1 Introduction

Object recognition is a fundamental problem in computer vision and keeps attracting more and more attention nowadays. Its concepts have been applied in multiple fields, such as manufacturing, surveillance system, optical character recognition, face recognition, etc.

Almost every object recognition algorithm proposed in the literature is based on the computation of certain features of the image, which allow to characterize the object depicted and to discriminate it from others. In particular, since objects can appear at different locations and with different sizes, it is desirable for such features to be invariant by translation, rotation and scale. These invariant features can be global, i.e., computed taking into account the whole image, or local, i.e., computed considering only neighborhoods of key-points in the image.

In this paper we focus on Fourier descriptors, an important class of global invariant features used since the seventies [19, 40] based on algebraic properties of the Fourier transform. In particular, inspired by some neurophysiological facts on the structure of the human primary visual cortex, we extend this theory to define features invariant to translation and rotation and we apply them for invariant object recognition in SVM context. These results are then compared with those obtained with another important class of global invariant features, the moment invariants (see Appendix 1), used e.g., in [10, 34, 35], and with two different local invariants. For more information on object recognition via local features we refer to [2, 12, 25, 27, 28, 30].

Our choice of a global approach is motivated by the better results obtained by these methods in presence of noise, luminance changes and other different alterations, with respect to algorithms based on local invariant features [7]. Indeed, under these conditions, key-points detectors used in the local approach produce key-points that are not relevant for the object recognition.

In the following, we briefly introduce the theory of Fourier descriptors, before discussing the framework used in this paper and our contributions.

1.1 Fourier Descriptors

The basic idea behind Fourier descriptors is that the action of an abelian locally compact group \({\mathbb {G}}\) on functions in \(L^2({\mathbb {G}})\) is much easier to treat at the level of their Fourier transforms. In the specific case of images, \(f,g\in L^2({\mathbb {R}}^2)\), this is expressed by the well-known equivalence for the translation of \(a\in {\mathbb {R}}^2\):

$$\begin{aligned} f(x)= & {} g(x-a) \quad \forall x\in {\mathbb {R}}^2\nonumber \\&\iff \hat{f}(\lambda ) = e^{i\langle a, \lambda \rangle } \hat{g}(\lambda ) \quad \forall \lambda \in {\mathbb {R}}^2, \end{aligned}$$
(1)

where the Fourier transform is definedFootnote 1 by

$$\begin{aligned} \hat{f}(\lambda ) = \int _{{\mathbb {R}}^2} f(x)\,e^{-i\langle \lambda , x\rangle }\,{dx}, \qquad \forall \lambda \in {\mathbb {R}}^2. \end{aligned}$$

In this setting, Fourier descriptors are quantities associated with functions of \(L^2({\mathbb {R}}^2)\) that can be easily computed starting from their Fourier representation and that are invariant under the action of translations. Ideally, a Fourier descriptor should be complete, meaning that for any couple of images \(f,g\in L^2({\mathbb {R}}^2)\) the equality of the Fourier descriptor is equivalent to the equality of f and g up to translations. Indeed, the lack of completeness could yield to problems in applications, notably to false positives in the classification.

However, a result as strong as completeness is usually out of reach and unnecessary for practical applications. In this case, one looks for Fourier descriptors that are at least weakly complete, meaning that they are complete on a sufficiently big subset of \(L^2({\mathbb {R}}^2)\), usually either open and dense or at least residual, i.e., the intersection of countably many open and dense sets. This guarantees that the Fourier descriptor will correctly classify a sufficiently large class of images.

Various Fourier descriptors have been defined in the literature [19, 24, 26, 38, 40]. In this work we are mainly interested in the following two, whose invariance w.r.t. translations can be checked via (1).

  • Power-spectrum: the quantity \({{\mathrm{PS}}}_f(\lambda ) := |\hat{f}(\lambda )|^2\) for \(\lambda \in {\mathbb {R}}^2\), which is the Fourier transform of the auto-correlation function

    $$\begin{aligned} a_f(x) := \int _{{\mathbb {R}}^2} \overline{f(y)} \, f(y+x)\, dy. \end{aligned}$$

    It is easy to show that the power-spectrum is not weakly complete, and indeed it is used in texture synthesis to identify the translation invariant Gaussian distribution of textures [18].

  • Bispectrum: an extension of the power-spectrum, it is the quantity \({{\mathrm{BS}}}_f(\lambda _1,\lambda _2):=\hat{f}(\lambda _1)\, \hat{f}(\lambda _2)\, \overline{\hat{f}(\lambda _1+\lambda _2)}\), or equivalently the Fourier transform of the triple correlation, defined as

    $$\begin{aligned} a_{3,f}(x_1,x_2) := \int _{{\mathbb {R}}^2} \overline{f(y)} \, f(y+x_1)\, f(y+x_2) \,dy. \end{aligned}$$

    These descriptors are complete on compactly supported functions of \(L^2({\mathbb {R}}^2)\) and are well established in statistical signal processing. See e.g., [14], where they are applied to sound texture recognition.

These two Fourier descriptors can be easily generalized to functions on \(L^2({\mathbb {G}})\) of a locally compact abelian group \({\mathbb {G}}\) to obtain invariants under the action of \({\mathbb {G}}\). This can be applied, for example, to 2D shape recognition. However, when working with images, these descriptors are unsatisfying. Indeed, they are invariant only under translations, and so cannot be used to classify images under the action of rotations.

1.2 Framework of the Paper

In this paper, following a line of research started in [38], we present a theoretical framework that allows us to build generalized Fourier descriptors which are invariant w.r.t. (semidiscrete) roto-translations of images. We exploit the following two facts:

  • It is possible to define a natural generalization of the power-spectrum and the bispectrum on non-commutative groups, as it has been done in [24, 38].

  • Contributions of some of the authors to a fairly recent model of the human primary visual cortex V1 [4, 5] have shown that the latter can be modeled as the semidiscrete group of roto-translations \(SE(2,N) = {\mathbb {Z}}_N\rtimes {\mathbb {R}}^2\). In this model, cortical stimuli are functions in \(L^2(SE(2,N))\), w.r.t. the Haar measure of SE(2, N), and images from the visual plane are lifted to cortical stimuli via a natural injective and left-invariant lift operation \(\mathcal{L}:L^2({\mathbb {R}}^2)\rightarrow L^2(SE(2,N))\). Such lift is defined as the wavelet transform w.r.t. a mother wavelet \(\varPsi \), see Sect. 2.

From these facts, a natural pipeline for invariant object recognition is the following:

  1. 1.

    Given an image \(f\in L^2({\mathbb {R}}^2)\) lift it to a cortical stimulus \(\mathcal{L}f\in L^2(SE(2,N))\).

  2. 2.

    Compute the generalized Fourier descriptors of \(\mathcal{L}f\) on the non-commutative group SE(2, N).

  3. 3.

    If the lift of another image \(g\in L^2({\mathbb {R}}^2)\) has the same Fourier descriptors as \(\mathcal{L}f\), deduce that \(\mathcal{L}f \approx \mathcal{L}g\) up to the action of SE(2, N).

  4. 4.

    Thanks to the left-invariance and injectivity of the lift \(\mathcal{L}\), obtain that also \(f\approx g\) up to the action of SE(2, N).

This pipeline was already investigated in [38], where the authors considered a non-left-invariant lift, the cyclic lift. For this lift they then proved a weak completeness result of the generalized bispectrum for images, represented as \(L^2({\mathbb {R}}^2)\) functions with support inside a fixed compact set.

