Face Tracking and Recognition in Video (Face Recognition Techniques) Part 2

Results for Database-0

We now consider affine transformation. Specifically, the motion is characterized bytmp35b0-563_thumb[2]are deformation parameters and

tmp35b0-564_thumb[2]are 2D translation parameters. It is a reasonable approximation because there is no significant out-of-plane motion as the subjects walk toward the camera. Regarding the photometric transformation, only the zero-mean-unit-variance operation is performed to compensate partially for contrast variations. The complete transformation ,tmp35b0-565_thumb[2]is processed as follows. Affine transform z usingtmp35b0-566_thumb[2]crop out the interested region at positiontmp35b0-567_thumb[2]with the same size as the still template in the gallery, and perform the zero-mean-unit-variance operation.


A time-invariant first-order Markov Gaussian model with constant velocity is used for modeling motion transition. Given that the subject is walking toward the camera, the scale increases with time. However, under perspective projection, this increase is no longer linear, causing the constant-velocity model to be not optimal. However, experimental results show that so long as the samples of θ can cover the motion, this model is sufficient.

The likelihood measurement is simply set as a “truncated” Laplacian:

tmp35b0-573_thumb[2]

wheretmp35b0-574_thumb[2]is sum of absolute distance,tmp35b0-575_thumb[2]are  manually  specified, and tmp35b0-578_thumb[2]

 

Database-1. First row: the face gallery with image size of 30 x 26. Second and third rows: four frames in one probe video with image size of 720 x 480; the actual face size ranged from approximately 20 x 20 in the first frame to 60 x 60 in the last frame. Note the significant illumination variations between the probe and the gallery

Fig. 13.1 Database-1. First row: the face gallery with image size of 30 x 26. Second and third rows: four frames in one probe video with image size of 720 x 480; the actual face size ranged from approximately 20 x 20 in the first frame to 60 x 60 in the last frame. Note the significant illumination variations between the probe and the gallery

Gaussian distribution is widely used as a noise model, accounting for sensor noise and digitization noise among others. However, given the observation equation:tmp35b0-580_thumb[2] tmp35b0-581_thumb[2]the dominant part of vt becomes the high-frequency residual if θt is not proper; and it is well known that the high-frequency residual of natural images is more Laplacian-like. The “truncated” Laplacian is used to give a “surviving” chance for samples to accommodate abrupt motion changes.

Table 13.1 summarizes the average recognition performance and computational time of the Condensation and the proposed algorithm when applied to Database-0. Both algorithms achieved 100% recognition rate with top match. However, the proposed algorithm is more than 10 times faster than the CONDENSATION algorithm.

Table 13.1 Recognition performance of algorithms when applied to Database-0

Algorithm

Condensation

Proposed

Recognition rate within top one match

100%

100%

Time per frame

7 seconds

0.5 seconds

Table 13.2 Performances of algorithms when applied to Database-1

Case

Case 1

Case 2

Case 3

Case 4

Case 5

Tracking accuracy

83%

87%

93%

100%

NA

Recognition within top 1 match

13%

NA

83%

93%

57%

Recognition within top 3 matches

43%

NA

97%

100%

83%

Results on Database-1

Case 1: Tracking and Recognition Using Laplacian Density We first investigate the performance using the same setting as described in Sect. 13.3.3.1. Table 13.2 shows that the recognition rate is poor: only 13% are correctly identified using the top match. The main reason is that the “truncated” Laplacian density is not able to capture the appearance difference between the probe and the gallery, indicating a need for more effective appearance modeling. Nevertheless, the tracking accuracy is reasonable, with 83% successfully tracked because we are using multiple face templates in the gallery to track the specific face in the probe video. After all, faces in both the gallery and the probe belong to the same class of human face, and it seems that the appearance change is within the class range.

