Nonholonomic Orthogonal Learning Algorithms for Blind Source Separation

Similar documents
Natural Gradient Learning for Over- and Under-Complete Bases in ICA

ON SOME EXTENSIONS OF THE NATURAL GRADIENT ALGORITHM. Brain Science Institute, RIKEN, Wako-shi, Saitama , Japan

To appear in Proceedings of the ICA'99, Aussois, France, A 2 R mn is an unknown mixture matrix of full rank, v(t) is the vector of noises. The

1 Introduction Independent component analysis (ICA) [10] is a statistical technique whose main applications are blind source separation, blind deconvo

MULTICHANNEL BLIND SEPARATION AND. Scott C. Douglas 1, Andrzej Cichocki 2, and Shun-ichi Amari 2

ICA [6] ICA) [7, 8] ICA ICA ICA [9, 10] J-F. Cardoso. [13] Matlab ICA. Comon[3], Amari & Cardoso[4] ICA ICA

Independent Component Analysis on the Basis of Helmholtz Machine

where A 2 IR m n is the mixing matrix, s(t) is the n-dimensional source vector (n» m), and v(t) is additive white noise that is statistically independ

PROPERTIES OF THE EMPIRICAL CHARACTERISTIC FUNCTION AND ITS APPLICATION TO TESTING FOR INDEPENDENCE. Noboru Murata

ON-LINE BLIND SEPARATION OF NON-STATIONARY SIGNALS

A METHOD OF ICA IN TIME-FREQUENCY DOMAIN

A Constrained EM Algorithm for Independent Component Analysis

Blind Extraction of Singularly Mixed Source Signals

Analytical solution of the blind source separation problem using derivatives

Independent Component Analysis and Blind Source Separation

CIFAR Lectures: Non-Gaussian statistics and natural images

Using Kernel PCA for Initialisation of Variational Bayesian Nonlinear Blind Source Separation Method

An Improved Cumulant Based Method for Independent Component Analysis

Gatsby Theoretical Neuroscience Lectures: Non-Gaussian statistics and natural images Parts I-II

TWO METHODS FOR ESTIMATING OVERCOMPLETE INDEPENDENT COMPONENT BASES. Mika Inki and Aapo Hyvärinen

Independent Component Analysis. Contents

Simultaneous Diagonalization in the Frequency Domain (SDIF) for Source Separation

A Cross-Associative Neural Network for SVD of Nonsquared Data Matrix in Signal Processing

Basic Principles of Unsupervised and Unsupervised

x 1 (t) Spectrogram t s

Independent Component Analysis and Its Applications. By Qing Xue, 10/15/2004

On Information Maximization and Blind Signal Deconvolution

Massoud BABAIE-ZADEH. Blind Source Separation (BSS) and Independent Componen Analysis (ICA) p.1/39

SPARSE REPRESENTATION AND BLIND DECONVOLUTION OF DYNAMICAL SYSTEMS. Liqing Zhang and Andrzej Cichocki

A Canonical Genetic Algorithm for Blind Inversion of Linear Channels

Non-Euclidean Independent Component Analysis and Oja's Learning

1 Introduction Blind source separation (BSS) is a fundamental problem which is encountered in a variety of signal processing problems where multiple s

Blind Source Separation with a Time-Varying Mixing Matrix

Fundamentals of Principal Component Analysis (PCA), Independent Component Analysis (ICA), and Independent Vector Analysis (IVA)

Neural Learning in Structured Parameter Spaces Natural Riemannian Gradient

BLIND SEPARATION OF INSTANTANEOUS MIXTURES OF NON STATIONARY SOURCES

One-unit Learning Rules for Independent Component Analysis

Equivalence Probability and Sparsity of Two Sparse Solutions in Sparse Representation

Independent Component Analysis of Incomplete Data

Robust techniques for independent component analysis (ICA) with noisy data

Constrained Projection Approximation Algorithms for Principal Component Analysis

Undercomplete Independent Component. Analysis for Signal Separation and. Dimension Reduction. Category: Algorithms and Architectures.

IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 54, NO. 2, FEBRUARY

FREQUENCY-DOMAIN IMPLEMENTATION OF BLOCK ADAPTIVE FILTERS FOR ICA-BASED MULTICHANNEL BLIND DECONVOLUTION

ADAPTIVE LATERAL INHIBITION FOR NON-NEGATIVE ICA. Mark Plumbley

Recursive Generalized Eigendecomposition for Independent Component Analysis

BLIND SEPARATION USING ABSOLUTE MOMENTS BASED ADAPTIVE ESTIMATING FUNCTION. Juha Karvanen and Visa Koivunen

BLIND SEPARATION OF POSITIVE SOURCES USING NON-NEGATIVE PCA

Robust extraction of specific signals with temporal structure

Variational Principal Components

ACENTRAL problem in neural-network research, as well

MINIMIZATION-PROJECTION (MP) APPROACH FOR BLIND SOURCE SEPARATION IN DIFFERENT MIXING MODELS

Estimation of linear non-gaussian acyclic models for latent factors

Wavelet de-noising for blind source separation in noisy mixtures.

Discriminative Direction for Kernel Classifiers

Independent Component Analysis

CONTROL SYSTEMS ANALYSIS VIA BLIND SOURCE DECONVOLUTION. Kenji Sugimoto and Yoshito Kikkawa

SPARSE signal representations have gained popularity in recent

Blind channel deconvolution of real world signals using source separation techniques

c Springer, Reprinted with permission.

Dimension Reduction Techniques. Presented by Jie (Jerry) Yu

Blind Machine Separation Te-Won Lee

NONLINEAR BLIND SOURCE SEPARATION USING KERNEL FEATURE SPACES.