In this paper, we consider the same question for left-invariant lifts, where the situation turns out to be more complicated. In particular, as explained in the following section, to ensure the weak completeness we are led to consider “stronger” invariants than the generalized bispectrum. However, as observed in Remark 4, the actual computation of these stronger invariants on lifted images requires N times less computational time and space w.r.t. the computation of the generalized bispectrum of cyclically lifted images.

1.3 Contributions of the Paper

Let \(K\subset {\mathbb {R}}^2\) be any compact set, representing the size of the images under consideration. According to the pipeline for object recognition introduced above, the weak completeness of the generalized Fourier descriptors on images can be proved in two steps:

  1. 1.

    Prove the completeness of the generalized Fourier descriptors on some residual set \({\mathcal {G}}\subset L^2({\mathbb {Z}}_N\times K)\) of cortical stimuli;

  2. 2.

    Prove that for some residual set \({\mathcal {R}}\subset L^2({\mathbb {R}}^2)\) of images with support in K we have \(\mathcal L({\mathcal {R}})\subset {\mathcal {G}}\).

The first point is addressed in Theorem 1, where is identified an open and dense set \(\mathcal{G}\subset L^2({\mathbb {Z}}_N\times K)\) on which the combination of the generalized power-spectrum and bispectrum holds. This generalizes the result in [38], where the same result was proved for a residual subset of the range of the cyclic lift.

Unfortunately, it turns out that for this set \(\mathcal{G}\) and a left-invariant lift \(\mathcal{L}\) there is no hope of finding a set \(\mathcal{R}\subset L^2({\mathbb {R}}^2)\) satisfying the second point above. We are then led to introduce stronger Fourier descriptors, the rotational power-spectrum and bispectrum, which are invariant only w.r.t. rotations. To solve this problem we preprocess images by centering them at their barycenter, a procedure that is essential also in [38]. Theorem 2 then shows that the resulting invariants are complete for an open and dense set of functions in \(L^2(K)\), for any compact \(K\subset {\mathbb {R}}^2\). The proof of this completeness requires fine technical tools from harmonic analysis and the theory of circulant operators, and for this reason we only present a sketch of it, evidencing the technical difficulties. A complete proof will be presented in a forthcoming paper by the second and last authors.

Finally, in Theorem 3 we show that, under mild assumptions on the mother wavelet \(\varPsi \), to check the equality of all these Fourier descriptors it is sufficient to compute simple quantities computed from the 2D Fourier transform of the image. This allows for an efficient implementation on regular hexagonal grids. After using these descriptors to feed an SVM-based classifier, we compare their performances with those of Hu and Zernike moments, the Fourier–Mellin transform and some well-known local descriptors. To this purpose, we test them on two large databases: the COIL-100Footnote 2 object recognition database, composed of 7200 objects presenting rotation and scale changes [31], and the ORLFootnote 3 face database, on which different human faces are subject to several kinds of variations.

1.4 Structure of the Paper

The remainder of the paper is organized as follows. In Sect. 2, we present the features of a mathematical model of the primary visual cortex V1 that are essential to our approach. In Sect. 3, we introduce some generalities on the Fourier Transform on the semidiscrete group of roto-translation SE(2, N). In Sect. 4, we describe the natural generalization of the power-spectrum and the bispectrum on \({\mathbb {R}}^2\) to SE(2, N). We then prove the weak completeness result (Theorem 1) and show that under the chosen lift operator this does not imply weak completeness for images. Finally, we introduce the rotational power-spectrum and bispectrum and sketch the proof of the corresponding weak completeness result (Theorem 2) for images. We end this section with some results on the practical computation of these descriptors. In Sect. 5, we illustrate some numerical results where these descriptors are compared with those obtained via global descriptors such as Zernike moments, Hu moments, Fourier–Mellin transform, and local ones like the SIFT and HoG descriptors. Finally, we conclude with some practical suggestions in Sect. 6.

2 A Mathematical Model of the Primary Visual Cortex V1

As mentioned in the introduction, the main novelty of our approach is its connection with a fairly recent model of the human primary visual cortex V1 due to Petitot and Citti-Sarti [11, 32] and our recent contributions [35, 33]. The theory of orientation scores introduced in [15, 16] is also strongly connected with this work, in particular for its exploitation of left-invariant lift operators. We also mention [37], where image invariants based on the structure of the roto-translation group SE(2) are introduced for textures. In this section, we present the features of this model that are essential to our approach.

Since it is well-known [23] that neurons in V1 are sensitive not only to positions in the visual field, but also to local orientations and that it is reasonable to assume these orientations to be finite, in [4] V1 has been modeled as the semidiscrete group of roto-translations \(SE(2,N) = {\mathbb {Z}}_N\rtimes {\mathbb {R}}^2\) for some even \(N\in {\mathbb {N}}\). Letting \(R_k\) be the rotation of \(2\pi /k\), the (non-commutative) group operation of SE(2, N) is

$$\begin{aligned} (x,k)(y,r) = (x+R_k y, {k+r}). \end{aligned}$$

Here, we are implicitly identifying \(k+r\) with \(k+r\mod N\).

Visual stimuli \(f\in L^2({\mathbb {R}}^2)\) are assumed to be lifted to activation patterns in \(L^2(SE(2,N))\) by a lift operator \(\mathcal{L}:L^2({\mathbb {R}}^2)\rightarrow L^2(SE(2,N))\). Motivated by neurophysiological evidence, we then assume that

  1. (H)

    The lift operator \(\mathcal{L}\) is linear and is defined as

    $$\begin{aligned} \mathcal{L}f(x,k) := \int _{{\mathbb {R}}^2} f(y) \bar{\varPsi }(R_{-k}(y-x))\, dy, \end{aligned}$$
    (2)

    for a given mother wavelet \(\varPsi \in L^2({\mathbb {R}}^2)\) such that \(\mathcal{L}\) is injective and bounded.

Remark 1

This assumption means that the lift operator under consideration is the wavelet transform w.r.t. \(\varPsi \) (see e.g., [17]). The fact that \(\mathcal{L}\) be injective and bounded is then equivalent to the fact that the mother wavelet \(\varPsi \) is weakly admissible, i.e., is such that the map

$$\begin{aligned} \lambda \in {\mathbb {R}}^2\mapsto \sum _{k\in {\mathbb {Z}}_N} |\hat{\varPsi }(R_{-k}\lambda )|^2 \end{aligned}$$

is strictly positive and essentially bounded.

As a consequence of the above assumption, the lift operation \(\mathcal{L}\) is left-invariant w.r.t. to the action of SE(2, N). Namely,

$$\begin{aligned} \varLambda (x,k)\circ \mathcal{L}= \mathcal{L}\circ \pi (x,k). \end{aligned}$$
(3)

Here \(\varLambda \) and \(\pi \) are the actions of SE(2, N) on \(L^2(SE(2,N))\) and \(L^2({\mathbb {R}}^2)\), respectively. That is,

$$\begin{aligned}&{[}\varLambda (x,k)\varphi {]}(y,r)=\varphi \left( (x,k)^{-1}(y,r)\right) \\&\quad =\varphi (R_{-k}(y-x),{r-k}),\\&{[}\pi (x,k)f{]}(y)=f\left( (x,k)^{-1}y\right) = f(R_{-k}( y-x)) . \end{aligned}$$

Formula (3) can be seen as a semidiscrete version of the shift-twist symmetry [6].

