lemma subsequently.
Lemma 3.
There exists an almost everywhere continuously differentiable function such that is singular-value and vectors of (for almost every ).
Proof.
The proof relies on the same arguments as [8, Lemma 8]. ∎
Calculus (see details in B.1) show that deriving the identities in Eq. (11) w.r.t. an entry of with results in
(14) |
where we recall that denotes the canonical vector in with . We further have the identity
(15) |
Denoting by the symmetric block-wise random matrix which appears in the matrix inverse in Eq. (14), the derivatives of and are therefore expressed in terms of the resolvent evaluated on , and we will see subsequently that the assessment of the spectral properties of boils down to the estimation of the normalized trace . As such, the matrix provides the associated random matrix model that encodes the spectral properties of . We will henceforth focus our analysis on this random matrix, in order to assess the spectral behavior of . More generally, we will be interested in studying random matrices from the -order block-wise tensor contraction ensemble for defined as
(16) |
where is the mapping
(17) |
We associate tensor to the random matrix
(18) |
where
and and are on the unit spheres , and respectively. Note that is symmetric and behaves as a so-called spiked random matrix model where the signal part correlates with the true signals and for a sufficiently large . However, the noise part (i.e., the term in Eq. (18)) has a non-trivial structure due to the statistical dependencies between the singular vectors and the tensor noise . Despite this statistical dependency, we will show in the next subsection that the spectral measure of (and that of ) converges to a deterministic measure (see Theorem 4) that coincides with the limiting spectral measure of where are unit vectors and independent of . Furthermore, the matrix admits as an eigenvalue (regardless of the value of ), which is a simple consequence of the identities in Eq. (11), since
(19) |
Note however that the expression in Eq. (14) exists only if is not an eigenvalue of . As discussed in [8] in the symmetric case, such condition is related to the locality of the maximum ML estimator. Indeed, we first have the following remark that concerns the Hessian of the underlying objective function.
Remark 2.
Recall the Lagrangian of the ML problem as and denote . If is a singular value of with associated singular vectors , then is an eigenpair of the Hessian evaluated at . Indeed, , thus since (from Eq. (19)).
From [18, Theorem 12.5], a necessary condition for to be a local maximum of the ML problem is that
which yields (by remark 2) the condition
(20) |
As such, for , must be the largest eigenvalue of as per Eq. (19), while its second largest eigenvalue cannot exceed as shown by Eq. (20). Thus our analysis applies only to some local optimum of the ML problem for which the corresponding singular value lies outside the bulk of , namely we suppose777We will find that such assumption is satisfied provided for some that we will determine that there exists a tuple verifying the identities in Eq. (11) such that is not an eigenvalue of . Moreover, this allows the existence of the matrix inverse in Eq. (14) and it is not restrictive in the sense that it is satisfied for a sufficiently large value of the SNR .
In the sequel, for any verifying the identities in Eq. (11), we find that the largest eigenvalue is always isolated from the bulk of independently of the SNR . In addition, there exists such that for , the observed spike is not informative (in the sense that the corresponding alignments will be null) and is visible (an isolated eigenvalue appears in the spectrum of , see Figure 4) because of the statistical dependencies between and . The same phenomenon has been observed in the case of symmetric random tensors by [8].
3.3 Limiting spectral measure of block-wise -order tensor contractions
We start by presenting our first result which characterizes the limiting spectral measure of the ensemble with . We characterize this measure in the limit when the tensor dimensions grow large as in the following assumption.
Assumption 1 (Growth rate).
As , the dimension ratios , and .
We have the following theorem which characterizes the spectrum of with arbitrary deterministic unit vectors .
Theorem 3.
Let be a sequence of random asymmetric Gaussian tensors and a sequence of deterministic vectors of increasing dimensions, following Assumption 1. Then the empirical spectral measure of converges weakly almost surely to a deterministic measure whose Stieltjes transform is defined as the solution to the equation such that for where, for satisfies for .
Proof.
See Appendix B.2. ∎
Remark 3.
The equation yields a polynomial equation of degree in which can be solved numerically or via an iteration procedure which converges to a fixed point for large enough.

In particular, in the cubic case when , the empirical spectral measure of converges to a semi-circle law as precisely stated by the following Corollary of Theorem 3.
Corollary 1.
Given the setting of Theorem 3 with , the empirical spectral measure of converges weakly almost surely to the semi-circle distribution supported on , whose density and Stieltjes transform write respectively as
Proof.
See Appendix B.3. ∎
Figure 3 depicts the spectrum of with and independent unit vectors , it particularly illustrates the convergence in law of this spectrum when the dimensions grow large.
Remark 4.
More generally, the spectral measure of converges to a deterministic measure with connected support if is almost surely full rank, i.e., if since the rank of is . In contrast if it is not full rank, its spectral measure converges to a deterministic measure with unconnected support (see the case of matrices in Corollary 4 subsequently).
The analysis of the tensor relies on describing the spectrum of where the singular vectors depend statistically on (the noise part of ). Despite these dependencies, it turns out that the spectrum of converges in law to the same deterministic measure described by Theorem 3. Besides, we need a further technical assumption on the singular value and vectors of .