Case 2: Pure Tracking Using Laplacian Density In Case 2, we measure the appearance change within the probe video as well as the noise in the background. To this end, we introduce a dummy template T0, a cut version in the first frame of the video. Define the observation likelihood for tracking as

tmp35b0-584_thumb[2]

where σ2 and t2 are set manually. The other setting, such as motion parameter and model, is the same as in Case 1. We still can run the Condensation algorithm to perform pure tracking. Table 13.2 shows that 87% are successfully tracked by this simple tracking model, which implies that the appearance within the video remains similar.

Case 3: Tracking and Recognition Using Probabilistic Subspace Density As mentioned in Case 1, we need a new appearance model to improve the recognition accuracy. of the many approaches suggested in the literature, we decided to use the approach suggested by Moghaddam et al. [25] because of its computational efficiency and high recognition accuracy. However, here we model only the intrapersonal variations.

We need at least two facial images for one identity to construct the intrapersonal space (IPS). Apart from the available gallery, we crop out the second image from the video ensuring no overlap with the frames actually used in probe videos.

We then fit a probabilistic subspace density on top of the IPS. It proceeds as follows: A regular PCA is performed for the IPS. Suppose the eigensystem for the IPS istmp35b0-585_thumb[2]where d is the number of pixels andtmp35b0-586_thumb[2]Only top r principal components corresponding to top r eigenvalues are then kept while the residual components are considered isotropic. The density is written as follows tmp35b0-589_thumb[2]

where the principal components y, the reconstruction errortmp35b0-590_thumb[2]and the isotropic noise variance ρ are defined as

tmp35b0-592_thumb[2]

It is easy to write the likelihood as follows:

tmp35b0-593_thumb[2]

Table 13.2 lists the performance using this new likelihood measurement. It turns out that the performance is significantly better than in Case 1, with 93% tracked successfully and 83% correctly recognized within the top match. If we consider the top three matches, 97% are correctly identified.

Case 4: Tracking and Recognition Using Combined Density In Case 2, we studied appearance changes within a video sequence. In Case 3, we studied the appearance change between the gallery and the probe. In Case 4, we attempt to take advantage of both cases by introducing a combined likelihood defined as follows.

tmp35b0-594_thumb[2]

Again, all other settings are the same as in Case 1. We now obtain the best performance so far: no tracking error, 93% are correctly recognized as the first match, and no error in recognition when the top three matches are considered.

Case 5: Still-to-Still Face Recognition We also performed an experiment for still-to-still face recognition. We selected the probe video frames with the best frontal face view (i.e., biggest frontal view) and cropped out the facial region by normalizing with respect to the eye coordinates manually specified. It turns out that the recognition result is 57% correct for the top match and 83% for the top three matches. Clearly, Case 4 is the best among all.

Video Gallery vs. Video Probes

Here we describe a parametric model for appearance and dynamics to understand the manifold structures of these models, which are then used to devise joint appearance and dynamic based recognition algorithms.

Parametric Modelfor Appearance and Dynamic Variations

A wide variety of spatio-temporal data have often been modeled as realizations of dynamical models. Examples include dynamic textures [11], human joint angle trajectories [6] and silhouettes [37]. A well-known dynamical model for such time-series data is the autoregressive and moving average (ARMA) model. Linear dynamical systems represent a class of parametric models for time-series. A wide variety of time series data such as dynamic textures, human joint angle trajectories, shape sequences, video based face recognition etc., are frequently modeled as autoregressive and moving average (ARMA) models [1, 6, 11, 37]. Let f(t) be a sequence of features extracted from a video indexed by time t . The ARMA model parametrizes the evolution of the features f(t) using the following equations:

tmp35b0-595_thumb[2]

where,tmp35b0-596_thumb[2]is    the    hidden state vector,tmp35b0-597_thumb[2]the    transition    matrix    andtmp35b0-598_thumb[2] tmp35b0-599_thumb[2]the measurement matrix.tmp35b0-600_thumb[2]represents the observed features while w and v are noise components modeled as normal with 0 mean and covariancestmp35b0-601_thumb[2] tmp35b0-602_thumb[2]respectively.