The main observation for our purposes is that (3) means that two images f and \(g\in L^2({\mathbb {R}}^2)\) can be deduced via roto-translation (i.e., \(f = \pi (x,k)g\) for some \((x,k)\in SE(2,N)\)) if and only if their lifts can be deduced via \(\varLambda (x,k)\).

3 Preliminaries on Non-commutative Harmonic Analysis

In this section we introduce some generalities on the (non-commutative) Fourier transform on SE(2, N), an essential tool to define and compute the Fourier descriptors we are interested in. We refer to [1, 20] for a general introduction to the topic.

Since SE(2, N) is a non-commutative unimodular group, the Fourier transform of \(\varphi \in L^2(SE(2,N))\) is an operator associating to each (continuous) irreducible unitary representations \(T^\lambda \) of SE(2, N) some Hilbert–Schmidt operator on the Hilbert space where \(T^\lambda \) acts. Here, \(\lambda \) is an index taking values in the dual object of SE(2, N), which is denoted by \(\widehat{SE(2,N)}\) and is the set of equivalence classes of irreducible unitary representations.

The set of irreducible representations of a semi-direct product group can be obtained via Mackey’s machinery (see e.g., [1, Ch. 17.1, Theorems 4 and 5]). Accordingly, \(\widehat{SE(2,N)}\) is parametrized by the orbits of the (contragredient) action of rotations \(\{R_k\}_{k\in {\mathbb {Z}}_N}\) on \({\mathbb {R}}^2\), i.e., by the slice \(\mathcal{S}\subset {\mathbb {R}}^2\) which in polar coordinates is \((0,+\infty )\times [0,2\pi /N)\). Additionally, corresponding to the origin, there are the characters of \({\mathbb {Z}}_N\). Namely, to each \(\lambda \in \mathcal{S}\) corresponds the representation \(T^\lambda \) acting on \({\mathbb {C}}^N\) via

$$\begin{aligned} T^\lambda (x,k)v= & {} \text {diag}_h(e^{i\langle \lambda , R_h x \rangle })\circ S^k v \nonumber \\= & {} \left( e^{i\langle \lambda , R_h x\rangle }v_{h+k} \right) _{h=0}^{N-1}, \end{aligned}$$
(4)

where we denoted by \({{\mathrm{diag}}}_h v_h\) the diagonal matrix of diagonal \(v\in {\mathbb {C}}^N\) and by S the shift operator \((Sv)_j = v_{j+1}\), so that \((S^kv)_j = v_{j+k}\). On the other hand, to each \(k\in {\mathbb {Z}}_N\) corresponds the representation on \({\mathbb {C}}\) given by \(z\mapsto e^{i \frac{2\pi k}{N}} z\). Since it is possible to show that to invert the Fourier transform it is enough to consider only the representations parametrized by \(\mathcal{S}\), we will henceforth ignore the \({\mathbb {Z}}_N\) part of the dual.

Finally, the matrix-valued Fourier coefficient of a function \(\varphi \in L^2(SE(2,N))\cap L^1(SE(2,N))\) for \(\lambda \in \widehat{SE(2,N)}\) is

$$\begin{aligned} \hat{\varphi }(T^\lambda ) = \int _{SE(2,N)} \varphi (a)\,T^\lambda (a^{-1})\,da, \end{aligned}$$
(5)

where da is the Haar measureFootnote 4 of SE(2, N). This is essentially the same formula for the Fourier transform on \({\mathbb {R}}^2\), which is a scalar and is obtained using the representations \(\lambda (x) = e^{i \langle x, \lambda \rangle }\) acting on \({\mathbb {C}}\).

Straightforward computations yield

$$\begin{aligned} \hat{\varphi }(T^\lambda )_{i,j} = \mathcal{F}(\varphi (\cdot ,i-j))(R_{-j}\lambda ), \end{aligned}$$
(6)

where we let \(\mathcal{F}\) denote the Fourier transform on \({\mathbb {R}}^2\).

As usual, the definition of the Fourier transform can be extended to the whole \(L^2(SE(2,N))\) by density arguments. Then, there exists a unique measure on \(\widehat{SE(2,N)}\), the Plancherel measure, supported on \(\mathcal{S}\) where it coincides with the restriction of the Lebesgue measure of \({\mathbb {R}}^2\), such that the Fourier transform is an isometry between \(L^2(SE(2,N))\) and \(L^2(\widehat{SE(2,N)})\). In particular, the following inversion formula holds

$$\begin{aligned} \varphi (x,k) = \int _{\mathcal{S}}\text {Tr}\left( \hat{\varphi }(T^\lambda )\circ T^\lambda (x,k)\right) \,d\lambda . \end{aligned}$$

The fundamental property of the non-commutative Fourier transform, generalizing (1), is that for all \(\varphi ,\eta \in L^2(SE(2,N))\) and \(a\in SE(2,N)\) it holds

$$\begin{aligned} \varphi (x,k)= & {} [\varLambda (a)\eta ](x,k) \quad \forall (x,k)\in SE(2,N) \iff \nonumber \\&\hat{\varphi }(T) = \hat{\eta }(T)\circ T^{-1}(a) \quad \forall T\in \widehat{SE(2,N)}. \end{aligned}$$
(7)

Namely, \(\varphi ,\eta \) can be deduced via the action of SE(2, N) if and only if their Fourier transforms at a representation T can be deduced via multiplication by T(a).

Remark 2

The fact that the Fourier transform in (5) be matrix-valued is a direct consequence of SE(2, N) being a Moore group, that is, that all the \(T^\lambda \) act on finite-dimensional spaces. This is not true for the roto-translation group SE(2). As a consequence, the Fourier transform on SE(2), takes values not in the finite dimensional space of complex \(N\times N\) matrices, but in the infinite dimensional space of operators over \(L^2({\mathbb {S}}^1)\). This is indeed the main theoretical advantage of considering SE(2, N).

3.1 Decomposition of Tensor Product Representations

Proofs of Sect. 4, will use a well-known fact on tensor product representations: the Induction–Reduction Theorem (see [1]). This theorem allows to decompose the tensor products of representations \(T^{\lambda _1}\otimes T^{\lambda _2}\), acting on \({\mathbb {C}}^N\otimes {\mathbb {C}}^N\cong {\mathbb {C}}^{N\times N}\), to an equivalent representation acting on \(\bigoplus _{k\in {\mathbb {Z}}_N} {\mathbb {C}}^N\), which is a block-diagonal operator whose block elements are of the form \(T^{\lambda _1+R_k\lambda _2}\). Moreover, the linear transformation realizing the equivalence is explicit.