Feature Extraction with Weighted Samples Based on Independent Component Analysis

Post-Nonlinear Blind Extraction in the Presence of Ill-Conditioned Mixing Wai Yie Leong and Danilo P. Mandic, Senior Member, IEEE

Independent Component Analysis

Emergence of Phase- and Shift-Invariant Features by Decomposition of Natural Images into Independent Feature Subspaces

Sensitivity Analysis for the Problem of Matrix Joint Diagonalization

Unsupervised learning: beyond simple clustering and PCA

Independent Component Analysis (ICA)

Blind Source Separation for Changing Source Number: A Neural Network Approach with a Variable Structure 1

Linking non-binned spike train kernels to several existing spike train metrics

Simple LU and QR based Non-Orthogonal Matrix Joint Diagonalization

Least square joint diagonalization of matrices under an intrinsic scale constraint

Local minima and plateaus in hierarchical structures of multilayer perceptrons

Statistical Pattern Recognition

Deep learning / Ian Goodfellow, Yoshua Bengio and Aaron Courville. - Cambridge, MA ; London, Spis treści

Bayesian ensemble learning of generative models

FASTGEO - A HISTOGRAM BASED APPROACH TO LINEAR GEOMETRIC ICA. Andreas Jung, Fabian J. Theis, Carlos G. Puntonet, Elmar W. Lang

Complete Blind Subspace Deconvolution

A Strict Stability Limit for Adaptive Gradient Type Algorithms

Speed and Accuracy Enhancement of Linear ICA Techniques Using Rational Nonlinear Functions

Distinguishing Causes from Effects using Nonlinear Acyclic Causal Models

Nonlinear Blind Source Separation Using a Radial Basis Function Network

A two-layer ICA-like model estimated by Score Matching

GENERALIZED DEFLATION ALGORITHMS FOR THE BLIND SOURCE-FACTOR SEPARATION OF MIMO-FIR CHANNELS. Mitsuru Kawamoto 1,2 and Yujiro Inouye 1

Automatic Rank Determination in Projective Nonnegative Matrix Factorization

Mark your answers ON THE EXAM ITSELF. If you are not sure of your answer you may wish to provide a brief explanation.

Problems in Linear Algebra and Representation Theory

Blind separation of sources that have spatiotemporal variance dependencies

Dictionary Learning for L1-Exact Sparse Coding

Neural Network Training

Independent Component Analysis Using an Extended Infomax Algorithm for Mixed Subgaussian and Supergaussian Sources

L26: Advanced dimensionality reduction

ICA. Independent Component Analysis. Zakariás Mátyás

BLIND DECONVOLUTION ALGORITHMS FOR MIMO-FIR SYSTEMS DRIVEN BY FOURTH-ORDER COLORED SIGNALS

Blind separation of instantaneous mixtures of dependent sources

Lecture 5: Logistic Regression. Neural Networks

CPSC 340: Machine Learning and Data Mining. More PCA Fall 2017

Transcription:

LETTER Communicated by Klaus Obermayer Nonholonomic Orthogonal Learning Algorithms for Blind Source Separation Shun-ichi Amari Tian-Ping Chen Andrzej Cichocki RIKEN Brain Science Institute, Brain-Style Information Systems Group, Japan Independent component analysis or blind source separation extracts independent signals from their linear mixtures without assuming prior knowledge of their mixing coefficients. It is known that the independent signals in the observed mixtures can be successfully extracted except for their order and scales. In order to resolve the indeterminacy of scales, most learning algorithms impose some constraints on the magnitudes of the recovered signals. However, when the source signals are nonstationary and their average magnitudes change rapidly, the constraints force a rapid change in the magnitude of the separating matrix. This is the case with most applications (e.g., speech sounds, electroencephalogram signals). It is known that this causes numerical instability in some cases. In order to resolve this difficulty, this article introduces new nonholonomic constraints in the learning algorithm. This is motivated by the geometrical consideration that the directions of change in the separating matrix should be orthogonal to the equivalence class of separating matrices due to the scaling indeterminacy. These constraints are proved to be nonholonomic, so that the proposed algorithm is able to adapt to rapid or intermittent changes in the magnitudes of the source signals. The proposed algorithm works well even when the number of the sources is overestimated, whereas the existent algorithms do not (assuming the sensor noise is negligibly small), because they amplify the null components not included in the sources. Computer simulations confirm this desirable property. 1 Introduction Blind source separation is the problem of recovering a number of independent random or deterministic signals when only their linear mixtures are available. Here, blind implies that the mixing coefficients, the original signals, and their probability distributions are not known a priori. The problem is also called independent component analysis (ICA) or blind extraction of source signals. The problem has become increasingly important because of many potential applications, such as those in speech recognition and en- Neural Computation 12, 1463 1484 (2) c 2 Massachusetts Institute of Technology

1464 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki hancements, telecommunication, and biomedical signal analysis and processing. However, there still remain open theoretical problems to be solved further. The source signals can be identified only up to a certain ambiguity, that is, up to a permutation and arbitrary scaling of original signals. In order to resolve the scaling indeterminacy, most learning algorithms developed so far impose some constraints on the power or amplitude of the separated signals (Jutten & Hérault, 1991; Hérault & Jutten, 1986; Bell & Sejnowski, 1995; Common, 1994; Douglas, Cichocki, & Amari, 1997; Sorouchyari, 1991; Chen & Chen, 1994; Cichocki, Unbehauen, & Rummert, 1994; Cichocki, Unbehauen, Moszczynski, & Rummert, 1994; Cichocki, Amari, Adachi, & Kasprzak, 1996; Cichocki, Bogner, Moszczynski, & Pope, 1997; Cardoso & Laheld, 1996; Girolami & Fyfe, 1997a, 1997b; Amari, Chen, & Cichocki, 1997; Amari, Cichocki, & Yang, 1995, 1996; Amari & Cardoso, 1997; Amari, 1967; Amari & Cichocki, 1998). When the source signals are stationary and their average magnitudes do not change over time, the constraints work well to extract the original signals in fixed magnitudes. However, the source signals change their local average magnitudes in many practical applications such as evoked potentials and speech by humans. In such cases, when a component source signal becomes very small suddenly, the corresponding row of the separating matrix tends to be large in the learning process to compensate for this change and to emit the output signal of the same amplitude. In particular, when one source signal becomes silent, the separating matrix diverges. This causes serious numerical instability. This article proposes an efficient learning algorithm with equivariant properties that works well even when the magnitudes of the source signals change quickly. To this end, we first introduce an equivalence relation in the set of all the demixing matrices. Two matrices are defined to be equivalent when the recovered signals differ only by scaling. When the trajectory of the learning dynamics of the demixing matrix moves in a single equivalence class, the recovered signals remain the same except for scaling. Therefore, such a motion is ineffective in learning. In order to reduce ineffective movements in the learning process, we modify the natural gradient learning rule (Amari, Cichocki, & Yang, 1996; Amari, 1998) such that the trajectory of learning is orthogonal to the equivalence class. This imposes constraints on the directions of movements in learning. However, such constraints are proved to be nonholonomic, implying that there are no integral submanifolds corresponding to the constraints. This implies that the trajectory can reach any demixing matrices in spite of the constraints, a property that is useful in robotic design. Nonholonomic dynamics is a current research topic in robotics. Since the trajectory governed by the proposed algorithm does not include useless (redundant) components, the algorithm can be more efficient (with respect to convergence and efficiency) than those algorithms with other