Assumption 2.
We assume that there exists a sequence of critical points satisfying Eq. (11) such that , , , with and .
We precisely have the following result.
Theorem 4.
Proof.
See Appendix B.4. ∎
However, the statistical dependency between and exhibits an isolated eigenvalue in the spectrum of at the value independently of the value of the SNR , which is a consequence of Eq. (19). Figure 4 depicts iterations of the power iteration method where the leftmost histogram corresponds to the fixed point solution, where one sees that the spike converges to the value .

Remark 5.
Let us denote the blocks of the resolvent of as
(21) |
where and . The following corollary of Theorem 4 will be useful subsequently.
Corollary 2.
Recall the setting and notations of Theorem 4. For all , we have
3.4 Concentration of the singular value and the alignments
When the dimensions of grow large under Assumption 1, the singular value and the alignments and converge almost surely to some deterministic limits. This can be shown by controlling the variances of these quantities using the Poincaré’s inequality (Lemma 2). Precisely, for , invoking Eq. (15) we have
Bounding higher order moments of similarly allows to obtain the concentration of , e.g. by Chebyshev’s inequality, we have for all
Similarly, with Eq. (14), there exists such that for all
For the remainder of the manuscript, we denote the almost sure limits of and the of alignments and respectively as
(22) |
3.5 Asymptotic singular value and alignments
Having the concentration result from the previous subsection, it remains to estimate the expectations and . Usually, the evaluation of these quantities using tools from random matrix theory relies on computing Cauchy integrals involving (the resolvent of ). Here, we take a different (yet analytically simpler) approach by taking directly the expectation of the identities in Eq. (11) and Eq. (12), then applying Stein’s Lemma (Lemma 1) and Lemma 3. For instance, for , we have
From Eq. (14), when , it turns out that the only contributing terms888Yielding non-vanishing terms in the expression of . of the derivatives and in the above sum are respectively
This yields
Therefore, the almost sure limit as of satisfies
where and are defined in Eq. (22). From Eq. (11), proceeding similarly as above with the identities
we obtain the following result.
Theorem 5.
Proof.
See Appendix B.5. ∎
Remark 6.
A more compact expression for the alignments is provided in Theorem 8.

Figure 5 depicts the predicted asymptotic dominant singular value of and the corresponding alignments from Theorem 5. As we can see, the result of Theorem 5 predicts that a non-zero correlation between the population signals and their estimated counterparts is possible only when , which corresponds to the value after which starts to increase with .
Remark 7.
Note that, given , the inverse formula expressing in terms of is explicit. Specifically, we have . In particular, this inverse formula provides an estimator for the SNR given the largest singular value of .
3.6 Cubic -order tensors: case
In this section, we study the particular case where all the tensor dimensions are equal. As such, the three alignments converge almost surely to the same quantity. In this case, the almost sure limits of and can be obtained explicitly in terms of the signal strength as per the following corollary of Theorem 5.
Proof.
See Appendix B.6. ∎
Figure 6 provides plots of the almost sure limits of the singular value and alignments when is cubic as per Corollary 3 (see Subsection A.2 for simulations supporting the above result). In particular, this result predicts a possible correlation between the singular vectors and the underlying signal components above the value with corresponding singular value with an alignment . In addition, we can easily check from the formulas above that and for large values of . Besides, for values of around the value , the expression of admits the following expansion
whereas the corresponding alignment expends as

4 Random tensors meet random matrices
In this section, we investigate the application of Theorem 5 to the particular case of spiked random matrices. Indeed, for instance when the spiked tensor model in Eq. (10) becomes a spiked matrix model which we will now denote as
(23) |
where again and are on the unit spheres and respectively, and is a Gaussian noise matrix with i.i.d. entries .
Our approach does not apply directly to the matrix above. Indeed, the singular vectors of corresponding to its largest singular value satisfy the identities
(24) |
and deriving w.r.t. an entry of will result in
(25) |
Now since is an eigenvalue of , the matrix is not invertible, which makes our approach inapplicable for . However, we can retrieve the behavior of the spiked matrix model in Eq. (23) by considering an order tensor and setting for instance as we will discuss subsequently.
4.1 Limiting spectral measure
Given the result of Theorem 5, the asymptotic singular value and alignments for the spiked matrix model in Eq. (23) correspond to the particular case . We start by characterizing the corresponding limiting spectral measure, we have the following corollary of Theorem 4 when .
Corollary 4.
Proof.
See Appendix B.7. ∎

Remark 8.
Corollary 4 describes the limiting spectral measure of for , and (a scalar), which is equivalent to random matrices of the form with .
Figure 7 depicts the limiting spectral measures as per Corollary 4 for various values of (see also simulations in Appendix A.1). In particular, the limiting measure is a semi-circle law for square matrices (i.e., ), while it decomposes into a two-mode distribution101010With a Dirac at weighted by , not shown in Figure 7. for due to the fact that the underlying random matrix model is not full rank ( if of rank ).