To avoid confusion, we will henceforth denote components of vectors \(v\in {\mathbb {C}}^N\) as \(v(0),\ldots , v(N-1)\), elements of \(\bigoplus _{k\in {\mathbb {Z}}_N} {\mathbb {C}}^N\) as \((w_k)_{k\in {\mathbb {Z}}_N}\) where \(w_k\in {\mathbb {C}}^N\), and the components of vectors \(\mathbf v \in {\mathbb {C}}^N\otimes {\mathbb {C}}^N\) as \(\mathbf v (k,h)\) for \(k,h=0,\ldots , N-1\). We also remark that linear operators \(\mathcal{B}\) on \(\bigoplus _{k\in {\mathbb {Z}}_N}{\mathbb {C}}^N\) can be decomposed as \(\mathcal{B}= (\mathcal{B}^{k,h})_{k,h\in {\mathbb {Z}}_N}\), where each \(\mathcal{B}^{k,h}\) is an \(N\times N\) complex matrix. Namely, we have

$$\begin{aligned} \mathcal{B}(w_k)_{k\in {\mathbb {Z}}_N} = \left( \sum _{h=0}^{N-1}\mathcal{B}^{k,h}w_h \right) _{k\in {\mathbb {Z}}_N}. \end{aligned}$$
(8)

Then, exploiting the commutation of the Fourier transform with equivalences of representation, the Induction–Reduction Theorem implies that for every \(\varphi \in L^2(SE(2,N))\) and any \(\lambda _1,\lambda _2\in \mathcal{S}\) it holds

$$\begin{aligned} A\circ \hat{\varphi }(T^{\lambda _1}\otimes T^{\lambda _2}) \circ A^{-1} = \bigoplus _{k\in {\mathbb {Z}}_N} \hat{\varphi } (T^{\lambda _1+R_k\lambda _2}). \end{aligned}$$
(9)

Here, \(A:{\mathbb {C}}^N\otimes {\mathbb {C}}^N \rightarrow \bigoplus _{k\in {\mathbb {Z}}_N} {\mathbb {C}}^N\) is given by

$$\begin{aligned} (A \mathbf v )_k(h) = (A_k \mathbf v )(h) = \mathbf v (h,h-k), \quad \forall \mathbf v \in {\mathbb {C}}^N\otimes {\mathbb {C}}^N. \end{aligned}$$
(10)

4 Fourier Descriptors on SE(2, N)

In the following sections, we introduce and study the Fourier descriptors on the group SE(2, N). As already mentioned, proving a general completeness result is essentially hopeless, and we will content ourselves to prove the weak completeness.

Let \(K\subset {\mathbb {R}}^2\) be a compact set. In the following, we will be mainly concerned with functions that are compactly supported either in K or in \(K\times {\mathbb {Z}}_N\subset SE(2,N)\).

4.1 Generalized Fourier Descriptors

Following [38], the power-spectrum and the bispectrum on \({\mathbb {R}}^2\) can be generalized to SE(2, N) as follows.

Definition 1

The generalized power-spectrum and bispectrum of \(\varphi \in L^2(SE(2,N))\) are the collections of matrices for any \(\lambda ,\lambda _1,\lambda _2\in \mathcal{S}\),

$$\begin{aligned}&{{\mathrm{PS}}}_\varphi (\lambda ) := \widehat{\varphi }(T^\lambda ) \circ \widehat{\varphi }(T^\lambda )^* \\&{{\mathrm{BS}}}_\varphi (\lambda _1,\lambda _2) := \widehat{\varphi }(T^{\lambda _1})\otimes \widehat{\varphi }(T^{\lambda _2}) \circ \widehat{\varphi }(T^{\lambda _1}\otimes T^{\lambda _2})^*. \end{aligned}$$

The next result generalizes, with a simplified proof, the result presented in [38]. Let us mention that this result is indeed true in a more general setting, as it will be shown in a forthcoming paper by Prandi and Gauthier.

Theorem 1

Let \(K\subset {\mathbb {R}}^2\) be a compact. The generalized power-spectrum and bispectrum are weakly complete on \(L^2({\mathbb {Z}}_N\times K)\). In particular, they discriminate on the open and dense set \(\mathcal{G}\subset L^2({\mathbb {Z}}_N\times K)\) of functions \(\varphi \) supported in \({\mathbb {Z}}_N\times K\) and whose Fourier transform \(\hat{\varphi }(T^\lambda )\) is invertible for an open and dense set of \(\lambda \)s. That is, \(\varphi _1,\varphi _2\in \mathcal{G}\) are such that \({{\mathrm{PS}}}_{\varphi _1}={{\mathrm{PS}}}_{\varphi _2}\) and \({{\mathrm{BS}}}_{\varphi _1}={{\mathrm{BS}}}_{\varphi _2}\) if and only if \(\varphi _1 = \varLambda (x,k)\varphi _2\) for some \((x,k)\in SE(2,N)\).

Proof

The fact that \(\mathcal{G}\) is open and dense is proved in Lemma 1 in Appendix 2. Let \(\varphi , \eta \in \mathcal{G}\) be such that \(PS_{\varphi _1}=PS_{\varphi _2}\) and \({{\mathrm{BS}}}_\varphi ={{\mathrm{BS}}}_\eta \). The equality of the generalized bispectrum implies that the set of \(\lambda \)s where \(\hat{\varphi }(T^\lambda )\) and \(\hat{\eta }(T^\lambda )\) fail to be invertible is the same. We will denote it by I and let

$$\begin{aligned} U(T^\lambda ):=\hat{\varphi }(T^\lambda )^{-1} \hat{\eta }(T^\lambda ) \qquad \forall \lambda \in I. \end{aligned}$$

In order to complete the proof of the statement, we will prove that \(U(T^\lambda )\) can be defined for all \(\lambda \)s in \({\mathbb {R}}^2\) and, moreover, that \(U(T^\lambda ) = T^\lambda (a)\) for some \(a\in SE(2,N)\). Indeed, by (7) this will readily imply that \(\varphi = \varLambda (a)\eta \) as announced.

We claim that \(U(T^\lambda )\) is unitary for all \(\lambda \in I\). Indeed, by the equality of the generalized power-spectrum we have

$$\begin{aligned} U(T^\lambda )^* U(T^\lambda ) = \eta (T^\lambda )^* {{\mathrm{PS}}}_\varphi (\lambda ) \eta (T^\lambda ) = \mathrm{I}. \end{aligned}$$

Observe that the equality of the generalized bispectrum and the definition of U, imply that for all \(\lambda _1,\lambda _2\in I\) it holds

$$\begin{aligned} {{\mathrm{BS}}}_\varphi (\lambda _1,\lambda _2)= & {} \hat{\varphi }(T^{\lambda _1}) \otimes \hat{\varphi }(T^{\lambda _2})\circ U(T^{\lambda _1})\\&\otimes U(T^{\lambda _2}) \circ \hat{\eta }(T^{\lambda _1}\otimes T^{\lambda _2})^*. \end{aligned}$$

By the invertibility of \(\hat{\varphi }(T^{\lambda _1})\otimes \hat{\varphi }(T^{\lambda _2})\) and the unitarity of U, this yields

$$\begin{aligned} \hat{\varphi }(T^{\lambda _1}\otimes T^{\lambda _2})\circ U(T^{\lambda _1})\otimes U(T^{\lambda _2}) = \hat{\eta }(T^{\lambda _1}\otimes T^{\lambda _2}). \end{aligned}$$
(11)

The announced result is then a consequence of the following three facts, which are proved in Appendix 2.

  1. 1.

    Lemma 2: The function \(\lambda \mapsto U(T^\lambda )\) is continuous on I.

  2. 2.

    Lemma 3: The function \(\lambda \mapsto U(T^\lambda )\) can be extended to a continuous function on \({\mathbb {R}}^2\) for which (11) is still true.

  3. 3.

    Lemma 4: There exists \(a\in SE(2,N)\) such that \(U(T^\lambda ) = T^\lambda (a)\). \(\square \)

An immediate corollary is the following.