Nonholonomic Orthogonal Learning Algorithms 1465 constraints. It is also robust against sudden changes in the magnitudes of the source signals. Even when some components disappear, the algorithm still works well. Therefore, our algorithm can be used even when the number of source signals is overestimated. These features have been confirmed by computer simulations. The stability conditions for the proposed algorithm are also derived. 2 Critical Review of Existing Adaptive Learning Algorithms The problem of blind source separation is formulated as follows (Hérault & Jutten, 1986; Jutten & Hérault, 1991). Let n source signals s 1 (t),...,s n (t) at time t be represented by a vector s(t) = [s 1 (t) s n (t)] T, where the superscript T denotes transposition of a vector or matrix. They are assumed to be stochastically independent. We assume that the source signals are not directly accessible, and we can instead observe n signals x 1 (t),...,x n (t) summarized in a vector x(t) = [x 1 (t) x n (t)] T, which are instantaneous linear mixtures of the n source signals. Therefore, we have x(t) = As(t), (2.1) where A = [a ij ] n n is an unknown nonsingular mixing matrix. The source signals s(1), s(2),...,s(t) are generated at discrete times t = 1, 2,...,T subject to unknown probability distributions. In order to recover the original source signals s(t), we make new linear mixtures of the sensor signals x(t) by y(t) = W (t)x(t), (2.2) where W = [w ij ] n n is called the demixing or separating matrix. If W (t) = A 1, the original source signals are recovered exactly. However, A or W is unknown, so we need to estimate W from the available observations x(1),...,x(t). Here, W (t) is an estimate. It is well known that the mixing matrix A, and hence source signals, are not completely identifiable. There are two types of indeterminacies or ambiguities: the order of the original source signals is indefinite, and the scales of the original signals are indefinite. The first indeterminacy implies that a permutation or reordering of recovered signals remains possible, but this is not critical, because what is wanted is to extract n independent signals; their ordering is not important. The second indeterminacy can be explained as follows. When the signals s are transformed as s = Λs, where Λ is a diagonal scaling matrix with nonzero diagonal entries, the observed signals are kept invariant if the mixing matrix A is changed into A = AΛ 1, because x = As = A s. This implies that W and W = ΛW are equivalent in the sense that they give the same estimated source signals except for their scales. Taking these indeterminacies into account, we can estimate

1466 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki W = A 1 P Λ at best, where P is a permutation matrix and Λ is a diagonal matrix. These intrinsic (unavoidable) ambiguities do not seem to be a severe limitation, because in most applications, most information is contained in the waveforms of the signals (for instance, speech signals and multisensor biomedical recordings) rather than their magnitudes and the order in which they are arranged. In order to make the scaling indeterminacies clearer, we introduce an equivalence relation in the n 2 -dimensional space of nonsingular matrices Gl(n) ={W }. We define W and ΛW as equivalent, W ΛW, (2.3) where Λ is an arbitrary nonsingular diagonal matrix. Given W, its equivalence class C W ={W W = ΛW } (2.4) is an n-dimensional subspace of Gl(n) consisting of all the matrices equivalent to W. Therefore, a learning algorithm need not search for the nonidentifiable W = A 1 but searches for the equivalence class C W that contains the true W = A 1 except for permutations. In existing algorithms, in order to determine the scales, one of two constraints is imposed explicitly or implicitly. The first is a rigid constraint. Let us introduce n restrictions on W, f i (W ) =, i = 1,...,n. (2.5) A typical one is f i (W ) = w ii 1, i = 1,...,n. Define the set Q r of W s satisfying the above constraints to be Q r ={W f i (W ) =, i = 1,...,n}. (2.6) The set Q r is supposed to form a (n 2 n)-dimensional manifold transversal to C W. In order to resolve the scaling indeterminacies, we require that W (t) = [ w ij (t) ] always be included in the Q r. Such a rigid constraint is implemented, for example, by the learning algorithm w ij (t + 1) = w ij (t) η(t)ϕ i (y i (t))y j (t), i j, (2.7) where ϕ(y i ) are suitably chosen nonlinear functions, η(t) is the step size, and all the diagonal elements w ii are fixed equal to 1.