4.2 Limiting singular value and alignments
Given the limiting Stieltjes transform from the previous subsection (Corollary 4), the limiting largest singular value of (in Eq. (23)) and the alignments of the corresponding singular vectors are obtained thanks to Theorem 5, yielding the following corollary.
Corollary 5.
Under Assumption 1 with , we have for
where are the singular vectors of corresponding to its largest singular value and is given by
Besides, for , and .
Proof.
See Appendix B.8. ∎

Figure 8 provides the curves of the asymptotic singular value and alignments of Corollary 5 (see Subsection A.1 for simulations supporting the above result). Unlike the tensor case from the previous section, we see that the asymptotic alignments are continuous and a positive alignment is observed for which corresponds to the classical BBP phase transition of spiked random matrix models [2, 5, 19].
Remark 9 (Application to tensor unfolding).
The tensor unfolding method consists in estimating the spike components of a given tensor with by applying an SVD to its unfolded matrices (for ) of size . Applying Corollary 5 to this case predicts a phase transition at with111111 In our assumptions, we assume that is some constant that does not depend on the tensor dimensions. Still, we believe that it can be relaxed in the same vein as in [4]. , which yields . After re-scaling the noise component (by multiplying Eq. (23) by with ) this yields , i.e., the phase transition for tensor unfolding obtained by [3] (see Theorem 3.3 therein). More generally, for any order- tensor of arbitrary dimensions , the tensor unfolding method succeeds provided that .
5 Generalization to arbitrary -order tensors
We now show that our approach can be generalized straightforwardly to the -order spiked tensor model of Eq. (1). Indeed, from Eq. (4), the -singular value and vectors , corresponding to the best rank-one approximation of the -order tensor in Eq. (1), satisfy the identities
(26) |
5.1 Associated random matrix ensemble
Let denote the matrix obtained by contracting the tensor with the singular vectors in , i.e.,
(27) |
As in the order case, from Eq. (26), the derivatives of the singular vectors with respect to the entry of the noise tensor express as
(28) |
where , and the derivative of w.r.t. writes as
(29) |
As such, the associated random matrix model of is the matrix appearing in the resolvent in Eq. (28). More generally, the -order block-wise tensor contraction ensemble for is defined as
(30) |
where is the mapping
(31) |
with .
5.2 Limiting spectral measure of block-wise -order tensor contractions
In this section, we characterize the limiting spectral measure of the ensemble for in the limit when all tensor dimensions grow as per the following assumption.
Assumption 3.
For all , assume that with .
We thus have the following result which characterizes the spectrum of for any deterministic unit norm vectors .
Theorem 6.
Let be a sequence of random asymmetric Gaussian tensors and a sequence of deterministic vectors of increasing dimensions, following Assumption 3. Then the empirical spectral measure of converges weakly almost surely to a deterministic measure whose Stieltjes transform is defined as the solution to the equation such that for where, for satisfies for .
Proof.
See Appendix B.9. ∎
Algorithm 1 provides a pseudo-code to compute the Stieltjes transform in Theorem 6 through an iterative solution to the fixed point equation . In particular, as for the order case, when all the tensor dimensions are equal (i.e., for all ), the spectral measure of converges to a semi-circle law. We have the following corollary of Theorem 6.
Corollary 6.
Proof.
See Appendix B.10. ∎

Figure 9 provides an illustration of the convergence in law (when the dimension grow large) of the spectrum of with a th-order tensor and independently sampled from the unit spheres respectively.
Remark 11.
is almost surely of rank . As in the order case, when is almost surely full rank, its spectral measure converges to a semi-circle law with connected support, while in general the limiting spectral measure has unconnected support.
The result of Theorem 6 still holds for where stand for the singular vectors of defined through Eq. (26). Indeed, the statistical dependencies between these singular vectors and the noise tensor do not affect the convergence of the spectrum of to the limiting measure described by Theorem 6. We further need the following assumption.
Assumption 4.
We assume that there exists a sequence of critical points satisfying Eq. (26) such that , with and .
Theorem 7.
Proof.
See Appendix B.12. ∎

Figure 10 depicts the spectrum of for an order tensor with . As we saw previously, an isolated eigenvalue pops out from the continuous bulk of because of the statistical dependencies between the tensor noise and the singular vectors . More generally, for an order- tensor , the spectrum of admits an isolated spike at the value independently of the signal strength . Indeed, is an eigenvalue of with corresponding eigenvector the concatenation of the singular vectors , i.e.,
(32) |
5.3 Asymptotic singular value and alignments of hyper-rectangular tensors
Similarly to the -order case studied previously, when the dimensions of grow large at a same rate, its singular value and the corresponding alignments for concentrate almost surely around some deterministic quantities which we denote and respectively. Applying again Stein’s Lemma (Lemma 1) to the identities in Eq. (26), we obtain the following theorem which characterizes the aforementioned deterministic limits.
Theorem 8.
Proof.
See Appendix B.13. ∎
Remark 12.
As in the order case, note that the inverse formula expressing in terms of is explicit. Specifically, we have . In particular, this inverse formula provides an estimator for the SNR given the largest singular value of . Algorithm 2 provides a pseudo-code to compute the asymptotic alignments.

Figure 11 depicts the asymptotic singular value and alignments for an order tensor as per Theorem 8. As discussed previously, the predicted alignments are discontinuous and a strictly positive correlation between the singular vectors and the underlying signals is possible for the considered local optimum of the ML estimator only above the minimal signal strength in the shown example. In the case of hyper-cubic tensors, i.e., all the dimensions are equal, Figure 12 depicts the minimal SNR values in terms of the tensor order and the corresponding asymptotic singular value and alignments. As such, when the tensor order increases the minimal SNR and singular value converge respectively to and , while the corresponding alignment121212corresponding to the minimal theoretical SNR . gets closer to . The expressions of and for hyper-cubic tensors of order are explicitly given as
(33) |

6 Generalization to rank tensor with orthogonal components
In this section we discuss the generalization of our previous findings to rank spiked tensor model with of the form
(34) |
where are the signal strengths, for , and . Supposing that the components are mutually orthogonal, i.e., for , the above -rank spiked tensor model can be treated through equivalent rank-one tensors defined for each component independently, by applying the result of the rank-one case established in the previous section. Precisely, the best rank- approximation of corresponds to
(35) |
and the ’s correlate with ’s as a result of the uniqueness of orthogonal tensor decomposition (see Theorem 4.1 in [1]). Therefore, the study of boils down to the study of the random matrices for which behave as the rank-one case treated previously with signal strength respectively. Indeed, since by definition and given the orthogonality condition on the ’s, for the inner product in high dimension, as such
(36) |
where . Meaning, that the study of is equivalent to consider the rank-one spiked tensor model .
7 Discussion
In this work, we characterized the asymptotic behavior of spiked asymmetric tensors by mapping them to equivalent (in spectral sense) random matrices. Our starting point is mainly the identities in Eq. (26) which are verified by all critical points of the ML problem. Quite surprisingly and as also discussed in [8] for symmetric tensors, we found that our asymptotic equations describe precisely the maximum of the ML problem which correlates with the true spike. Extrapolating the findings from [12, 8] in the symmetric case, we conjuncture the existence of an order one threshold above which our equations describe the behavior of the global maximum of the ML problem. Unfortunately, it is still unclear how we can characterize such with our present approach, which remains an open question. The same question concerns the characterization of the algorithmic threshold which is more interesting from a practical standpoint as computing the ML solution is NP-hard for .
In the present work, our results were derived under a Gaussian assumption on the tensor noise components. We believe that the derived formulas are universal in the sense that they extend to other distributions provided that the fourth order moment is finite (as assumed by [3] for long random matrices). Other extensions concern the generalization to higher-ranks with arbitrary components and possibly correlated noise components since the present RMT-tools are more flexible than the use of tools form statistical physics.
[Acknowledgments] This work was supported by the MIAILargeDATA Chair at University Grenoble Alpes led by R. Couillet and the UGA-HUAWEI LarDist project led by M. Guillaud. We would like to thank Henrique Goulart, Pierre Common and Gérard Ben-Aroud for valuable discussions on the topic of random tensors.
Appendix A Simulations
In this section we provide simulations to support our findings.
A.1 Matrix case
We start by considering the spiked random matrix model of the form
(37) |
and are unitary vectors of dimensions and respectively. Figure 13 depicts the spectrum of for for different values of the dimensions , and the predicted limiting spectral measure as per Corollary 4. Figure 14 shows a comparison between the asymptotic singular value and alignments obtained in Corollary 5 and their simulated counterparts through SVD applied on . Figure 15 further provides comparison between theory and simulations in the case of long random matrices (), which corresponds to tensor unfolding as per Remark 9, where a perfect matching is also observed between theory and simulations.



A.2 Order tensors
Now we consider a -order random tensor model of the form
(38) |
and are unitary vectors of dimensions and respectively. In our simulations the estimation of the singular vectors of is performed using the power iteration method described by Algorithm 3. We consider three initialization strategies for Algorithm 3:
-
(i)
Random initialization by randomly sampling from the unitary spheres and respectively. We refer to this strategy in the figures legends by “Random init.”.
-
(ii)
Initialization with the true components . We refer to this strategy in the figures legends by “Init. with ”.
- (iii)
Figure 16 provides a comparison between the asymptotic singular value and alignments of (obtained by Corollary 3) with their simulation counterparts where the singular vectors are estimated by Algorithm 3 with the initialization strategies (i) in yellow dots and (ii) in green dots. As we can see, as the tensor dimensions grow, the numerical estimates approach their asymptotic counterparts for (ii). Besides, the random initialization (i) yields poor convergence for around its minimal value when the tensor dimension is large enough (see ). This phenomenon is related to the algorithmic time complexity of the power method which is known to succeed in recovering the underlying signal in polynomial time provided that [17, 6], which is also noticeable in our simulations (the algorithmic phase transition seems to grow with the tensor dimensions). Figure 17 further depicts comparison between theory and simulations when Algorithm 3 is initialized following strategy (iii) which allows to follow the trajectory of the global maximum of the ML problem, where we can notice a good matching between the asymptotic curves and their simulated counterparts.










Appendix B Proofs
B.1 Derivative of tensor singular value and vectors
Deriving the identities in Eq. (11) and Eq. (12) w.r.t. the entry of the tensor noise , we obtain the following set of equations
Writing as , we can apply again the identities in Eq. (11) which results in . Doing similarly with and , we have
Furthermore, since , we have
Thus the derivative of writes simply as
Hence, we find that
Yielding the expression in Eq. (14). The same calculations apply to the more general -order tensor case yielding the identity in Eq. (28).
B.2 Proof of Theorem 3
Denote the matrix model as
where we recall and are independent of . We further denote the resolvent matrix of as
In order to characterize the limiting Stieltjes transform of , we need to estimate the quantity (as a consequence of Theorem 2). We further introduce the following limits
From the identity in Eq. (7), we have
from which we particularly have
or
(39) |
We thus need to compute the expectations of and , which develop as
where the last equality is obtained by applying Stein’s lemma (Lemma 1). For continuing the derivations, we need to express the derivative of the resolvent with respect to an entry of the tensor noise . Indeed, since , we have
from which we get
where
and we finally obtain the following derivatives
In particular,
Going back to the computation of , we will see that the only contributing term in the derivative is . Indeed,
Now, since the vectors are of bounded norms and assuming is of bounded spectral norm (see condition in Eq. (8)), under Assumption 1 (as ), the terms , , , and are vanishing almost surely. As such, we find that
Similarly, we find that
From Eq. (39), satisfies
where we recall . Similarly, and satisfy
where we recall again and . Moreover, by definition, , thus we have for each
yielding
with solution of the equation satisfying for with (see Property 2 of the Stieltjes transform in Subsection 2.2).
B.3 Proof of Corollary 1
B.4 Proof of Theorem 4
Given the random tensor model in Eq. (10) and its singular vectors characterized by Eq. (11), we denote the associated random matrix model as
where
We further denote the resolvents of and respectively as
By Woodbury matrix identity (Lemma 5), we have
(40) |
In particular, taking the normalized trace operator, we get
since the matrix is of bounded spectral norm (assuming is bounded, see condition in Eq. (8)) and being of finite size ( matrix). As such, the asymptotic spectral measure of is the same as the one of which can be estimated though . Comparing to the result from Appendix B.2, now the singular vectors depend statistically on the tensor noise which needs to be handled.
From the identity Eq. (7), we have , from which
(41) |
We thus need to compute the expectations of and . In particular,
where the last equality is obtained by Stein’s lemma (Lemma 1). Due to the statistical dependency between and , the above sum decomposes in two terms which are
where the first term has already been handled in the previous subsection (if replacing with ). We now show that the second term is asymptotically vanishing under Assumption 1. Indeed, by Eq. (14), we have
As such decomposes in three terms , where
as , since the singular vectors are of bounded norms and assuming the resolvent and are of bounded spectral norms ( has bounded spectral norm by Assumption 2 and Eq. (8)). Similarly, we further have
And finally,
Therefore,
As in the previous subsection, the derivative of w.r.t. the entry expresses as but now with
where is of vanishing spectral norm. Indeed, from Eq. (14), there exists independent of such that yielding that the spectral norm of is bounded by for some constant independent of . Therefore, we find that (with the almost sure limits of and respectively), thus yielding the same limiting Stieltjes transform as the one obtained in Appendix B.2.
B.5 Proof of Theorem 5
Given the identities in Eq. (11), we have
(42) |
with and converging almost surely to their asymptotic limits and respectively given the concentration properties in Subsection 3.4. To characterize such limits we need to evaluate the expectation of . Indeed,
where the last equality is obtained by Stein’s lemma (Lemma 1). From Eq. (14), we have
Hence decomposes in three terms and . The terms and will be vanishing asymptotically and only contains non-vanishing terms. Indeed,
as since and are of bounded norms and being of bounded spectral norm for outside the support of (through the identity in Eq. (40)). Similarly with , we have
Now is not vanishing, precisely,
where the last line results from the fact that as we saw in the previous subsection. Similarly, we find that
Therefore, by Eq. (42), the almost sure limits and satisfy the equation
Hence,
with . Similarly, we find that
with . Solving the above system of equations provides the final asymptotic alignments. Proceeding similarly with Eq. (12) we obtain an estimate of the asymptotic singular value , thereby ending the proof.
B.6 Proof of Corollary 3
By Corollary 1, the limiting Stieltjes transform is given by
and since with all equal, then for all . Hence, for each , defined in Theorem 5 writes as
and is solution to the equation
First we compute the critical value of by solving the above equation in and taking the limit when tends to the right edge of the semi-circle law (i.e., ). Indeed, solving the above equation in yields
hence
Now to express in terms of , we solve the equation in and choose the unique non-decreasing (in ) and positive solution, which yields
Plugging the above expression of into the expressions of the asymptotic alignments in Theorem 5, we obtain for all
yielding the final result.
B.7 Proof of Corollary 4
Setting and , we get
And since , then satisfies the equation
the solution of which belongs to
thus the limiting Stieltjes transform (having ) is given by
In particular, the edges of the support of the corresponding limiting distribution are the roots of , yielding
And the density function of is obtained by computing the limit yielding
where . The Dirac component in the above expression corresponds to the fact that the corresponding matrix model is of rank .
B.8 Proof of Corollary 5
From Subsection B.7, plugging the expression of the limiting Stieltjes transform into the expressions defining and yields
Thus, by Theorem 5, we have
Then solving the equation in provides the almost sure limit of in terms of . Specifically,
Plugging the above expression of into by replacing with the expression of provides the asymptotic alignment as , with given by the following expression
From the identities
simplifies as
The asymptotic alignment is given by since the dimension ratio of the component is . Moreover, the critical value of is obtained by solving the equation , i.e., when the limiting singular value gets closer to the right edge of the support of the corresponding limiting spectral measure (see Corollary 4). Finally, similarly as above, we can check that is equal to for (i.e., the alignment along the third dimension is , corresponding to the component in Eq. (10)).
B.9 Proof of Theorem 6
Denote the matrix model as
with and are independent of . We further denote the resolvent of as
By Borel-Cantelli lemma, we have and for all , . Applying the identity in Eq. (7) to the symmetric matrix we get , from which we particularly get
or
(43) |
where we recall that .
We thus need to compute the expectation of which develops as
where the last equality follows from Stein’s lemma (Lemma 1). In particular, as in Appendix B.2 for the -order case, it turns out that the only contributing term in the derivative is with the other terms yielding quantities of order . Therefore, we find that
From Eq. (43), satisfies with . Similarly, for all , satisfies with . And since, , we have for each , , yielding
with solution to the equation satisfying for .
B.10 Proof of Corollary 6
Given Theorem 6 and setting for all , we have , thus satisfies the equation
Solving in yields
and the limiting Stieltjes transform with is
B.11 Existence of
Define the following function for and
with . By concavity of the function we have for all . Therefore, is bounded as
From Section B.10, we have, for any , there exists such that . Besides, we further have and hence by continuity, there exists such that for large enough (e.g., ).
B.12 Proof of Theorem 7
Given the random tensor model in Eq. (1) and its singular vectors characterized by Eq. (26), we denote the associated random matrix model as
where , with entries
We further denote the resolvent of and respectively as
Similarly as in -order case, by Woodbury matrix identity (Lemma 5), we have
since the perturbation matrix is of bounded spectral norm (if is bounded, see condition in Eq. (8)) and has finite size ( matrix). Therefore, the characterization of the spectrum of boils down to the estimation of . Now we are left to handle the statistical dependency between the tensor noise and the singular vectors of . Recalling the proof of Appendix B.9, we have again
with . Taking the expectation of yields
where we already computed the first term () in Appendix B.9. Now we will show that is asymptotically vanishing under Assumption 3. Indeed, by Eq. (28), the higher order terms arise from the term , we thus only show that the contribution of this term is also vanishing. Precisely,
where denotes the vector with entries . As such, is vanishing asymptotically since the singular vectors are unitary and since their entries are bounded by .
We finally need to check that the derivative of w.r.t. the entry has the same expression asymptotically as the one in the independent case of Appendix B.9. Indeed, we have with
where and is of vanishing norm. Indeed, by Eq. (28), there exists independent of such that , therefore, the spectral norm of is bounded by for some constant independent of . Finally, (with the almost sure limits of and respectively), hence yielding the same limiting Stieltjes transform as the one obtained in Appendix B.9.
B.13 Proof of Theorem 8
Given the identities in Eq. (26), we have for all
with and concentrate almost surely around their asymptotic denoted and respectively. Taking the expectation of the first term and applying Stein’s lemma (Lemma 1), we get
where the only contributing term in the expression of from Eq. (28) is which yields
Therefore, the almost sure limits and for each satisfy
therefore
since . To solve the above equation, we simply write and by omitting the dependence on . We therefore have
from which we have
thus
and we remark that , hence is given by
which ends the proof.
Alternative expression of
From the above, we have and
Therefore,
with since satisfies . Hence, we find
B.14 Additional lemmas
Lemma 4 (Spectral norm of random Gaussian tensors).
Let , then the spectral norm of can be bounded, with probability at least for , as
Proof.
By definition, the spectral norm of is given as
(44) |
For some , let be -nets of respectively. Since is compact, there exists a maximizer of Eq. (44). And with the -net argument, there exists for each such that
such that for . Therefore, one has
For , one has . As such, the spectral norm of can be bounded as
(45) |
Since the entries of are i.i.d. standard Gaussian random variables, we have , hence using Hoeffding’s inequality we have for any with
Minimizing the right-hand side w.r.t. yields . With the same arguments, we further have . And taking the union of the two cases yields
Back to Eq. (45), since , involving the union bound gives us
which yields the final bound for an appropriate choice of . ∎
Lemma 5 (Woodbury matrix identity).