Corollary 1

Let \(\tilde{\mathcal{L}}: L^2({\mathbb {R}}^2)\rightarrow L^2(SE(2,N)) \) be an injective lift operator (not necessarily satisfying (2)). Assume that there exists a residual set \(\mathcal{R}\subset L^2({\mathbb {R}}^2)\) such that \(\tilde{\mathcal{L}}(\mathcal{R})\cap \mathcal{G}\) is residual. Then, the generalized power-spectrum and bispectrum are weakly complete on \(L^2({\mathbb {R}}^2)\). Namely, for any \(f,g\in \mathcal{R}\) it holds that \({{\mathrm{BS}}}_{\tilde{\mathcal{L}} f} = {{\mathrm{BS}}}_{\tilde{\mathcal{L}} g}\) if and only if \(f=\pi (x,k)g\) for some \((x,k)\in SE(2,N)\).

Remark 3

In [38], the authors applied their version of Theorem 1 to a non-left-invariant lift \(\tilde{\mathcal{L}}\), called cyclic lift. Indeed, for this cyclic lift, when N is odd, it is possible to prove that, for any compact \(K\subset {\mathbb {R}}^2\), there exists a residual set \(\mathcal{R}\subset L^2(K)\) satisfying the assumptions of Corollary 1.

Unfortunately, Corollary 1 can never be applied to lifts of the form (2). In fact, as proved in Appendix 3, letting \(\omega _f(\lambda ):=(\hat{f}(R_{-k}\lambda ))_{k=0}^{N-1}\in {\mathbb {C}}^N\), we have that

$$\begin{aligned} \widehat{\mathcal{L}f}(T^\lambda ) = \omega _\varPsi (\lambda )^* \otimes \omega _f(\lambda )^*, \end{aligned}$$
(12)

where for \(v,w\in {\mathbb {C}}^N\), we let \(v^* = (\overline{v_k})_k\) and \((v\otimes w)_{k,h} = v_k\,\overline{w_h}\), so that \((v\otimes w)u = \langle w, u\rangle \, v\) for all \(u\in {\mathbb {C}}^N\). This immediately implies that \(\text {rank } \widehat{\mathcal{L}f}(T^\lambda )\le 1\) and hence that \(\text {range }\mathcal{L}\cap \mathcal{G}= \emptyset \) whenever \(N>1\).

4.2 Rotational Fourier Descriptors

To bypass the difficulty posed by the non-invertibility of the Fourier transform for lifted functions, we are led to consider the following stronger descriptors.

Definition 2

The rotational power-spectrum and bispectrum of \(\varphi \in L^2(SE(2,N))\) are the collections of matrices, for any \(\lambda ,\lambda _1,\lambda _2\in \mathcal{S}\) and \(h\in {\mathbb {Z}}_N\),

$$\begin{aligned}&{{\mathrm{RPS}}}\varphi (\lambda ,h) := \widehat{\varphi }(T^{R_h\lambda }) \circ \widehat{\varphi }(T^\lambda )^* \\&{{\mathrm{RBS}}}_\varphi (\lambda _1,\lambda _2,h) := \widehat{\varphi }(T^{R_h{\lambda _1}})\otimes \widehat{\varphi }(T^{\lambda _2}) \circ \widehat{\varphi }(T^{\lambda _1}\otimes T^{\lambda _2})^*. \end{aligned}$$

As already mentioned in the introduction, the rotational descriptors are invariant only under the action of \({\mathbb {Z}}_N\subset SE(2,N)\) but not under translations. To avoid this problem, let us fix a compact \(K\subset {\mathbb {R}}^2\) and consider the set \(\mathcal{A}\subset L^2({\mathbb {R}}^2)\) of functions compactly supported in K, with non-zero averageFootnote 5. Observe that this is an open and dense subset of \(L^2(K)\). We can then define the barycenter \(c_f\in {\mathbb {R}}^2\) of \(f\in \mathcal{A}\) as

$$\begin{aligned} c_f=\frac{1}{\text {avg } f} \left( \int _{{\mathbb {R}}^2}x_1 f(x)\,dx, \int _{{\mathbb {R}}^2}x_2 f(x)\,dx\right) , \end{aligned}$$

and the centering operator \(\varPhi :\mathcal{A}\rightarrow \mathcal{A}\) as

$$\begin{aligned} \varPhi f(x) :=f(x-c_f). \end{aligned}$$
(13)

Then, considering the centered lift \(\mathcal{L}_c = \mathcal{L}\circ \varPhi \), we have that \(\mathcal{L}_c f = \mathcal{L}_c g\) if and only if g is a translate of f. In particular,

$$\begin{aligned} \mathcal{L}_c f= & {} \varLambda (0,k)\mathcal{L}_c g \\&\iff f = \pi (x,k) g \quad \text {for some }x\in {\mathbb {R}}^2. \end{aligned}$$

Let us consider the following set of functions.

Definition 3

Let \(\mathcal{R}\subset L^2({\mathbb {R}}^2)\) be the set of real-valued functions f supported in K, such that \(\hat{f}(\lambda )\ne 0\) for a.e. \(\lambda \in {\mathbb {R}}^2\) and the family \(\varOmega _f = \{S^k \omega _f(\lambda )\}_{k=0}^{N-1}\) is a basis for \({\mathbb {C}}^N\), if N is odd, or, if N is even, for

$$\begin{aligned} \mathcal{X}= \{ v\in {\mathbb {C}}^N \mid v(h) = \overline{v(h+N/2)} \quad \forall h\in {\mathbb {Z}}_N \}. \end{aligned}$$

The dependence of this definition on the parity of N comes from the well-known fact that \(\hat{f}(\lambda ) = \overline{\hat{f}(-\lambda )}\). Indeed, for N even, this implies that \(S^k\omega _f(\lambda )\in \mathcal{X}\) for any \(k\in {\mathbb {Z}}_N\). As such, there is no hope for the family \(\varOmega _f\) to generate the whole \({\mathbb {C}}^N\).

Finally, we have the following theorem.

Theorem 2

For any compact \(K\subset {\mathbb {R}}^2\), if the mother wavelet \(\varPsi \in \mathcal{R}\), the rotational power-spectrum and bispectrum are weakly complete on \(L^2(K)\cap \mathcal{A}\). Namely, the set \(\mathcal{R}\) is open and dense in \(L^2(K)\) and for any \(f,g\in \mathcal{R}\cap \mathcal{A}\) it holds that \({{\mathrm{RPS}}}_{\mathcal{L}_c f}={{\mathrm{RPS}}}_{\mathcal{L}_c g}\) and \({{\mathrm{RBS}}}_{\mathcal{L}_c f}={{\mathrm{RBS}}}_{\mathcal{L}_c g}\) if and only if \(f = \pi (x,k)g\) for some \((x,k)\in SE(2,N)\).

Here, we content ourselves to present only a sketch of the proof of this result for the case N odd. The parity of N does not introduce essential problems, up to exploit the fact that \(\text {range } \widehat{Lf}(T^\lambda )\subset \mathcal{X}\) for all \(f\in \mathcal{R}\) and \(\lambda \in \mathcal{S}\) and that the equivalence A of the Induction–Reduction Theorem quotients nicely to an equivalence between \(\mathcal{X}\times \mathcal{X}\) and \(\bigoplus _{k\in {\mathbb {Z}}_N}\mathcal{X}\). However, in order to prove the key technical point (14) we need a much finer study of the properties of circulant operators, which is outside the scope of this work and we defer to a forthcoming paper by Prandi and Gauthier.