Nonholonomic Orthogonal Learning Algorithms 1467 This idea of reducing the complexity of a learning algorithm slightly is a simple one and has been used in several papers (Hérault & Jutten, 1986; Jutten & Hérault, 1991; Sorouchyari, 1991; Chen & Chen, 1994; Moreau & Macchi, 1995; Matsuoka, Ohya, & Kawamoto, 1995; Oja & Karhunen, 1995). However, the constraints (see equation 2.5) do not allow equivariant properties (Cardoso & Laheld, 1996). Therefore, when A is ill conditioned, the algorithm does not work well. The second constraint specifies the power or energy of the recovered (estimated) signals. Such soft constraints can be expressed in a general form as E[h i (y i )] =, i = 1,...,n, (2.8) where E denotes the expectation and h i (y i ) is a nonlinear function; typically or h i (y i ) = 1 ϕ i (y i )y i (2.9) h i (y i ) = 1 y 2 i. (2.1) The latter guarantees that the variances or powers of the recovered signals y i are 1. Such soft constraints are implemented automatically in the adaptive learning rule, W (t + 1) = W (t) + η(t)f [y(t)]w (t), (2.11) where F [y(t)] isann n matrix of the form (Cichocki, Unbehauen, & Rummert, 1994; Amari et al., 1996) F [y(t)] = I ϕ(y(t))y T (t), (2.12) (or Amari et al., 1997) F[y(t)] = I αϕ(y(t))y T (t) + βy(t)ϕ T (y(t)) (2.13) with adaptively or adequately chosen constants α and β, or (Cardoso & Laheld, 1996), F [y(t)] = I y(t)y T (t) ϕ(y(t))y T (t) + y(t)ϕ(y T (t)). (2.14) The diagonal elements of the matrix F (y) in equation 2.11 can be regarded as specifying the constraints, E [ h i (y i ) ] =. (2.15)

1468 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki Then, although W (t) can move freely in C W in the learning process, any equilibrium point of the averaged learning equation satisfies the constraints of equation 2.15. Let us define the set Q s of W s by Q s ={W E[h i (y i )] =, i = 1,...,n}. (2.16) Then Q s is an (n 2 n)-dimensional attracting submanifold and W (t) is attracted toward Q s, although it does not stay exactly in Q s (t) because of stochastic fluctuations. Summarizing, rigid constraints do not permit algorithms with equivariant properties (Cardoso & Laheld 1996). Soft constraints enable us to construct algorithms with equivariant properties. However, for source signals whose amplitudes or powers change over time (e.g., signals with long silent periods and spikes or rapid jumps like speech signals), it is more difficult to achieve rapid convergence to equilibria because of inadequate specification of constraints than in the stationary case. Special tricks or techniques may be employed, like random selection of training samples and dilation of silence periods (Bell & Sejnowski, 1995; Amari, Douglas, & Cichocki, 1998). Performance may be poorer for such nonstationary signals. Moreover, it is easy to show that such algorithms 2.11 through 2.14 are unstable (in the sense that weights w ij (t) diverge to infinity) in cases when the number of the true sources is fewer than the number of the observed signals, where W is a quadratic n n matrix. Such a case arises often in practice when one or more source signals are silent or decay to zero in an unpredictable way. In this article, we will elucidate the roles of such constraints geometrically. We then show that such constraints and drawbacks can be lifted or at least relaxed by introducing a new algorithm. In other words, we impose no explicit constraints for the separating matrix W and search for the equivalence class C W directly. This article is theoretical in nature, and simulations are limited. Larger-scale simulations studies and applications to real data are necessary for proving the practical usefulness of the proposed method. 3 Derivation of a New Learning Algorithm In order to derive a new algorithm, we recapitulate the derivation of the natural gradient algorithm due to Amari, Chen, & Cichocki (1997). To begin, we consider the loss function, l(y, W ) = log det(w ) m log p i (y i ), (3.1) i=1 where p i (y i ) are some positive probability density functions (p.d.f.) of output

Nonholonomic Orthogonal Learning Algorithms 1469 signals and det(w ) denotes the determinant of matrix W. We can then apply the stochastic gradient descent learning method to derive a learning rule. In order to calculate the gradient of l, we derive the total differential dl of l when W is changed from W to W + dw. In component form, dl = l(y, W + dw ) l(y, W ) = i,j l w ij dw ij, (3.2) where the coefficients l/ w ij of dw ij represent the gradient of l. Simple algebraic and differential calculus yields dl = tr(dww 1 ) + ϕ(y) T dy, (3.3) where tr is the trace of a matrix and ϕ(y) is a column vector whose components are ϕ i (y i ) = d log(p i(y i ) dy = ṗi(y i ) i p i (y i ), (3.4) denoting differentiation. From y = Wx, we have dy = dwx= dww 1 y. (3.5) Hence, we define dx = dww 1, (3.6) whose components dx ij are linear combinations of dw ij. The differentials dx ij form a basis of the tangent space at W of Gl(n), since they are linear combinations of the basis dw ij. The space Gl(n) of all the nonsingular matrices is a Lie group. An element W is mapped to the identity I when W 1 is multiplied from the right. A small deviation dw at W is mapped to dx = dww 1 at I. By using dx, we can compare deviations dw 1 and dw 2 at two different points W 1 and W 2, using dx 1 = dw 1 W 1 1 and dx 2 = dw 2 W 1 2. The magnitude of deviation dx is measured by the inner product ( dx, dx I = tr dxdx T) = ( ) 2 dxij. (3.7) i,j This implies that the inner product of dw at W is given by { ( dw, dw W = tr dww 1 dww 1) } T, (3.8)