For high-dimensional time-series data (dynamic textures etc), the most common approach is to first learn a lower-dimensional embedding of the observations via PCA, and learn temporal dynamics in the lower-dimensional space. Closed form solutions for learning the model parameters (A, C) from the feature sequence (fi:T) have been proposed by [11, 27] and are widely used in the computer vision community. Let observationstmp35b0-603_thumb[2]represent the features for the time indicestmp35b0-604_thumb[2]be    the    singular    value    decomposition of thetmp35b0-605_thumb[2]data. Thentmp35b0-606_thumb[2]where

tmp35b0-617_thumb[2]

The model parameters (A, C) do not lie in a vector space. The transition matrix A is only constrained to be stable with eigenvalues inside the unit circle. The observation matrix C is constrained to be an orthonormal matrix. For comparison of models, the most commonly used distance metric is based on subspace angles between column-spaces of the observability matrices [10]. For the ARMA model of (13.20), starting from an initial condition z(0), it can be shown that the expected observation sequence is given by

tmp35b0-618_thumb[2]

Thus, the expected observation sequence generated by a time-invariant model M = (A,C) lies in the column space of the extended observability matrix given by

tmp35b0-619_thumb[2]

In experimental implementations, we approximate the extended observability matrix by the finite observability matrix as is commonly done [33]

tmp35b0-620_thumb[2]

The size of this matrix is mp x d. The column space of this matrix is a d -dimensional subspace of Rmp, where d is the dimension of the state-space z in (13.20). d is typically of the order of 5-10.

Thus, given a database of videos, we estimate the model parameters as described above for each video. The finite observability matrix is computed as in (13.23). To represent the subspace spanned by the columns of this matrix, we store an orthonormal basis computed by Gram-Schmidt orthonormalization. Since, a subspace is a point on a Grassmann manifold [35, 36], a linear dynamical system can be alternately identified as a point on the Grassmann manifold corresponding to the column space of the observability matrix. The goal now is to devise methods for classification and recognition using these model parameters. Given a set of videos for a given class, we would like to compute a parametric or non-parametric class-conditional density. Then, the maximum likelihood classification for each test instance can be performed using these class conditional distributions. To enable these, we need to understand the geometry of the Grassmann manifold.

The Manifold Structure of Subspaces

The set of all d-dimensional linear subspaces of Rn is called the Grassmann manifold which will be denoted astmp35b0-621_thumb[2]The set of all n x d orthonormal matrices is called the Stiefel manifold and shall be denoted astmp35b0-623_thumb[2]As discussed in the applications above, we are interested in computing statistical models over the Grassmann manifold. Lettmp35b0-624_thumb[2]be    some    previously    estimated points ontmp35b0-625_thumb[2]and we seek their sample mean, an average, for defining a probability model on ,tmp35b0-626_thumb[2]Recall that these Uis are tall, orthogonal matrices. It is easy to see that the Euclidean sample meantmp35b0-627_thumb[2]is not a valid operation, because the resultant mean does not have the property of orthonormality. This is becausetmp35b0-628_thumb[2]is not a vector space. Similarly, many of the standard tools in estimation and modeling theory do not directly apply to such spaces but can be adapted by accounting for the underlying nonlinear geometry.

A subspace is stored as an orthonormal matrix which forms a basis for the subspace. As mentioned earlier, orthonormal matrices are points on the Stiefel manifold. However, since the choice of basis for a subspace is not unique, any notion of distance and statistics should be invariant to this choice. This requires us to interpret each point on the Grassmann manifold as an equivalence of points on the Stiefel manifold, where all orthonormal matrices that span the same subspace are considered equivalent. This interpretation is more formally described as a quotient interpretation that is, the Grassmann manifold is considered a quotient space of the Stiefel manifold. Quotient interpretations allow us to extend the results of the base manifold such as tangent spaces, geodesics etc to the new quotient manifold. In our case, it turns out that the Stiefel manifold itself can be interpreted as a quotient of a more basic manifold—the special orthogonal group SO(n). A quotient of Stiefel is thus a quotient of SO(n) as well.