Proof

(Sketch in the case N odd) The fact that \(\mathcal{R}\) is open and dense in \(L^2(K)\) follows from the same arguments in Lemma 1.

Let \({{\mathrm{Circ}}}v\) be the circulant matrix associated with v, that is, \({{\mathrm{Circ}}}v = [v,Sv,\ldots ,S^{N-1}v]\). Then the condition on \(\varOmega _f\) for \(f\in \mathcal{R}\) is equivalent to the invertibility of \({{\mathrm{Circ}}}\omega _f(\lambda )\) for an open and dense set of \(\lambda \)s. By the properties of the Fourier transform on \({\mathbb {R}}^2\) w.r.t. translations it follows that

$$\begin{aligned} \omega _{\varPhi f}(\lambda ) = {{\mathrm{diag}}}_k\left( e^{-i\langle \lambda , R_k c_f \rangle } \right) \omega _f(\lambda ). \end{aligned}$$

This entails that \({{\mathrm{Circ}}}\omega _f(\lambda )\) is invertible if and only if \({{\mathrm{Circ}}}\omega _{\varPhi f}(\lambda )\) is. Hence, the statement is equivalent to the fact that for any couple \(f,g\in \mathcal{R}\) we have \({{\mathrm{RBS}}}_{\mathcal{L}f}={{\mathrm{RBS}}}_{\mathcal{L}g}\) if and only if \(f = R_k g\) for some \(k\in {\mathbb {Z}}_N\).

The proof is similar to the one of Theorem 1, but with additional technical difficulties. Let I be the set where \({{\mathrm{Circ}}}\omega _f(\lambda )\) and \({{\mathrm{Circ}}}\omega _g(\lambda )\) are invertible. By assumption I is open and dense. To overcome the non-invertibility of \(\widehat{\mathcal{L}f}\) in the definition the candidate intertwining representation U, we exploit the invertibility of the circulant matrices \({{\mathrm{Circ}}}\omega _f(\lambda )\) and \({{\mathrm{Circ}}}\omega _g(\lambda )\) on an open and dense set. Namely, for any \(\lambda \in I\) we let

$$\begin{aligned} U(T^\lambda )^* := {{\mathrm{Circ}}}\omega _g(\lambda )\,\left( {{\mathrm{Circ}}}\omega _f(\lambda ) \right) ^{-1}. \end{aligned}$$

By definition, \(U(T^\lambda )\) is circulant and \(U(T^\lambda )^* S^k\omega _f(\lambda ) = S^k \omega _g(\lambda )\) for any \(k\in {\mathbb {Z}}_N\). Moreover, by (12), this is equivalent to

$$\begin{aligned} \widehat{\mathcal{L}f}(T^{R_k\lambda }) \, U(T^\lambda ) = \widehat{\mathcal{L}g}(T^{R_k\lambda }), \qquad \forall k\in {\mathbb {Z}}_N. \end{aligned}$$

In particular, \(\lambda \mapsto U(T^{\lambda })\) is constant on orbits \(\{R_k\lambda \}_{k\in {\mathbb {Z}}_N}\). Finally, \(U(T^\lambda )\) is unitary as a consequence, e.g., of Theorem 3.

The main difficulty in the proof is now to derive the equivalent of identity (11), that is, that for an open and dense set of couples \((\lambda _1,\lambda _2)\) we have

$$\begin{aligned}&\widehat{\mathcal{L}f}(T^{R_k\lambda _1}\otimes T^{R_k\lambda _2}) \, U(T^{\lambda _1})\otimes U(T^{\lambda _2}) \nonumber \\&\quad = \widehat{\mathcal{L}g}(T^{R_k\lambda _1}\otimes T^{R_k\lambda _2}), \quad \forall k\in {\mathbb {Z}}_N. \end{aligned}$$
(14)

As already mentioned, the proof of this identity requires a deep use of properties of circulant operators, which is outside the scope of this paper. We thus defer it to a forthcoming paper.

Once (14) is known, the statement follows applying the same arguments as those in Theorem 1. Namely,

  1. 1.

    The function \(\lambda \mapsto U(T^\lambda )\) is continuous on I. This can be done via the same arguments as in Lemma 2.

  2. 2.

    The function \(\lambda \mapsto U(T^\lambda )\) can be extended to a continuous function on \(\mathcal{S}\) still satisfying (14). This can be done exactly as in Lemma 3.

  3. 3.

    There exists \(k\in {\mathbb {Z}}_N\) such that \(U(T^\lambda ) = T^\lambda (0,k)\). This is proved following Lemma 4. Indeed, the fact that now \(\lambda \mapsto U(T^\lambda )\) is constant on the orbits \(\left\{ R_k \lambda \right\} _{k\in {\mathbb {Z}}_N}\) implies that the \(\varphi _k\)s obtained there have to be independent of k. Since \(\varphi _k(\lambda )=e^{i\langle R_k x_0, \lambda \rangle }\) for some \(x_0\in {\mathbb {R}}^2\), this implies that \(x_0 =0\) and hence \(\varphi _k\equiv 0\). Obviously, this proves that \(U(T^\lambda )=S^k = T^\lambda (0,k)\), for some \(k\in {\mathbb {Z}}_N\). \(\square \)

4.3 Practical Computation of the Fourier Descriptors

Here, we present some explicit formulae for the computation of the Fourier descriptors presented in this section.

In the following, we show that, under some assumptions on the mother wavelet \(\varPsi \), the concrete computation of the generalized power-spectrum and bispectrum and of their rotational counterparts, depend only on the 2D Fourier transform of f.

Theorem 3

Assume that the mother wavelet \(\varPsi \in \mathcal{R}\). Then:

  • For any \(f\in \mathcal{R}\), the generalized power-spectrum and bispectrum of \(\mathcal{L}f\) are, respectively, determined by the quantities, for a.e. \(\lambda ,\lambda _1,\lambda _2\in \mathcal{S}\),

    $$\begin{aligned}&I_1^{\lambda }(f)= \Vert \omega _f(\lambda )\Vert ^2 = \sum _{k=0}^{N-1} |\hat{f}(R_{-k}\lambda )|^2 \\&I_1^{\lambda _1,\lambda _2}(f) = \langle \omega _f(\lambda _1)\odot \omega _f(\lambda _2), \omega _f(\lambda _1+\lambda _2)\rangle \\&\quad =\sum _{k=0}^{N-1} \hat{f}(R_{-k}\lambda _1)\hat{f}(R_{-k}\lambda _2) \overline{\hat{f}(R_{-k}(\lambda _1+\lambda _2))}. \end{aligned}$$
  • For any \(f\in \mathcal{A}\cap \mathcal{R}\), the rotational power-spectrum and bispectrum of \(\mathcal{L}_c f\) are, respectively, determined by the quantities, for a.e. \(\lambda ,\lambda _1,\lambda _2\in \mathcal{S}\) and \(h\in {\mathbb {Z}}_N\),

    $$\begin{aligned}&I_2^{\lambda ,h}(f)=\,\langle \omega _{\Phi f}(R_h\lambda ), \omega _{\Phi f}(\lambda )\rangle \\&\quad =\sum _{k=0}^{N-1} \overline{\hat{f}(R_{-k+h}\lambda )} \hat{f}(R_{-k}\lambda ), \\&I_2^{\lambda _1,\lambda _2,h}(f)=\langle \omega _{\Phi f}(R_h\lambda _1)\odot \omega _{\Phi f}(\lambda _2), \omega _{\Phi f}(\lambda _1+\lambda _2)\rangle \\&\quad =\sum _{k=0}^{N-1} \hat{f}(R_{-k+h}\lambda _1)\hat{f}(R_{-k}\lambda _2) \overline{\hat{f}(R_{-k}(\lambda _1+\lambda _2))}. \end{aligned}$$

    Here, \(\varPhi :\mathcal{A}\rightarrow \mathcal{A}\) is the centering operator defined in (13).