Let , , and , we have
References
- [1] {barticle}[author] \bauthor\bsnmAnandkumar, \bfnmAnimashree\binitsA., \bauthor\bsnmGe, \bfnmRong\binitsR., \bauthor\bsnmHsu, \bfnmDaniel\binitsD., \bauthor\bsnmKakade, \bfnmSham M\binitsS. M. and \bauthor\bsnmTelgarsky, \bfnmMatus\binitsM. (\byear2014). \btitleTensor decompositions for learning latent variable models. \bjournalJournal of machine learning research \bvolume15 \bpages2773–2832. \endbibitem
- [2] {barticle}[author] \bauthor\bsnmBaik, \bfnmJinho\binitsJ., \bauthor\bsnmArous, \bfnmGérard Ben\binitsG. B. and \bauthor\bsnmPéché, \bfnmSandrine\binitsS. (\byear2005). \btitlePhase transition of the largest eigenvalue for nonnull complex sample covariance matrices. \bjournalThe Annals of Probability \bvolume33 \bpages1643–1697. \endbibitem
- [3] {barticle}[author] \bauthor\bsnmBen Arous, \bfnmGérard\binitsG., \bauthor\bsnmHuang, \bfnmDaniel Zhengyu\binitsD. Z. and \bauthor\bsnmHuang, \bfnmJiaoyang\binitsJ. (\byear2021). \btitleLong Random Matrices and Tensor Unfolding. \bjournalarXiv preprint arXiv:2110.10210. \endbibitem
- [4] {barticle}[author] \bauthor\bsnmBen Arous, \bfnmGerard\binitsG., \bauthor\bsnmMei, \bfnmSong\binitsS., \bauthor\bsnmMontanari, \bfnmAndrea\binitsA. and \bauthor\bsnmNica, \bfnmMihai\binitsM. (\byear2019). \btitleThe landscape of the spiked tensor model. \bjournalCommunications on Pure and Applied Mathematics \bvolume72 \bpages2282–2330. \endbibitem
- [5] {barticle}[author] \bauthor\bsnmBenaych-Georges, \bfnmFlorent\binitsF. and \bauthor\bsnmNadakuditi, \bfnmRaj Rao\binitsR. R. (\byear2011). \btitleThe eigenvalues and eigenvectors of finite, low rank perturbations of large random matrices. \bjournalAdvances in Mathematics \bvolume227 \bpages494–521. \endbibitem
- [6] {barticle}[author] \bauthor\bsnmBiroli, \bfnmGiulio\binitsG., \bauthor\bsnmCammarota, \bfnmChiara\binitsC. and \bauthor\bsnmRicci-Tersenghi, \bfnmFederico\binitsF. (\byear2020). \btitleHow to iron out rough landscapes and get optimal performances: averaged gradient descent and its application to tensor PCA. \bjournalJournal of Physics A: Mathematical and Theoretical \bvolume53 \bpages174003. \endbibitem
- [7] {barticle}[author] \bauthor\bsnmCapitaine, \bfnmMireille\binitsM., \bauthor\bsnmDonati-Martin, \bfnmCatherine\binitsC. and \bauthor\bsnmFéral, \bfnmDelphine\binitsD. (\byear2009). \btitleThe largest eigenvalues of finite rank deformation of large Wigner matrices: convergence and nonuniversality of the fluctuations. \bjournalThe Annals of Probability \bvolume37 \bpages1–47. \endbibitem
- [8] {barticle}[author] \bauthor\bsnmGoulart, \bfnmJosé Henrique de Morais\binitsJ. H. d. M., \bauthor\bsnmCouillet, \bfnmRomain\binitsR. and \bauthor\bsnmComon, \bfnmPierre\binitsP. (\byear2021). \btitleA Random Matrix Perspective on Random Tensors. \bjournalarXiv preprint arXiv:2108.00774. \endbibitem
- [9] {bphdthesis}[author] \bauthor\bsnmHandschy, \bfnmMadeline Curtis\binitsM. C. (\byear2019). \btitlePhase Transition in Random Tensors with Multiple Spikes, \btypePhD thesis, \bpublisherUniversity of Minnesota. \endbibitem
- [10] {barticle}[author] \bauthor\bsnmHitchcock, \bfnmFrank L\binitsF. L. (\byear1927). \btitleThe expression of a tensor or a polyadic as a sum of products. \bjournalJournal of Mathematics and Physics \bvolume6 \bpages164–189. \endbibitem
- [11] {barticle}[author] \bauthor\bsnmHuang, \bfnmJiaoyang\binitsJ., \bauthor\bsnmHuang, \bfnmDaniel Z\binitsD. Z., \bauthor\bsnmYang, \bfnmQing\binitsQ. and \bauthor\bsnmCheng, \bfnmGuang\binitsG. (\byear2020). \btitlePower iteration for tensor pca. \bjournalarXiv preprint arXiv:2012.13669. \endbibitem
- [12] {barticle}[author] \bauthor\bsnmJagannath, \bfnmAukosh\binitsA., \bauthor\bsnmLopatto, \bfnmPatrick\binitsP. and \bauthor\bsnmMiolane, \bfnmLeo\binitsL. (\byear2020). \btitleStatistical thresholds for tensor PCA. \bjournalThe Annals of Applied Probability \bvolume30 \bpages1910–1933. \endbibitem
- [13] {barticle}[author] \bauthor\bsnmLandsberg, \bfnmJoseph M\binitsJ. M. (\byear2012). \btitleTensors: geometry and applications. \bjournalRepresentation theory \bvolume381 \bpages3. \endbibitem
- [14] {binproceedings}[author] \bauthor\bsnmLesieur, \bfnmThibault\binitsT., \bauthor\bsnmMiolane, \bfnmLéo\binitsL., \bauthor\bsnmLelarge, \bfnmMarc\binitsM., \bauthor\bsnmKrzakala, \bfnmFlorent\binitsF. and \bauthor\bsnmZdeborová, \bfnmLenka\binitsL. (\byear2017). \btitleStatistical and computational phase transitions in spiked tensor estimation. In \bbooktitleProc. IEEE International Symposium on Information Theory (ISIT) \bpages511–515. \endbibitem
- [15] {binproceedings}[author] \bauthor\bsnmLim, \bfnmLek-Heng\binitsL.-H. (\byear2005). \btitleSingular values and eigenvalues of tensors: a variational approach. In \bbooktitleProc. IEEE International Workshop on Computational Advances in Multi-Sensor Adaptive Processing \bpages129–132. \endbibitem
- [16] {barticle}[author] \bauthor\bsnmMarčenko, \bfnmVladimir A\binitsV. A. and \bauthor\bsnmPastur, \bfnmLeonid Andreevich\binitsL. A. (\byear1967). \btitleDistribution of eigenvalues for some sets of random matrices. \bjournalMathematics of the USSR-Sbornik \bvolume1 \bpages457. \endbibitem
- [17] {barticle}[author] \bauthor\bsnmMontanari, \bfnmAndrea\binitsA. and \bauthor\bsnmRichard, \bfnmEmile\binitsE. (\byear2014). \btitleA statistical model for tensor PCA. \bjournalarXiv preprint arXiv:1411.1076. \endbibitem
- [18] {bbook}[author] \bauthor\bsnmNocedal, \bfnmJorge\binitsJ. and \bauthor\bsnmWright, \bfnmStephen\binitsS. (\byear2006). \btitleNumerical optimization. \bpublisherSpringer Science & Business Media. \endbibitem
- [19] {barticle}[author] \bauthor\bsnmPéché, \bfnmSandrine\binitsS. (\byear2006). \btitleThe largest eigenvalue of small rank perturbations of Hermitian random matrices. \bjournalProbability Theory and Related Fields \bvolume134 \bpages127–173. \endbibitem
- [20] {binproceedings}[author] \bauthor\bsnmPerry, \bfnmAmelia\binitsA., \bauthor\bsnmWein, \bfnmAlexander S\binitsA. S. and \bauthor\bsnmBandeira, \bfnmAfonso S\binitsA. S. (\byear2020). \btitleStatistical limits of spiked tensor models. In \bbooktitleAnnales de l’Institut Henri Poincaré, Probabilités et Statistiques \bvolume56 \bpages230–264. \bpublisherInstitut Henri Poincaré. \endbibitem
- [21] {barticle}[author] \bauthor\bsnmRabanser, \bfnmStephan\binitsS., \bauthor\bsnmShchur, \bfnmOleksandr\binitsO. and \bauthor\bsnmGünnemann, \bfnmStephan\binitsS. (\byear2017). \btitleIntroduction to tensor decompositions and their applications in machine learning. \bjournalarXiv preprint arXiv:1711.10781. \endbibitem
- [22] {barticle}[author] \bauthor\bsnmStein, \bfnmCharles M\binitsC. M. (\byear1981). \btitleEstimation of the mean of a multivariate normal distribution. \bjournalThe annals of Statistics \bpages1135–1151. \endbibitem
- [23] {barticle}[author] \bauthor\bsnmSun, \bfnmWill Wei\binitsW. W., \bauthor\bsnmHao, \bfnmBotao\binitsB. and \bauthor\bsnmLi, \bfnmLexin\binitsL. (\byear2014). \btitleTensors in Modern Statistical Learning. \bjournalWiley StatsRef: Statistics Reference Online \bpages1–25. \endbibitem
- [24] {bbook}[author] \bauthor\bsnmTao, \bfnmTerence\binitsT. (\byear2012). \btitleTopics in random matrix theory \bvolume132. \bpublisherAmerican Mathematical Soc. \endbibitem
- [25] {barticle}[author] \bauthor\bsnmZare, \bfnmAli\binitsA., \bauthor\bsnmOzdemir, \bfnmAlp\binitsA., \bauthor\bsnmIwen, \bfnmMark A\binitsM. A. and \bauthor\bsnmAviyente, \bfnmSelin\binitsS. (\byear2018). \btitleExtension of PCA to higher order data structures: An introduction to tensors, tensor decompositions, and tensor PCA. \bjournalProceedings of the IEEE \bvolume106 \bpages1341–1358. \endbibitem