A point U on ,tmp35b0-629_thumb[2]is represented as a tall-thin n x d orthonormal matrix. The corresponding equivalence class of n x d matricestmp35b0-630_thumb[2]is called the Procrustes representation of the Stiefel manifold. Thus, to compare two points intmp35b0-631_thumb[2]we simply compare the smallest squared distance between the corresponding equivalence classes on the Stiefel manifold according to the Procrustes representation. Given matricestmp35b0-632_thumb[2]the smallest squared Euclidean distance between the corresponding equivalence classes is given by

tmp35b0-643_thumb[2]

When R varies over the orthogonal group O(d), the minimum is attained at

tmp35b0-644_thumb[2]is the singular value decomposition of A. We refer the reader to [8] for proofs and alternate cases. Given several examples from a class (U1 ,U2 ,…,Un) on the manifold, the class conditional density can be estimated using an appropriate kernel function. We first assume that an appropriate choice of a divergence on the manifold has been made such as the one above. For the Procrustes measure, the density estimate is given by [8] as

tmp35b0-646_thumb[2]

where K(T) is the kernel function, M is a d x d positive definite matrix which plays the role of the kernel width or a smoothing parameter. C(M) is a normalizing factor chosen so that the estimated density integrates to unity. The matrix valued kernel function K(T) can be chosen in several ways. We have used K(T) = exp(-tr(T)) in all the experiments reported in this topic. In this non-parametric method for density estimation, the choice of kernel width M becomes important. Thus, though this is a non-iterative procedure, the optimal choice of the kernel width can have a large impact on the final results. In general, there is no standard way to choose this parameter except for cross-validation. In the experiments reported here, we use M = I ,the d x d identity matrix.

In addition to such nonparametric methods, there are principled methods to devise parametric densities on manifolds. Here, we simply refer the reader to [36] for mathematical details. In brief, using the tangent structure of the manifold, it is possible to define the well-known parametric densities such as multi-variate Gaussian, mixture-of-Gaussians etc., on the tangent spaces and wrap them back to the manifold. Densities defined in such a manner are called ‘wrapped’-densities. In the experiments section, we use a wrapped-Gaussian to model class-condition densities on the Grassmann manifold. This is compared to the simpler nonparametric method described above.

Video-Based Face Recognition Experiments

We performed a recognition experiment on the NIST’s Multiple Biometric Grand Challenge (MBGC) dataset. The MBGC Video Challenge dataset consists of a large number of subjects walking towards a camera in a variety of illumination conditions. Face regions are manually tracked and a sequence of cropped images is obtained. There were a total of 143 subjects with the number of videos per subject ranging from 1 to 5. In our experiments, we took subsets of the dataset which contained at least 2 sequences per person denoted as S2, at least 3 sequences per person denoted as S3 etc. Each of the face-images was first preprocessed to zero-mean and unity variance. In each of these subsets, we performed a leave-one-out testing. The results of the leave one out testing are shown in Table 13.3. Also reported are the total number of distinct subjects and the total number of video sequences in each of the subsets. In the comparisons, we show results using the ‘arc-length’ metric between subspaces [13]. This metric computes the subspace angles between two subspaces and takes the Frobenius norm of the angles as a distance measure [13]. We also show comparisons with the Procrustes measure, the Kernel density estimate with M = I and a parametric wrapped Gaussian density on the manifold. The wrapped Gaussian is estimated on the tangent-plane centered at the mean-point of the dataset. The mean, more formally defined as the Karcher mean, is defined as the point that minimizes the sum of squared geodesic distances to all other points. The tangent-plane being a vector space allows the use of multi-variate statistics to define class-conditional densities. We refer the reader to [36] for mathematical details.