Remark 4

Theorem 3 shows in particular that the result of Theorem 2 is indeed stronger than the completeness result for the generalized bispectrum of the cyclic lift obtained in [38]. Indeed, in that work is proved that the latter (for odd N) is determined exactly by the quantities, for a.e. \(\lambda _1,\lambda _2\in \mathcal{S}\) and \(h,k\in {\mathbb {Z}}_N\),

$$\begin{aligned}&\tilde{I}_2^{\lambda _1,\lambda _2, k, h} = \langle \omega _{\phi f}(R_h\lambda _1)\odot \omega _{\phi f}(R_k\lambda _2), \\&\quad \omega _{\phi f}(\lambda _1+R_{h+k}\lambda _2)\rangle . \end{aligned}$$

In particular, for each \(\lambda _1,\lambda _2\in \mathcal{S}\) one has to compute N times more quantities than those for the rotational bispectrum.

As a corollary of Theorem 3 we show that, in order to compare the power-spectra and bispectra, it is usually enough to compare only the latter.

Corollary 2

Let \(\varPsi \in \mathcal{R}\) and \(f,g\in \mathcal{R}\cap \mathcal{A}\). Then, if \(\mathcal{L}f\) and \(\mathcal{L}g\) have the same generalized (resp. rotational) bispectrum, they have also the same generalized (resp. rotational) power-spectrum.

Proof

We only prove the result for the rotational descriptors. In order to prove the one for the generalized descriptors, it will be enough to fix \(h=0\) in the following. By Theorem 3 it is enough to show that whenever \(I^{\lambda _1,\lambda _2,h}_2(f)= I^{\lambda _1,\lambda _2,h}_h(g)\) for a.e. \(\lambda _1,\lambda _2\in \mathcal{S}\) and any \(h\in {\mathbb {Z}}_N\), then \(I^{\lambda ,h}_2(f) = I^{\lambda ,h}_1(g)\) for a.e. \(\lambda \in \mathcal{S}\) and any \(h\in {\mathbb {Z}}_N\). We start by observing that by the Paley–Wiener Theorem all these quantities are analytic, since f and g are compactly supported. Moreover,

$$\begin{aligned} \lim _{\lambda _1,\lambda _2\downarrow 0} I_2^{\lambda _1,\lambda _2,h}(f) = N \, \hat{f}(0) |\hat{f}(0)|^2 = N \, {{\mathrm{avg}}}(f)^3, \end{aligned}$$

and the same is true for g. Thus, \({{\mathrm{avg}}}(f) = {{\mathrm{avg}}}(g)\). Finally, the result follows observing that

$$\begin{aligned} \lim _{\lambda _2\downarrow 0} I_2^{\lambda _1,\lambda _2,h}(f) = {{\mathrm{avg}}}(f)\,I^{\lambda _1,h}_2(f). \end{aligned}$$

\(\square \)

5 Experimental Results

The goal of this section is to evaluate the performance of the invariant Fourier descriptors defined in the previous section on a large image database for object recognition. In addition to the generalized power-spectrum (PS) and bispectrum (BS) and the rotational power-spectrum (RPS) and bispectrum (RBS), we also consider the combination of the RPS and BS descriptors. Indeed, combining these two descriptors seems to be a good compromise between the theoretical result of completeness given by Theorem 2, which only holds for the RBS, and computational demands, as the results on the COIL-100 database will show.

After showing how to efficiently compute these descriptors and presenting the image data-set, we analyze some experimental results. In order to estimate the features capabilities, we use a support vector machine (SVM) [39] as supervised classification method. The recognition performances of the different descriptors regarding invariance to rotation, discrimination capability and robustness against noise are compared.

Fig. 1
figure 1

Steps of computing the invariant descriptors. (S1) computation of the shifted FFT of the image f, (S2) generation of the hexagonal grid, (S3) extraction of different hexagons, (S4) evaluation of the FFT of f on each extracted hexagon, (S5) generation of the vector \(\omega _f (\lambda )\) and (S6) computation of the four invariants

5.1 Implementation

As proved in Theorem 3, the equality of the Fourier descriptors we introduced does not depend on the choice of the mother wavelet \(\varPsi \). Accordingly, in our implementation we only computed the quantities introduced in Theorem 3, whose complexity is reduced to the efficient computation of the vector \(\omega _f (\lambda )\), for a given \(\lambda \in \mathcal{S}\). We recall that this vector is obtained by evaluating the Fourier transform of f on the orbit of \(\lambda \) under the action of discrete rotations \(R_{-k}\) for \(k \in {\mathbb {Z}}_N\).

Let us remark that, although in our implementation we chose this approach, in principle fixing a specific mother wavelet could be useful to appropriately weight descriptors depending on the associated frequencies. Indeed, preliminary tests with a Gabor mother wavelet (which can be easily shown to be in \(\mathcal{R}\)) showed slightly better results at a bigger computational cost.

For the implementation we chose to consider \(N = 6\) and to work with images composed of hexagonal pixels. There are two reasons for this choice:

  • It is well-known that retinal cells are distributed in a hexagonal grid, and thus it is reasonable to assume that cortical activations reflect this fact.

  • Hexagonal grids are invariant under the action of \({\mathbb {Z}}_6\) and discretized translations, which is the most we can get in the line of the invariance w.r.t. SE(2, 6). Indeed, apart from the hexagonal lattice, the only other lattices on \({\mathbb {R}}^2\) which are invariant by some \({\mathbb {Z}}_N\) and appropriate discrete translations are obtained with \(N=2,3,4\).

The different steps of computation of the descriptorsFootnote 6 are described in Fig. 1 and given as follows:

  1. 1.

    The input image is converted to grayscale mode, the Fourier transform is computed via FFT, and the zero-frequency component is shifted to the center of the spectrum (Fig. 1, S1).

  2. 2.

    For cost computational reasons and since we are dealing with natural images, for which the relevant frequencies are the low ones, we extract a grid of \(16\times 16\) pixels around the origin (Fig. 1, S2).

  3. 3.

    The invariants of Theorem 3 are computed from the shifted Fourier transform values, on all frequencies in an hexagonal grid inside this \(16\times 16\) pixels square. A bilinear interpolation is applied to obtain the correct values of \(\omega _f (\lambda )\) (Fig. 1, S3, S4, S5, S6). The final dimension of the feature vector is given in Table 1.

Table 1 Dimension of the feature vectors for the Fourier descriptors under consideration

5.2 Test Protocol

We use the Fourier descriptors to feed an SVM classifier, via the MATLAB Statistics and Machine Learning Toolbox, applying it on a database of 7200 objects extracted from the Columbia Object Image Library (COIL-100) and a database of 400 faces extracted from ORL face database. Finally, we compare the results obtained with those obtained using traditional descriptors.

The result of the training step consists of the set of support vectors determined by the SVM-based method. During the decision step, the classifier computes the Fourier descriptors and the model determined during the training step is used to perform the SVM decision. The output is the image class.

For COIL-100 database, two cases are studied: a case without noise and another with noise. In the first one, tests have been performed using 75% of the COIL-100 database images for training and 25% for testing.In the second one, we have used a learning data-set composed of all the 7200 images (100 objects with 72 views) without noise and a testing data-set composed of 15 randomly selected views per object to which an additive Gaussian noise with \(S_d\) of 5, 10 and 20 was added (see Fig. 4).

We evaluate separately the recognition rate obtained using the four previous invariant descriptors and the combination of the RPS & BS invariants to test their complementarity. Then, we compare their performance with the Hu’s moments (HM), the Zernike’s moments (ZM), the Fourier–Mellin transform (FM), described in Appendix 1, and the local SIFT and HOG descriptors [12] whose performance under the same conditions has been tested in [7],

Since we use the RBF kernel in the SVM classification process, this depends on the kernel size \(\sigma \). The results presented here are obtained by choosing empirically the value \(\sigma _{opt}\) that provided maximum recognition rate.

5.3 Experiments

The performances of the different invariant descriptors are analyzed with respect to the recognition rate given a learning set. Hence, for a given ratio, the learning and testing sets have been built by splitting randomly all examples. Then, due to randomness of this procedure, multiple trials have been performed with different random draws of the learning and testing set. In the case of an added noise, since as mentioned before the learning set comprised all images, this procedure is applied only to the testing set.

The parameters of our experiments are the following:

  1. 1.

    The learning set \(c_i\) corresponding to the values of an invariant descriptor computed on an image from the database;

  2. 2.

    The classes \(\hat{c}_i \in \left\{ {1,100} \right\} \) corresponding to the object class.

  3. 3.

    Algorithm performance: the efficiency is given through a percentage of the well recognized objects composing the testing set.

  4. 4.

    Number of random trials: fixed to 5.

  5. 5.

    Kernel K: a Gaussian kernel of bandwidth \(\sigma \) is chosen

    $$\begin{aligned} K(x,y) = e^{\frac{{ - \left\| {x - y} \right\| ^2 }}{{2\sigma ^2 }}} \end{aligned}$$

    x and y correspond to the descriptors vectors of objects.

For solving a multi-class problem, the two most popular approaches are the one-against-all (OAA) method and the one-against-one (OAO) method [29]. For our purpose, we chose an OAO SVM, because it is substantially faster to train and seems preferable for problems with a very large number of classes.

5.3.1 COIL-100 Databases

The Columbia Object Image Library (COIL-100, Fig. 2) is a database of color images of 100 different objects, where 72 images of each object were taken at pose intervals of \(5^\circ \).

5.3.2 Classification Performance

Table 2 presents results obtained testing our object recognition method with the COIL-100 database. The best results were achieved using the local SIFT descriptor. The RBS comes in the second place and the local HOG features come third. Indeed it has been demonstrated in the literature, these local methods currently give the best results. However, if noise is added on the image, the use of global approach is better than the use of local ones. The main reason is that the key-points detector used in the local method produce in these cases many key-points that are nor relevant for object recognition. This will be shown in the next subsection.

Fig. 2
figure 2

Sample objects of COIL-100 database

Table 2 Recognition rate for each descriptor using the COIL-100 database. The test results for ZM, HM, FM, and SIFT are taken from [7]
Fig. 3
figure 3

Classification rate for different size of the training database. The test results for ZM, HM, FM, and SIFT are taken from [7]

Table 3 Classification rate on COIL-100 noisy database. The test results for ZM, HM, FM, and SIFT are taken from [7]

In Fig. 3 we present the recognition rate as a function of the size of the training set. As expected, this is an increasing function and we remark that the RBS and the combination of the RPS and the BS give better results than the other global invariant descriptors.

5.3.3 Robustness Against Noise

Also in this case, test results for ZM, HM, FM, and SIFT are taken from [7].

Results presented in Table 3 show that noise has little influence on classification performance when we use a global descriptor such as RBS, BS, the combination of BS & RPS, ZM, HM and FM. It has however a sensible effect on the SIFT local descriptor, and a big one on the HOG local descriptor.

5.3.4 The ORL Database

The Cambridge University ORL face database (Fig. 5) is composed of 400 gray level images of ten different patterns for each of 40 persons. The variations of the images are across time, size, pose and facial expression (open/closed eyes, smiling/not smiling), and facial details (glasses/no glasses).

In the literature, the protocol used for training and testing is different from one paper to another. In [36], a hidden Markov model (HMM) based approach is used, and the best model resulted in recognition rate of 95%, with high computational cost. In [21], Hjelmas reached a 85% recognition rate using the ORL database and feature vector consisting of Gabor coefficients.

We perform experiments on the ORL database using the RBS, BS, PS, RPS, ZM, HU, FM, and the combination of the RPS & BS descriptors. Since the local descriptors SIFT and HOG obtained, predictably, almost perfect scores, we do not present them. The results are shown in Table 4, where we clearly see that the RBS invariant descriptor gives the best recognition rate \(c = 89.8\%\), faring far better than before w.r.t. the combination of RPS and BS descriptors.

Fig. 4
figure 4

Sample of COIL-100 noisy object

Fig. 5
figure 5

Face samples from the ORL database

Table 4 Recognition rate for each descriptor using the ORL database

6 Conclusion and Perspectives

In this paper, we presented four Fourier descriptors over the semidiscrete roto-translation group SE(2, N). Then, we proved that the generalized power-spectrum (\({{\mathrm{PS}}}\)) and bispectrum (\({{\mathrm{BS}}}\))—and thus the rotational power-spectrum (\({{\mathrm{RPS}}}\)) and bispectrum (\({{\mathrm{RBS}}}\))—are weakly complete, in the sense that they allow to distinguish over an open and dense set of compactly supported functions \(\varphi \in L^2(SE(2,N))\) up to the SE(2, N) action. This generalizes a result of [38]. We then considered a framework for the application of these Fourier descriptors to roto-translation invariant object recognition, inspired by some neurophysiological facts on the human primary visual cortex. In this framework, we showed that the rotational bispectrum is indeed a weakly complete roto-translation invariant for planar images. Moreover, although the proposed Fourier descriptors are given in terms of complex mathematical objects, we showed that they can be implemented in a straightforward way as linear combinations of the values of the 2D Fourier transform of the image.

In the second part of the paper, we proposed an evaluation of the performances of these Fourier descriptors in object recognition and we presented the results obtained on different databases: the COIL-100 database, composed of several objects undergoing 3D rotation and scales changes, and the ORL-database, on which different human faces are subject to several kinds of variations. For both these databases, the global Fourier descriptors introduced in this paper are the most efficient global descriptors tested, equalled only, for noisy images, by the Zernike Moments. Although for unperturbed images the local SIFT descriptor gives better recognition rate, the addition of noise leads to the global descriptors outperforming the local ones. These results thus show the rotational bispectrum (\({{\mathrm{RBS}}}\)) to be a very good Fourier descriptor for object recognition, consistently with the theoretical weak completeness result. When the dimension of the feature vector is an issue, the \({{\mathrm{RBS}}}\) can be replaced by a combination of the generalized bispectrum (\({{\mathrm{BS}}}\)) and the rotational power-spectrum (\({{\mathrm{RPS}}}\)), which yields slightly worse results with a feature vector of length almost one-third.

An extension of the object recognition method presented in this paper to an AdaBoost framework for the problem of object detection is currently ongoing.