147 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki which is different from (dw ij ) 2. Therefore, the space Gl(n) has a Riemannian metric. It should be noted that dx = dww 1 is a nonintegrable differential form so that we do not have a matrix function X(W ) which gives equation 3.6. Nevertheless, the nonholonomic basis dx/ is useful. Since the basis dx is orthonormal in the sense that dx 2 is given by equation 3.7, it is effective to analyze the learning equation in terms of dx. The natural Riemannian gradient (Amari, Cichocki, & Yang, 1995; Amari et al., 1996, 1998) is automatically implemented by it and the equivariant properties automatically hold. It is easy to rewrite the results in terms of dw by using equation 3.6. The gradient dl in equation 3.3 is expressed in the differential form: dl = tr(dx) + ϕ(y) T dxy. (3.9) This leads to the stochastic gradient learning algorithm, X(t)=X(t + 1) X(t)= η(t) dl [ ] dx =η(t) I ϕ(y(t))y T (t), (3.1) in terms of X(t) = W (t)w 1 (t). This is rewritten as [ ] W (t + 1) = W (t) + η(t) I ϕ(y(t))y T (t) W (t) (3.11) in terms of W (t). It was shown that the diagonal term in equation 3.11 can be set arbitrarily (Cichocki, Unbehauen, Moszczynski, & Rummert, 1994; Cichocki & Unbehauen, 1996; Amari & Cardoso, 1997). The problem has been analyzed also based on the information geometry of semiparametric statistical models (Amari & Kawanabe, 1997a, 1997b; Bickel, Klaassen, Ritov, & Wellner, 1993). Therefore, the above algorithm can be generalized in a more flexible and universal form as [ ] W (t + 1) = W (t) + η(t) Λ(t) ϕ(y(t))y T (t) W (t), (3.12) where Λ(t) is any positive definite scaling diagonal matrix. By determining Λ correctly depending on y, we have various algorithms. We propose Λ(t) = diag { y 1 (t)ϕ 1 (y 1 (t)),...,y n (t)ϕ n (y n (t)) }. This algorithm is effective when the nonlinear activation functions ϕ i (y i ) are suitably chosen. This corresponds to setting new constraints dx ii (t) =, i = 1,...,n. (3.13)

Nonholonomic Orthogonal Learning Algorithms 1471 The new constraints lead to a new learning algorithm, written as W (t + 1) = W (t) η(t)f [y(t)]w (t), (3.14) where all the elements on the main diagonal of the n n matrix F (y) are put equal to zero, that is, and f ii = (3.15) f ij = ϕ i (y i )y j, i j. (3.16) The learning rule, 3.14, is called the orthogonal nonholonomic natural gradient descent algorithm for the reason shown in the next section. Algorithm 3.14 can be generalized as follows (Amari et al. 1997), W (t + 1) = W (t) η(t)f {y(t)}w (t), (3.17) with entries of the matrix F (y) as f ij = δ ij λ ii + α 1 y i y j + α 2 ϕ i (y i )y j α 3 y i ϕ i (y j ), (3.18) where δ ij is Kronecker delta, λ ii = α 1 y 2 i α 2ϕ i (y i )y i + α 3 y i ϕ i (y i ) and α 1,α 2, and α 3 are parameters that may be determined adaptively. 4 Nonholonomicity and Orthogonality The space Gl(n) ={W } is partitioned into equivalence classes C W, where any W belonging to the same C W is regarded as an equivalent demixing matrix. Let dw C be a direction tangent to C W, that is, W + dw C and W belong to the same equivalence class. The learning equation determines W (t) depending on the current y and W. When W (t) includes components belonging to the tangent directions to C W, such components are ineffective because they drive W within the equivalent class. Therefore, it is better to design a learning rule such that its trajectories W (t) are always orthogonal to the equivalence classes. Since C W are n-dimensional subspaces, if we could find a family of n 2 n-dimensional subspaces Q s such that C W and Q are orthogonal, we could impose the constraints that the learning trajectories would belong to Q. Therefore, one interesting question arises: Is there a family of (n 2 n)-dimensional submanifolds Q that are orthogonal to the equivalence classes C W? The answer is no. We can prove now that there does

1472 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki not exist a submanifold that is orthogonal to the families of submanifolds C W s. Theorem 1. The direction dw is orthogonal to C W, if and only if dx ii =, i = 1,...,n (4.1) where dx = dww 1. Proof. Since the equivalent class C W consists of matrices ΛW, Λ is regarded as a coordinate system in C W. A small deviation of W in C W is written as dw C = dλw, (4.2) where dλ is diag(dλ 1,...,dλ n ). The tangent space of C W is spanned by them. The inner product of dw and dw C is given by dw, dw C W = dww 1, dw C W 1 I = dx, dλ I = i,j dx ij dλ ij = dx ii dλ ii. (4.3) Therefore, dw is orthogonal to C W when and only when dw satisfies dx ii = for all i. We next show that dx is not integrable, that is, there are no matrix functions G(W ) such that dx = G W where G W dw = i,j dw, (4.4) G w ij dw ij. (4.5) If such G exists, X = G(W ) defines another coordinate system in Gl(n). Even when such G does not exist, dx is well defined, and it forms a basis in the tangent space of Gl(n) at W. Such a basis is called a nonholonomic basis (Schouten, 1954; Frankel, 1997). Theorem 2. The basis defined by dx = dww 1 is nonholonomic.

Nonholonomic Orthogonal Learning Algorithms 1473 Proof. Let us assume that there exists a function G(W ) such that dx = dg(w ) = dww 1. (4.6) We now consider another small deviation from W to W + δw. We then have δdx = δdg = dw δ(w 1 ). (4.7) We have Hence, δ(w 1 ) = (W + δw ) 1 W 1 W 1 δww 1. (4.8) δdx = dww 1 δww 1. (4.9) Since matrices are in general noncommutative, we have δdx dδx. (4.1) This shows that there does not exist a matrix X because dδg = δdg always hold when a matrix G exists. Our constraints dx ii =, i = 1,...,n (4.11) restrict the possible directions of W and define (n 2 n)-dimensional movable directions at each point W of Gl(n). These directions are orthogonal to C W. However, by the same reasoning, there do not exist functions g i (W ) such that dx ii = dg i (W ) = j,k g i (W ) w jk dw jk. (4.12) This implies that there exist no subspaces defined by g i (W ) =. Such constraints are said to be nonholonomic. The learning equation, 3.14, with equation 3.15 defines a learning dynamics with nonholonomic constraints. At each point W, W is constrained in (n 2 n) directions. However, the trajectories are not constrained in any (n 2 n)-dimensional subspace, and they can reach any points in Gl(n).

1474 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki This property is important when the amplitudes of s i (t) change over time. If the constraints are E [ h i (y i ) ] =, (4.13) for example, [ ] E y 2 i 1 =, (4.14) when E[s 2 i ] suddenly become 1 times smaller, the learning dynamics with the soft constraints make the ith row of W 1 times larger in order to compensate for this change and keep E[y 2 i ] = 1. Therefore, even when W converges to the true C W, large fluctuations emerge from the ineffective movements caused by changes in amplitude of s i. This sometimes causes numerical instability. On the other hand, our nonholonomic dynamics is always orthogonal to C W so that such ineffective fluctuations are suppressed. This is confirmed by computer simulations a few of which are presented in section 6. Remark 1. There is a great deal of research on nonholonomic dynamics. Classical analytical dynamics uses nonholonomic bases to analyze the spinning gyro (Whittaker, 194). It was also used in relativity (Whittaker, 194). Kron (1948) used the nonholonomic constraints to present a general theory of electromechanical dynamics of generators and motors. Nonholonomic properties play a fundamental role in continuum mechanics of distributed dislocations (see, e.g., Amari, 1962). An excellent explanation is found in Brockett (1993), where the controllability of nonlinear dynamical systems is shown by using the related Lie algebra. Recently robotics researchers have been eager to analyze dynamics with nonholonomic constraints (Suzuki & Nakamura 1995). Remark 2. Theorem 2 shows that the orthogonal natural gradient descent algorithm evolves along a trajectory path that does not include redundant (useless) components in the directions of C W. Therefore, it seems likely that the orthogonal algorithm can be more efficient than other algorithms, as has been confirmed by preliminary computer simulations. 5 Stability Analysis In this section, we discuss the local stability of the orthogonal natural gradient descent algorithm, 3.14, defined by W (t) = η(t)f {y(t)}w (t), (5.1)

Nonholonomic Orthogonal Learning Algorithms 1475 with f ii = and f ij = ϕ i (y i )y j,ifi j, in the vicinity of the desired equilibrium submanifold C A 1. Similar to a theorem established in (Amari et al., 1997a), we formulate the following theorem: Theorem 3. When the zero-mean independent source signals satisfy conditions E(ϕ i (s i )s j ) =, if i j (5.2) E( ϕ i (s i )s j s k ) =, if k j (5.3) and when the following inequalities hold, E( ϕ i (s i )s 2 j )E( ϕ j(s j )s 2 i )>E(s iϕ i (s i ))E(s j ϕ i (s j )), i, j = 1,...,n, (5.4) then the desired equivalence class C W, W = A 1 P Λ is a stable equilibrium. However, if conditions 5.2 and 5.3 are still satisfied and E( ϕ i (s i )s 2 j )E( ϕ j(s j )s 2 i )<E(s iϕ i (s i ))E(s j ϕ i (s j )), (5.5) then replacing f ij = ϕ(y i )y j by f ij = y i ϕ j (y j ), i j (5.6) it is guaranteed that the class C W, W = A 1 P Λ, contains stable separating equilibria. Proof. Let W = A 1 and (dw ) i denote the ith row of dw and (W 1 ) j be the jth column of W 1. Then taking into account that dx ij = ϕ i (y i )y j and dx ii =, we evaluate the second differentials in the Hessian matrix as follows: d 2 x ij = E{d(ϕ i (y i )y j )} (5.7) = E{ ϕ i (y i )y j dy i }+E{ϕ i (y i )dy j } = E{ ϕ i (y i )y j (dw ) i (W 1 s)}+e{ϕ i (y i )(dw ) j (W 1 s)} = E{ ϕ i (y i )y j (dw ) i (W 1 ) j }+E{ϕ i (y i )y i (dw ) j (W 1 ) i } = E{ ϕ i (s i )sj 2 }dx ij E{ϕ i (s i )s i }dx ji.

1476 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki Hence d 2 x ij = E{ ϕ i (s i )s 2 j }dx ij E{ϕ i (s i )s i }dx ji (5.8) d 2 x ji = E{ϕ j (s j )s j }dx ij E{ ϕ j (s j )s 2 i }dx ji (5.9) d 2 x ii =. (5.1) By the assumptions made in theorem 3, we see that the equilibrium point W = A 1 is stable. The proof for W = A 1 P Λ is the same and is omitted here for brevity. The proof for the other cases is similar. From the proof of theorem 2, we have Corollary. Let us consider the special case that all neurons have the same activation function: the cubic nonlinearity (ϕ i (y i ) = y 3 i ). In this case, if all the source signals are subgaussian signals (with negative kurtosis), then the learning algorithm, 3.14, with f ii = and f ij = y 3 i y j,ifi j ensures local stability for the desired equilibrium points: W = A 1 P Λ. Analogically, if all the source signals are supergaussian (with positive kurtosis), then the algorithm, 3.14, with f ii = and f ij = y i yj 3,ifi j ensures that all the desired equilibrium points W = A 1 P Λ are stable equilibrium separating points. 6 Computer Simulations In our computer experiments we have assumed that source signals have generalized gaussian distributions, and corresponding nonlinear activation functions have the form ϕ(y i ) = y i r i 1 sgn(y i ) for r i 1 and ϕ(y i ) = y i /[ y i (2 r i) + ɛ] for r 1 < 1, where ɛ is small positive constant added to prevent singularity of the nonlinear function (Cichocki, Bogner, Moszczynski, & Pope, 1997; Cichocki, Sabala, & Amari, 1998). For spiky supergaussian sources the optimal parameter r i changes between zero and one. In the illustrative examples that follow, we have assumed that r i =.5 and the learning rate is fixed equal to η =.1. Preliminary computer simulation experiments have confirmed the high performance of the proposed algorithm. We will give two illustrative examples to show the efficiency of the orthogonal natural gradient adaptive algorithm and verify the validity of our theorems. We also show that the orthogonal natural gradient descent algorithm is usually more efficient than soft constraint natural gradient descent algorithms for nonstationary

Nonholonomic Orthogonal Learning Algorithms 1477 5 S O U R C E S I G N A L S 5 5 5 1 5 5 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (a) 2 S E N S O R S I G N A L S 1 1 2 1 1 2 1 1 2 1 1 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (b) Figure 1: Source and sensor signals for example 1. (a) Original source signals. (b) Mixing (sensors) signals. signals, especially when the number of sources changes dynamically in time. 6.1 Example 1. To show the effectiveness of our algorithm, we assume that the number of sources is unknown (but their number is not greater than the number of the sensors). Three natural speech signals shown in Figure 1a are mixed in four sensors by using an ill-conditioned 4 3 mixing matrix. It is assumed that the mixing matrix and the number of the sources are unknown. It should be noted that the sensor signals look almost identical (see Figure 1b). Recovered signals are shown in Figure 2a.

1478 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki 1 1 1 1 1 1 2 2 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (a) 5 5 1 1 1 1.5.5 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (b) Figure 2: Separated signals for example 1. (a) Separated source signals for zero noise using ONGD algorithm. (b) Separated signals using the standard algorithm (3.11). The algorithm was able after 5 iterations not only to estimate the waveform of original speech signals but also their number. It should be noted that one output signal quickly decays to zero. For comparison we applied the standard independent component analysis learning algorithm, 3.11, to the same noiseless data. The standard algorithm also recovered the original three signals well, but it creates a false signal, as shown in Figure 2b. Moreover, the norm of separating matrix W is increasing exponentially, as illustrated in Figure 3a. This problem can be circumvented by computing preliminarily the rank of covariance matrix of the sensor signals and then taking only the number of sensors equal to the rank, provided noise

Nonholonomic Orthogonal Learning Algorithms 1479 Figure 3: Convergence dynamic behavior of the Frobenious norm of a separating matrix W as a function of the number of iterations for (a) the standard algorithm (3.11) for noiseless case, (b) the standard algorithm (3.11) for 1% additive gaussian noise, and (c) the orthogonal natural gradient descent algorithm (3.14).

148 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki is negligible. However, such an approach is not very practical for a nonstationary environment when the number of sources changes continuously in time. When some small (1%) additive noise is added to sensors, the standard algorithm can be stabilized, as shown in Figure 3b. However, a fluctuation of the norm of separating matrix W is larger than for the orthogonal natural gradient descent algorithm (see Figures 3b and 3c). 6.2 Example 2. In this simulation, we separated seven highly nonstationary natural speech signals, as shown in Figure 4a, using seven sensors (microphones), assuming that two of sources are temporally zero (silence period). It is assumed again that the number of active source signals and the randomly chosen mixing matrix A are completely unknown. The new learning algorithm, 3.14, was able to separate all signals, as illustrated in Figure 4c. Again we have found by simulation that the standard learning algorithm, 3.11, demonstrates higher fluctuation of synaptic weights of the separating matrix. Moreover, the absolute values of the weights increase exponentially to infinity for the noiseless case, so we are not able to compare the performance of the two algorithms, because the standard ICA algorithm, 3.11, is principally unstable in the noiseless case (without computing the rank of covariance matrix of sensor signals). Even for small noise, the weights of the separating matrix take very high values and are unstable in the case of the standard ICA method. 7 Conclusion We have pointed out several drawbacks in existing adaptive algorithms for blind separation of signals for example, the problem of separation of nonstationary signals and the problem of stability if some sources decay to zero. To avoid these problems, we developed a novel orthogonal natural gradient descent algorithm with the equivariant property. We have demonstrated by computer simulation experiments that the proposed algorithm ensures a good convergence speed even for ill-conditioned problems. The algorithm solves the problem of on-line estimation of the waveforms of the source signals even when the number of sensors is larger than the number of sources that is unknown. Moreover, the algorithm works properly and precisely for highly nonstationary signals like speech signals and biomedical signals. It is proved theoretically that the trajectories governed by the algorithm do not include redundant (useless) components and the algorithm is potentially more efficient than algorithms with soft or rigid constraints. The main feature of the algorithm is its orthogonality and that it converges to an equivalent class of the separating matrices rather than isolated critical points. Local stability conditions for the developed algorithm are also given.

Nonholonomic Orthogonal Learning Algorithms 1481 5 5 5 5 5 5 5 5 1 1 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 (a) x 1 4 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (b) 2 2 1 1 2 2 1 1 2 2 2 2 2 2 1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 x 1 4 (c) Figure 4: Computer simulation for example 2. (a) Original active speech signals. (b) Mixed (microphone) signals. (c) Separated (estimated) source signals.

1482 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki References Amari, S. (1962). On some primary structures of non-riemannian elasticitytheory. In K. Kondo (Ed.), RAAG memoirs of the unifying study of basic problems in engineering and physical sciences (Vol. 3, pp. 99 18). Tokyo: Gakujutsu Bunken Fukyu-kai. Amari, S. (1967). A theory of adaptive pattern classifiers. IEEE Trans. on Electronic Computers, 3, 299 37. Amari, S. (1998). Natural gradient works efficiently in learning. Neural Computation, 1, 2692 27. Amari, S., & Cardoso, J. F. (1997). Blind source separation Semi-parametric statistical approach. IEEE Trans. on Signal Processing, 45(11), 2692 27. Amari, S., Chen, T. P., & Cichocki, A. (1997). Stability analysis of adaptive blind source separation. Neural Networks, 1(8), 1345 1351. Amari, S., & Cichocki, A. (1998). Blind signal processing neural network approaches. Proceedings IEEE, special issue on Blind Identification and Estimation, 86, 226 248. Amari, S., Cichocki, A., & Yang, H. H. (1995). Recurrent neural networks for blind separation of sources. In Proceedings of International. Symposium on Nonlinear Theory and its Applications, NOLTA-95 (pp. 37 42). Amari, S., Cichocki, A., & Yang, H. H. (1996). A new learning algorithm for blind signal separation. In D. S. Touretzky, M. C. Mozer, & M. E. Hasselmo (Eds.), Advances in neural information processing systems, 8 (pp. 457 763). Cambridge, MA: MIT Press. Amari, S., Douglas, S. C., & Cichocki, A. (1998). Multichannel blind deconvolution and equalization using the natural gradient. Manuscript submitted for publication. Amari, S., & Kawanabe, M. (1997a). Information geometry of estimating functions in semiparametric statistical models. Bernoulli, 3, 29 54. Amari, S., & Kawanabe, M. (1997b). Estimating functions in semi-parametric statistical models. In I. V. Basawa, V. P. Godambe, & R. L. Taylor (Eds.), Estimating functions (Vol. 3, pp. 65 81). Bell, A. J., & Sejnowski, T. J. (1995). An information-maximization approach to blind separation and blind deconvolution. Neural Computation, 7, 1129 1159. Bickel, P. J., Klaassen, C. A. J., Ritov, Y., & Wellner, J. A. (1993). Efficient and adaptive estimation for semi-parametric models. Baltimore, MD: Johns Hopkins University Press. Brockett, R. W. (1993). Differential geometry and design of gradient algorithms. In R. E. Green & S. T. Yau (Eds.), Differential geometry: Partial differential equations on manifolds (pp. 69 92). Providence, RI: American Mathematical Society. Cardoso, J. F., & Laheld, B. (1996). Equivariant adaptive source separation. IEEE Trans. on Signal Processing, 12, 317 33. Chen, T. P., & Chen, R. (1994). Blind extraction of stochastic and deterministic signals by neural network approach. In Proc. of 28th Asilomar Conference on Signals, Systems and Computer (pp. 892 896).

Nonholonomic Orthogonal Learning Algorithms 1483 Cichocki, A., Amari, S., Adachi, S., & Kasprzak, W. (1996). Self-adaptive neural networks for blind separation of sources. In Proceedings of Int. Symp. on Circuits and Systems (ISCAS-96) (Vol. 2, pp. 157 161). Cichocki, A., Bogner, R. E., Moszczynski, L., & Pope, K. (1997). Modified Hérault- Jutten algorithms for blind separation of sources. Digital Signal Processing, 7(2), 8 93. Cichocki, A., Sabala, I., & Amari, S. (1998). Intelligent neural networks for blind signal separation with unknown number of sources. In Proc. of Conf. Engineering of Intelligent Systems, ESI-98 (pp. 148 154). Cichocki, A., Sabala, I., Choi, S., Orsier, B., & Szupiluk, R. (1997). Self adaptive independent component analysis for sub-gaussian and super-gaussian mixtures with unknown number of sources and additive noise. In Proceedings of 1997 International Symposium on Nonlinear Theory and Its Applications (NOLTA-97) (Vol. 2, pp. 731 734). Cichocki, A., & Unbehauen, R. (1996). Robust neural networks with on-line learning for blind identification and blind separation of sources. IEEE Transaction on Circuits and Systems I: Fundamental Theory and Applications, 43, 894 96. Cichocki, A., Unbehauen, R., Moszczynski, L., & Rummert, E. (1994). A new on-line adaptive algorithm for blind separation of source signals. In Proc. of 1994 Int. Symposium on Artificial Neural Networks ISANN-94 (pp. 46 411). Cichocki, A., Unbehauen, R., & Rummert, E. (1994). Robust learning algorithm for blind separation of signals. Electronics Letters, 3(17), 1386 1387. Common, P. (1994). Independent component analysis, a new concept? Signal Processing, 36, 287 314. Douglas, S. C., Cichocki, A., & Amari, S. (1997). Multichannel blind separation and deconvolution of sources with arbitrary distributions. In Proc. IEEE Workshop on Neural Networks for Signal Processing (pp. 436 444). New York: IEEE Press. Frankel, T. (1997). The geometry of physics. Cambridge: Cambridge University Press. Girolami, M., & Fyfe, C. (1997a). Extraction of independent signal sources using a deflationary exploratory projection pursuit network with lateral inhibition. I.E.E. Proceedings on Vision, Image and Signal Processing Journal, 14, 299 36. Girolami, M., & Fyfe, C. (1997b). Independence are far from normal. In Proceeding of European Symposium on Artificial Neural Networks (pp. 297 32). Hérault, J., & Jutten, C. (1986). Space or time adaptive signal processing by neural network models. In J. S. Denker (Ed.), Neural networks for computing: Proceedings of AIP Conference (pp. 26 211). New York: American Institute of Physics. Jutten, C., & Hérault, J. (1991). Blind separation of sources, part I: An adaptive algorithm based on neuromimetic architecture. Signal Processing, 24, 1 2. Kron, G. E. (1948). Geometrical theory of electrodynamics. New York: Wiley. Matsuoka, K., Ohya, M., & Kawamoto, M. (1995). A neural net for blind separation of non-stationary signals. Neural Networks, 8(3), 411 419.

1484 Shun-ichi Amari, Tian-Ping Chen, and Andrzej Cichocki Moreau, E., & Macchi, O. (1995). High order contrasts for self-adaptive source separation. International Journal of Adaptive Control and Signal Processing, 1(1), 19 46. Oja, E., & Karhunen, J. (1995). Signal separation by nonlinear Hebbian learning. In M. Palaniswami et al. (Eds.), Computational intelligence A dynamic system perspective (pp. 83 97). New York: IEEE Press. Schouten, J. A. (1954). Ricci calculus. Berlin: Springer-Verlag. Sorouchyari, E. (1991). Blind separation of sources, part III: Stability analysis. Signal Processing, 24, 21 3. Suzuki, T., & Nakamura, Y. (1995). Spiral motion of nonholonomic space robots. Journal of the Robotics Society of Japan, 13, 12 129. Whittaker, E. T. (194). Analytical dynamics. Proc.Roy. Soc., Edinburgh. Received September 15, 1997; accepted June 4, 1998.