Table 13.3 Comparison of video based face recognition approaches using (a) Subspace Angles + Arc-length metric, (b) Procrustes Distance, (c) kernel density, (d) Wrapped Normal on Tangent Plane

Subset

Distinct

Subjects

Total Sequences

Arc-length

Metric

Procrustes

Metric

Kernel

density

Wrapped

Normal

S2

143

395

38.48

43.79

39.74

63.79

S3

55

219

48.85

53.88

50.22

74.88

S4

54

216

48.61

53.70

50.46

75

Avg.

45.31%

50.45%

46.80%

71.22%

As can be seen, statistical methods outperform nearest-neighbor based approaches. As one would expect, the results improve when more examples per class are available. Since the optimal kernel-width is not known in advance, this might explain the relatively poor performance of the kernel density method. More examples of statistical inference on the Grassmann manifold for image and video-based recognition can be found in [35].

Face Recognition in Camera Network

Video-based face recognition algorithms exploit information temporally across the video sequence to improve recognition performance. With camera networks, we can capture multi-view videos which allow us to further integrate information spatially across view angles. It is worth noting that this is different from traditional face recognition of single-camera videos in which various face poses exhibit. In that case, one usually needs to model the dynamics of pose changes in the training phase and estimate pose in the testing phase. For example, in [20], Lee et al. train a representation for the face appearance manifold. The manifold consists of locally linear subspaces for different poses. A transition probability matrix is also trained to characterize the temporal dynamics for this representation. In [23], the dynamics are encoded in the learned Hidden Markov Models (HMMs). The mean observations of hidden states are shown to represent facial images at various poses. These approaches are designed to work with a single camera.

On the other hand, in camera network deployments there are multiple images of the face in different poses at a given time instant. These images could include a mix of frontal and nonfrontal images of the face, or, in some cases, a mix of nonfrontal images (see Fig. 13.2). Videos captured in such a mode have natural advantages in providing persistent sensing over a large area and stronger cues for handling pose variations. Nonetheless, if we do not leverage the collaboration among cameras, the power of multi-view data over single-views cannot be fully exploited. For example, if we extend the single-view video-based methods, such as [20] and [23], to a camera network, they have to function in such a mode that cameras do not collaborate with each other except at the final fusion stage.

Images acquired by a multi-camera network. Each column corresponds to a different camera, and each row corresponds to a different time instant and subject. Note that, under unconstrained acquisition, it is entirely possible that none of the images are frontal in spite of using five cameras to observe the subject [32]

Fig. 13.2 Images acquired by a multi-camera network. Each column corresponds to a different camera, and each row corresponds to a different time instant and subject. Note that, under unconstrained acquisition, it is entirely possible that none of the images are frontal in spite of using five cameras to observe the subject [32]

In general, there are some principles one should follow in developing a videobased face recognition algorithm for camera networks: First, the method should be able to collaboratively utilize information collected by multiple cameras and arrive at a multi-view representation from it, as opposed to perform recognition for each view individually and then fusing the result. Second, the method should be able to tackle pose variations effectively, as this is the major concern of a multi-view face recognition system. Third, the method should work on data whose acquisition conditions are as close to practical surveillance situations as possible. These conditions include: reasonable distance between subject and cameras, relatively low resolution in the face region, uncontrolled pose variations, uncontrolled subject motion, and possible interruptions in acquisition (say, the subject moves out of the field of view of a camera) etc.

Next, we will introduce a video-based face tracking and recognition framework following these principles. The system first tracks a subject’s head from multiview videos and back-projects textures to a spherical head-model. Then a rotation-invariant feature based on spherical harmonic (SH) transform is constructed from the texture maps. Finally, video-based recognition is achieved through measurement of ensemble similarity.

Next post:

Previous post: