You are on page 1of 115

1

Probabilistic Modeling for Semantic Scene Classification


Matthew R. Boutell
URCS Technical Report 862
May, 2005

This thesis was proposed in April, 2003. The dissertation of the thesis will
be published in May, 2005. While the dissertation subsumes and modifies much
of the material in this proposal, I have made it available (as URCS TR 862)
as a historical supplement to document the details and results of the MASSES
(Material and Spatial Experimental Scenes) prototype.
Probabilistic Modeling for Semantic
Scene Classification
by

Matthew R. Boutell

Thesis Proposal

for the Degree

Doctor of Philosophy

Supervised by

Christopher M. Brown

Department of Computer Science


The College
Arts and Sciences

University of Rochester
Rochester, New York

2005
ii

Abstract
Scene classification, the automatic categorization of images into semantic classes
such as beach, field, or party, is useful in applications such as content-based im-
age organization and context-sensitive digital enhancement. Most current scene-
classification systems use low-level features and pattern recognition techniques;
they achieve some success on limited domains.

Several contemporary classifiers, including some developed in Rochester, in-


corporate semantic material and object detectors. Classification performance im-
proves because because the gap between the features and the image semantics is
narrowed. We propose that spatial relationships between the objects or materials
can help by distinguishing between certain types of scenes and by mitigating the
effects of detector failures. While past work on spatial modeling has used logic-
or rule-based models, we propose a probabilistic framework to handle the loose
spatial relationships that exist in many scene types.

To this end, we have developed MASSES, an experimental testbed that can


generate virtual scenes. MASSES can be used to experiment with different spatial
models, different detector characteristics, and different learning parameters. Using
a tree-structured Bayesian network for inference on a series of simulated natural
scenes, we have shown that the presence of key materials can effectively distinguish
certain scene types. However, spatial relationships are needed to disambiguate
other types of scenes, achieving a gain of 7% in one case.

However, our simple Bayes net is not expressive enough to model the faulty
detection at the level of individual regions. As future work, we propose first to
iii

evaluate full (DAG) Bayesian networks and Markov Random Fields as potential
probabilistic frameworks. We then plan to extend the chosen framework for our
problem. Finally, we will compare our results on real and simulated sets of images
with those obtained by other systems using spatial features represented implicitly.
iv

Table of Contents

Abstract ii

List of Tables vii

List of Figures viii

1 Introduction 1

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 The Problem of Scene Classification . . . . . . . . . . . . . . . . . 3

1.2.1 Scene Classification vs. Full-scale Image Understanding . . 4

1.2.2 Scene Classification vs. Not Object Recognition . . . . . . 5

1.3 Past Work in Scene Classification . . . . . . . . . . . . . . . . . . 6

1.4 Statement of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.4.1 Philosophy . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.5 Summary of Preliminary Work . . . . . . . . . . . . . . . . . . . . 8

1.6 Organization of Proposal . . . . . . . . . . . . . . . . . . . . . . . 9

2 Related Work 10

2.1 Design Space of Scene Classification . . . . . . . . . . . . . . . . . 11

2.1.1 Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
v

2.1.2 Learning and Inference Engines . . . . . . . . . . . . . . . 11

2.2 Scene Classification Systems . . . . . . . . . . . . . . . . . . . . . 13

2.2.1 Low-level Features and Implicit Spatial Relationships . . . 13

2.2.2 Low-level Features and Explicit Spatial Relationships . . . 16

2.2.3 Mid-level Features and Implicit Spatial Relationships . . . 19

2.2.4 Semantic Features without Spatial Relationships . . . . . . 20

2.2.5 Semantic Features and Explicit Spatial Relationships . . . 21

2.2.6 Summary of Scene Classification Systems . . . . . . . . . . 22

2.3 Options for Computing Spatial Relationships . . . . . . . . . . . . 23

2.3.1 Computing Qualitative Spatial Relationships . . . . . . . . 23

2.3.2 Computing Quantitative Spatial Relationships . . . . . . . 26

2.4 Probabilistic Graphical Models . . . . . . . . . . . . . . . . . . . 27

2.4.1 Bayesian Networks . . . . . . . . . . . . . . . . . . . . . . 29

2.4.2 Markov Random Fields . . . . . . . . . . . . . . . . . . . . 33

2.4.3 Relative Merits . . . . . . . . . . . . . . . . . . . . . . . . 42

3 Methodology 44

3.1 Statistics of Natural Images . . . . . . . . . . . . . . . . . . . . . 45

3.1.1 Ground Truth Collection Process . . . . . . . . . . . . . . 46

3.1.2 Scene Prototypes . . . . . . . . . . . . . . . . . . . . . . . 48

3.1.3 Spatial Relationships . . . . . . . . . . . . . . . . . . . . . 48

3.1.4 Scene-specific Spatial Relationship Statistics . . . . . . . . 51

3.2 Experimental Environment . . . . . . . . . . . . . . . . . . . . . . 51

3.2.1 Advantages of Working in Simulation . . . . . . . . . . . . 52

3.2.2 MASSES Prototype . . . . . . . . . . . . . . . . . . . . . . 53


vi

3.2.3 Simulating Faulty Detectors . . . . . . . . . . . . . . . . . 58

3.2.4 Background Regions . . . . . . . . . . . . . . . . . . . . . 64

4 Experimental Results 66

4.1 Best-case Detection in MASSES . . . . . . . . . . . . . . . . . . . 66

4.2 Best-case Detection on Beach Photographs . . . . . . . . . . . . . 68

4.3 Faulty Detection on Beach Photographs . . . . . . . . . . . . . . 70

5 Proposed Research 73

5.1 Research Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

5.1.1 Why a New Inference Mechanism? . . . . . . . . . . . . . 76

5.1.2 Bayes Nets Vs. MRFs . . . . . . . . . . . . . . . . . . . . 77

5.1.3 Extend the Chosen Framework . . . . . . . . . . . . . . . 79

5.1.4 Analyze the Effect of Detector Quality . . . . . . . . . . . 82

5.1.5 Evaluate Explicit Spatial Relationships . . . . . . . . . . . 82

5.1.6 Evaluate Semantic Features . . . . . . . . . . . . . . . . . 83

5.1.7 Generalize to Other Scene Classes . . . . . . . . . . . . . . 83

5.1.8 Explore Potential Long-term Directions . . . . . . . . . . . 84

5.2 Issues Not Addressed in This Thesis . . . . . . . . . . . . . . . . . 85

5.3 Research Schedule . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

6 Acknowledgments 87

Bibliography 88

A Natural Scene Statistics 98

B Detector Characteristics 103


vii

List of Tables

2.1 Options for features to use in scene classification. . . . . . . . . . 12

2.2 Potential classifiers to use in scene classification. . . . . . . . . . . 14

2.3 Related work in scene classification, organized by feature type and


use of spatial information. . . . . . . . . . . . . . . . . . . . . . . 15

3.1 Scene class definitions. . . . . . . . . . . . . . . . . . . . . . . . . 47

3.2 Distribution resulting from offline sampling procedure. . . . . . . 64

4.1 MASSES with best-case material detection: Accuracy with and


without spatial information. . . . . . . . . . . . . . . . . . . . . . 67

4.2 MASSES with best-case material detection: Accuracy with and


without spatial information. . . . . . . . . . . . . . . . . . . . . . 69

4.3 MASSES with faulty material detection: Accuracy with and with-
out spatial information. . . . . . . . . . . . . . . . . . . . . . . . . 70
viii

List of Figures

1.1 Content-ignorant color balancing can destroys the brilliance of sun-


set images, such as those pictured, which have the same global color
distribution as indoor, incandescent-illuminated images. . . . . . . 3

2.1 A Bayes Net with a loop. . . . . . . . . . . . . . . . . . . . . . . . 32

2.2 Portion of a typical two-layer MRF. In low-level computer vision


problems, the top layer (black) represents the external evidence of
the observed image while the bottom layer (white) expresses the a
prioriknowledge about relationships between parts of the underly-
ing scene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.1 Ground-truth labeling of a beach scene. Sky, water, and sand re-
gions are clearly shown. . . . . . . . . . . . . . . . . . . . . . . . 46

3.2 Prototypical beach scenes. (a) A simple beach scene without back-
ground objects. (b) Because we make no attempt to detect it, we
consider the sailboat to be “background”. (c) A more complicated
scene: a developed beachfront. (d) A scene from a more distant
field-of-view. (e) A crowded beach. . . . . . . . . . . . . . . . . . 49
ix

3.3 Prototypical urban scenes. (a) The most common urban scene,
containing sky, buildings, and roads. (b),(c) The sky is not simply
above the buildings in these images. (d) Roads are not necessary.
(e) Perspective views induce varied spatial relationships. (f) Close
views can preclude the presence of sky. . . . . . . . . . . . . . . . 50

3.4 An example of spatial relationships involving a split region. . . . . 50

3.5 The MASSES environment. Statistics from labeled scenes are used
to bootstrap the generative model, which can then produce new
virtual scenes for training or testing the inference module. . . . . 54

3.6 Single-level Bayesian network used for MASSES . . . . . . . . . . 54

3.7 Sampling the scene type yields class C. Then we sample to find
the materials present in the image, in this case, M1 , M3 , and M4 .
Finally, we sample to find the relationships between each pair of
these material regions. . . . . . . . . . . . . . . . . . . . . . . . . 57

3.8 Bayesian network subgraph showing relationship between regions


and detectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.1 Images incorrectly classified due to spatial relationships. . . . . . 69

4.2 Images incorrectly classified using faulty detectors and no spatial


relationships. The actual materials are shown the top row; the
detected materials are shown below each. . . . . . . . . . . . . . . 70

4.3 Images incorrectly classified using faulty detectors. . . . . . . . . 71

4.4 Images incorrectly classified using faulty detectors. . . . . . . . . 72

5.1 Classifier input (labeled regions) and output (classification and con-
fidence). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.2 An image in which a region is mis-detected, creating contradictory


spatial relationships in the material-based inference scheme. . . . 76
x

5.3 Proposed DAG Bayesian network. Note that separate material and
region layers are needed. . . . . . . . . . . . . . . . . . . . . . . . 77
1

1 Introduction

Semantic scene classification, the process of categorizing images into semantic


classes such as beaches, sunsets or parties is a useful endeavor. As humans, we
can quickly determine the classification of a scene, even without recognizing every
one of the details present. Even the gist of a scene is worth much in terms of
communication.

1.1 Motivation

Automatic semantic classification of digital images finds many applications. We


describe two major ones briefly: content-based image organization and retrieval
(CBIR) and digital enhancement.

With digital libraries growing in size so quickly, accurate and efficient tech-
niques for CBIR become more and more important. Many current systems allow a
user to specify an image and then search for images “similar” to it, where similarity
is often defined only by color or texture properties. Because a score is computed
on each image in the potentially-large database, it is somewhat inefficient (though
individual calculations vary in complexity).

Furthermore, this so-called “query by example” has often proven to be return


2

inadequate results [68]. Sometimes the match between the retrieved and the query
images is hard to understand, while other times, the match is understandable, but
contains no semantic value. For instance, with simple color features, a query for a
rose can return a picture of a man wearing a red shirt, especially if the background
colors are similar as well.

Knowledge of the semantic category of a scene helps narrow the search space
dramatically [37]. If the categories of the query image and the database images
have been assigned (either manually or by an algorithm), they can be exploited
to improve both efficiency and accuracy. For example, knowing what constitutes
a party scene allows us to consider only potential party scenes in our search and
thus helps to answer the query “find photos of Mary’s birthday party”. This way,
search time is reduced, the hit rate is higher, and the false alarm rate is expected
to be lower. Visual examples can be found in [76].

Knowledge about the scene category can find also application in digital en-
hancement [73]. Digital photofinishing processes involve three steps: digitizing
the image if necessary (if the original source was film), applying enhancement
algorithms, and outputting the image in either hardcopy or electronic form. En-
hancement consists primarily of color balancing, exposure enhancement, and noise
reduction. Currently, enhancement is generic (i.e. without knowledge of the scene
content). Unfortunately, while a balancing algorithm might enhance the quality
of some classes of pictures, it degrades others.

Take color balancing as an example. Photographs captured under incandescent


lighting without flash tend to be yellowish in color. Color balancing removes
the yellow cast. However, applying the same color balancing to a sunset image
(containing the same overall yellow color) destroys the desired brilliance (Figure
1.1).

Other images that are negatively affected by color balancing are those con-
taining skin-type colors. Correctly balanced skin colors are important to human
3

Figure 1.1: Content-ignorant color balancing can destroys the brilliance of sunset
images, such as those pictured, which have the same global color distribution as
indoor, incandescent-illuminated images.

perception [64], and it is important to balance them. However, causing non-skin


objects with similar colors to look like skin is a conspicuous error.

Rather than applying generic color balancing and exposure adjustment to all
images, knowledge of the scene’s semantic classification allows us to customize
them to the scene. Following the example above, we could retain or boost sun-
set scenes’ brilliant colors while reducing a tungsten-illuminated indoor scene’s
yellowish cast.

1.2 The Problem of Scene Classification

On one hand, isn’t scene classification preceded by image understanding, the “holy
grail” of vision? What makes us think we can achieve results? On the other hand,
isn’t scene classification just an extension of object recognition, for which many
techniques have been proposed with varying success? How is scene classification
different from these two related fields?
4

1.2.1 Scene Classification vs. Full-scale Image Understand-


ing

As usually defined, image understanding is the process of converting “pixels to


predicates”: (iconic) image representations to another (symbolic) form of knowl-
edge [2]. Image understanding is the highest (most abstract) processing level in
computer vision [71], as opposed to image processing techniques, which convert
one image representation to another. (For instance, using a mask to convert raw
pixels to an edge image is much more concrete than identifying the expression
on a person’s face in the image!) Lower-level image processing techniques such
as segmentation are used to create regions that can then be identified as objects.
Various control strategies are used to order the processing steps and can vary [3].
The end result desired is for the vision to support high-level reasoning about the
objects and their relationships to meet a goal.

While image understanding in unconstrained environments is still very much


an open problem [71; 77], much progress is currently being made in scene clas-
sification. Because scenes can often be classified without full knowledge of every
object in the image, the goal is not as ambitious. For instance, if a person recog-
nizes trees at the top of a photo, grass on the bottom, and people in the middle,
he may hypothesize that he is looking at a park scene, even if he cannot see every
detail in the image. Or on a different level, if there are lots of sharp vertical and
horizontal edges, he may be looking at an urban scene.

It may be possible in some cases to use low-level information, such as color


or texture, to classify some scene types accurately. In other cases, perhaps ob-
ject recognition is necessary, but not necessarily of every object in the scene.
In general, classification seems to be an easier problem than unconstrained im-
age understanding; early results have confirmed this for certain scene types in
constrained environments [74; 77]. Scene classification is a subset of the image
5

understanding problem, and can be used to ease other image understanding tasks
[75]. For example, knowing that a scene is of a beach constrains where in the
scene one should look for people.

Obtaining image understanding in unconstrained environments is a lofty goal,


and one worthy of pursuit. However, given the state of image understanding, we
see semantic scene classification as a necessary stepping-stone in pursuit of the
“grail”.

1.2.2 Scene Classification vs. Not Object Recognition

However, scene classification is a different beast than object recognition. Detection


of rigid objects can rely upon geometrical relationships within the objects, and
various techniques [21; 63] can be used to achieve invariance to affine transforms
and changes in scene luminance. Detection of non-rigid objects is less constrained
physically, since the relationships are looser [12]. Scene classification is even less
constrained, since the components of a scene are varied. For instance, while
humans might find it easy to recognize a scene of a child’s birthday party, the
objects and people that populate the scene can vary widely, and the cues that
determine the birthday scene class (such as special decorations, articles marked
with the age of the child, and facial expressions on the attenders) can be subtle.
Even the more obvious cues, like a birthday cake, may be difficult to determine.

Again, the areas of scene classification and object recognition are related;
knowing the identity of some of the scene’s objects will certainly help to classify the
scene, while knowing the scene type affects the expected likelihood and location
of the objects it contains.
6

1.3 Past Work in Scene Classification

Most of the current systems primarily use low-level features to classify scenes
and achieve some success on constrained problems. These systems tend to be
exemplar-based, in which features are extracted from images, and pattern recog-
nition techniques are used to learn the statistics of a training set and to classify
novel test images. Very few systems are model-based, in which the expected
configuration of the objects in the scenes is specified by a human expert.

1.4 Statement of Thesis

The limited success of scene classification systems using low-level features forces us
to look for other solutions. Currently, good semantic material detectors and object
recognizers are available [70; 38; 63] and have begun to be successfully applied to
scene classification [37]. However, the presence or absence of certain objects is
not always enough to determine a scene type. Furthermore, object detection is
still developing and is far from perfect. Faulty detection causes brittle rule-based
systems to break.

Our central claim is that Spatial modeling of semantic objects and materials
can be used to disambiguate certain scene types as well as mitigate the effects of
faulty detectors. Furthermore, an appropriate probabilistic inference mechanism
must be developed to handle the loose spatial structure found in real images.

Current research into spatial modeling relies on (fuzzy) logic and subgraph
matching [44; 83]. While we have found no research that incorporates spatial
modeling in a probabilistic framework, we argue that a probabilistic approach
would be more appropriate. First, logic (even fuzzy variants) is not equipped to
handle exceptions efficiently [50], a concern we address in more detail in Section
2.4. Second, semantic material detectors often yield belief in the materials. While
7

it is not obvious how to use belief values, it seems desirable to exploit the uncer-
tainty in calculating the overall belief in each scene type. A nice side effect of true
belief-based calculation is the ease in which a “don’t know” option can be added
to the classifier: simply threshold the final belief value.

The first interesting problem is the appropriate choice of a probabilistic frame-


work. Both Bayesian networks [66] and Markov Random Fields [30; 16; 15] have
been applied to other vision problems in the past.

We also propose to investigate the effects of spatial modeling. In our experi-


mentation, we plan to compare the following:

1) Baseline (no spatial relationships). Use naive Bayes classification rules using
the presence or absence of materials only.

2a) Qualitative spatial relationships. Incorporate relations such as above or


beside between regions. This would be appropriate for use in a Bayesian Network
framework.

2b) Quantitative spatial relationships. Use distance and direction between


regions. This may potentially be more accurate, due to the increase in the infor-
mation used, but requires more training data. These may work particularly well
within a Markov Random Field framework.

One of our major foreseen contributions will be to validate the hypothesized


gain due to spatial modeling.

1.4.1 Philosophy

The success of our approach seems to hinge on the strength of the underlying
detectors. Consider two scenarios. First, if the detectors are reasonably accurate,
then we can expect to overcome some faults using spatial relationships. However,
if they are extremely weak, we would be left with a very ambitious goal: from a
very pessimistic view of the world (loose spatial structure and weak detectors),
8

pull order from the chaos and accurately apply a discrete label to a configuration
of objects.

In this latter case, prior research seems to confirm that the task does not sound
promising. For instance, Selinger found that if an object could be recognized
with moderate success using a single camera view, additional views could improve
recognition substantially. However, if the single view gave weak detection, then
multiple views could not redeem the problem. She states [62] (p. 106):

The result is in concert with general expertise in the field of recog-


nition concerning the difficulty of leveraging multiple sources of weak
evidence into strong hypotheses.

Therefore, while we cannot expect to use our technique to classify scenes for
which the detectors are completely inaccurate, we stand a reasonable chance if
improving accuracy if the detectors are reasonably strong themselves.

1.5 Summary of Preliminary Work

We have performed our experiments in a simulated abstract world. The materials


and spatial relationships used are based on statistics captured from real scenes.
This provides us with a rich test bed, in which we can develop algorithms, compare
approaches, quickly experiment with parameters, and explore “what-if” situations.

With a prototype scene simulator we developed, using a single-level, tree-


structured Bayesian network for inference on a series of simulated natural scenes,
we have shown that the presence of key materials can effectively distinguish certain
scene types. However, spatial relationships are needed to disambiguate other types
of scenes, achieving a gain of 7% in one case.
9

However, when simulating faulty detectors, we found that the network is not
expressive enough to capture the necessary information, actually leading to lower
accuracy when spatial relationships were used.

1.6 Organization of Proposal

Chapter 2 contains an overview of the relevant literature in scene classification,


spatial modeling, and probabilistic frameworks. In Chapter 3, we describe our
methodology, both for the detector and the simulator. Chapter 4 is a summary of
our experiments and results (using best-case and faulty detectors). We conclude
in Chapter 5, in which we describe our research plan and proposed contributions.
10

2 Related Work

Scene classification is a young, emerging field. The first section of this chapter
is taken in large part from our earlier review of the state of the art in scene
classification [6]; because this thesis is a work in progress, there is much overlap
between the two. Here we focus our attention on systems using approaches directly
related to our proposed thesis. Readers desiring a more comprehensive survey or
more detail are referred to the original review.

All systems classifying scenes must extract appropriate features and use some
sort of learning or inference engine to classify the image. We start by outlining
the options available for features and classifiers. We then present a number of
systems which we have deemed to be good representations of the field.

We augment our review of the literature by discussing two computational


models of spatial relationships and then discussing in detail two graphical models
we could use for probabilistic inference: Bayesian Networks and Markov Random
Fields, each of which will be explored in the thesis.
11

2.1 Design Space of Scene Classification

The literature reveals two approaches to scene classification: exemplar-based and


model-based. On one hand, exemplar-based approaches use pattern recognition
techniques on vectors of low-level image features (such as color, texture, or edges)
or semantic features (such as sky, faces or grass). The exemplars are thought
to fall into clusters, which can then be used to classify novel test images, using
an appropriate distance metric. Most systems use an exemplar-based approach,
perhaps due to recent advances in pattern recognition techniques. On the other
hand, model-based approaches are designed using expert knowledge of the scene
such as the expected configuration of a scene. A scene’s configuration is the layout
(relative location and sizes) of its objects. While it seems as though this should
be very important, very little research has been done in this area.

In either case, appropriate features must be extracted for accurate classifica-


tion. What makes a certain feature appropriate for a given task? For pattern
classification, one wants the inter-class distance to maximized and the intra-class
distances to be minimized. Many choices are intuitive, e.g. edge features should
help separate city and landscape scenes [78].

2.1.1 Features

In our review [6], we described features we found in similar systems, or which


we thought could be potentially useful. Table 2.1 is a summary of that set of
descriptions.

2.1.2 Learning and Inference Engines

Pattern recognition systems classify samples represented by feature vectors (see a


good review in [28]). Features are extracted from each of a set of training images,
12

Table 2.1: Options for features to use in scene classification.


Feature Description
Color Histograms [72], Coherence vectors [49], Moments [79]
Texture [51] Wavelets[42; 65], MSAR [73], Fractal dimension [71]
Filter Output Fourier & discrete cosine transforms [46; 73; 74; 75; 77],
Gabor [59], Spatio-temporal [53]
Edges Direction histograms [77], Direction coherence vectors
Context Patch Dominant edge with neighboring edges [63]
Object Geometry Area, Eccentricity, Orientation [10]
Object Detection Output from belief-based material detectors[37; 66],
rigid object detectors [63], face detectors [60]
IU Output Output of other image understanding systems
e.g., Main Subject Detection [66]
Context Within images (scale, focus, pose) [75]
Between images (adjacent images on film or video)
Mid-level “Spatial envelope” features [47]
Meta-data Time-stamp, Flash firing, Focal length, text [36; 24]
Statistical Measures Dimensionality reduction [18; 58]
13

or exemplars. In most classifiers, a statistical inference engine then extracts infor-


mation from the processed training data. Finally, to classify a novel test image,
this type of system extracts the same features from the test image and compares
them to those in the training set [18]. This exemplar-based approach is used by
most of the current systems.

The classifiers used in these type of systems differ in how they extract infor-
mation from the training data. In Table 2.2, we present a summary of the major
systems used in the realm of scene classification.

2.2 Scene Classification Systems

As stated, many of the systems proposed in the literature for scene classification
are exemplar-based, but a few are model-based, relying on expert knowledge to
model scene types, usually in terms of the expected configuration of objects in
the scene. In this section, we describe briefly some of these systems and point
out some of their limitations en route to differentiating our proposed method. We
organize the systems by feature type and in the use of spatial information, as
shown in Table 2.3. Features are grouped into low-level, mid-level, and high-level
(semantic) features, while spatial information is grouped into those that model
the spatial relationships explicitly in the inference stage and those that do not.

2.2.1 Low-level Features and Implicit Spatial Relationships

A number of researchers have used low-level features sampled at regular spatial


locations (e.g. blocks in a rectangular grid). In this way, spatial features are
encoded implicitly, since the features computed on each location are mapped to
fixed dimensions in the feature space.
14

Table 2.2: Potential classifiers to use in scene classification.

Classifier Description
1-Nearest-Neighbor Classifies test sample with same class as the exemplar
(1NN) closest to it in the feature space.
K-Nearest-Neighbor Generalization of 1NN in which the sample is given
(kNN) [18] the label of the majority of the k closest exemplars.
Learning Vector A representative set of exemplars, called a codebook,
Quantization (LVQ) is extracted. The codebook size and learning rate
[31; 32] must be chosen in advance.
Maximum a Posteriori Combines the class likelihoods (which must be
(MAP) [77] modeled, e.g., with a mixture of Gaussians) with
class priors using Bayes rule.
Support Vector Find an optimal hyperplane separating two classes.
Machine (SVM) Maps data into higher dimensions, using a kernel
[8; 61] function,to increase separability. The kernel and
associated parameters must be chosen in advance.
Artificial Neural Function approximators in which the inputs are
Networks (ANN) [1] mapped, through a series of linear combinations
and non-linear activation functions to outputs.
The weights are learned using a technique
called backpropagation.
15

Table 2.3: Related work in scene classification, organized by feature type and use
of spatial information.
Spatial Information
Feature Type Implicit/None Explicit
Low-level Vailaya, et al. Lipson, et al.
Oliva, et al. Ratan & Grimson
Szummer & Picard Smith & Li
Serrano, et al.
Paek & Chang
Carson, et al.
Wang, et al.
Mid-level Oliva, et al.
High-level Luo, et al. Mulhem, et al.
(Semantic) Song & Zhang Proposed Method

The problems addressed include indoor vs. outdoor classification (Szummer


and Picard [73], Paek and Chang [48], and Serrano et al. [65]), outdoor scene
classification (Vailaya et al. [77]), and image orientation detection [79; 80] 1 .

The indoor vs. outdoor classifiers’ accuracy approaches 90% on tough (e.g.,
consumer) image sets. On the outdoor scene classification problem, mid-90%
accuracy is reported. This may be due to the use of constrained data sets (e.g.
from the Corel stock photo library), because on less constrained (e.g., consumer)
image sets, we found the results to be lower. The generalizability of the technique
is also called into question by the discrepancies in the numbers reported for image
orientation detection by some of the same researchers [79; 80].
1
While image orientation detection is a different level of semantic classification, many of the
techniques used are similar.
16

Pseudo-Object Features

The Blobworld system, developed at Berkeley, was developed primarily for content-
based indexing and retrieval. However, it is used for scene classification problem
in [9]. The researchers segment the image and use statistics computed for each re-
gion (e.g., color, texture, location with respect to a 3 × 3 grid) without performing
object recognition. Admittedly, this is a more general approach for scene types
containing no recognizable objects. However, we can hope for more using object
recognition. Finally, a maximum likelihood classifier performs the classification.

Wang’s SIMPLIcity (Semantics-sensitive Integrated Matching for Picture LI-


braries) system [80] also uses segmentation to match pseudo-objects. The system
uses a fuzzy method called “Integrated Region Matching” to effectively compen-
sate for potentially poor segmentation, allowing a region in one image to match
with several in another image. However, spatial relationships between regions are
not used and the framework is used only for CBIR, not scene classification.

2.2.2 Low-level Features and Explicit Spatial Relationships

The systems above either ignore spatial information or encode it implicitly using
a feature vector. However, other bodies of research imply that explicitly-encoded
spatial information is valuable and should be encoded explicitly and used by the
inference engine. In this section, we review this body of research, describing a
number of systems using spatial information to model the expected configuration
of the scene.

Configural Recognition

Lipson, Grimson, and Sinha at MIT use an approach they call “configural recog-
nition” [34; 35], using relative spatial and color relationships between pixels in low
resolution images to match the images with class models.
17

The specific features extracted are very simple. The image is smoothed and
subsampled at a low resolution (ranging from 8 × 8 to 32 × 32). Each pixel
represents the average color of a block in the original image; no segmentation
is performed. For each pixel, only its luminance, RGB values, and position are
extracted.

The hand-crafted models are also extremely simple. For example, a template
for a snowy mountain image is a blue region over a white region over a dark
region; one for a field image is a large bluish region over a large greener region.
In general, the model contains relative x- and y-coordinates, relative R-, G-, B-,
and luminance values, and relative sizes of regions in the image.

The matching process uses the relative values of the colors in an attempt to
achieve illumination invariance. Furthermore, using relative positions mimics the
performance of a deformable template: as the model is compared to the image,
the model can be deformed by moving the patch around so that it best matches
the image. A model-image match occurs if any one configuration of the model
matches the image. However, this criterion may be extended to include the degree
of deformation and multiple matches depending on how well the model is expected
to match the scene.

Classification is binary for each classifier. On a test set containing 700 pro-
fessional images (the Corel Fields, Sunsets and Sunrises, Glaciers and Mountains,
Coasts, California Coasts, Waterfalls, and Lakes and Rivers CDs), the authors
report recall using four classifiers: fields (80%), snowy mountains (75%), snowy
mountains with lakes (67%), and waterfalls (33%). Unfortunately, exact precision
numbers cannot be calculated from the results given.

The strength of the system lies in the flexibility of the template, in terms of
both luminance and position. However, one limitation the authors state is that
each class model captured only a narrow band of images within the class and that
multiple models were needed to span a class.
18

Learning the Model Parameters In a follow-up study by Ratan and Grimson


[54], they also used the same model, but learned the model parameters from
exemplars. They reported similar results to the hand-crafted models used by
Lipson. However, the method was computationally expensive [83].

Combining Configurations with Statistical Learning In another varia-


tion on the previous research, Yu and Grimson adapt the configural approach to
a statistical, feature-vector based approach, treating configurations like words ap-
pearing in a document [83]. Set representations, e.g. attributed graphs, contain
parts and relations. In this framework, the configurations of relative brightness,
positions, and sizes are subgraphs. However, inference is computationally costly.

Vector representations allow for efficient learning of visual concepts (using


the rich theory of supervised learning). Encoding configural information in the
features overcomes the limited ability of vector representations to preserve relevant
information about spatial layout. [83]

Within a CBIR framework with two query images, configurations as extracted


as follows. Because configurations contained in both images are most informative,
an extension of the maximum clique method is used to extract common subgraphs
from the two images. The essence of the method is that configurations are grown
from the best matching pairs (e.g., highly contrasting regions) in each image.

During the query process, the common configurations are broken into smaller
parts and converted to a vector format, in which feature i corresponds to the
probability that sub-configuration i is present in the image.

A naive (i.e., single-level, tree-structured) Bayesian network is trained on-line


for image retrieval. A set of query images is used for training, with likelihood
parameters estimated by EM. Database images are then retrieved in order of
their posterior probability.

On a subset of 1000 Corel images, a single waterfall query is shown to have


19

better retrieval performance than other measures such as color histograms, wavelet
coefficients, and Gabor filter outputs.

Note that the spatial information is explicitly encoded in the features, but is
used directly in the inference process.

In the subgraph extraction process above, if extracting a common configuration


from more than two images is desired, one can use Hong, et al.’s method [26].

Composite Region Templates (CRT)

CRTs are configurations of segmented image regions [69]. The configurations are
limited to those occurring in the vertical direction: each vertical column is stored
as a region string and statistics are computed for various sequences occurring in
the strings. While an interesting approach, one unfortunate limitation of their
experimental work is that the size of the training and testing sets were both
extremely limited.

2.2.3 Mid-level Features and Implicit Spatial Relationships

Oliva and Torralba [46; 47] propose what they call a “scene-centered” description
of images. They use an underlying framework of low-level features (multiscale
Gabor filters), coupled with supervised learning to estimate the “spatial envelope”
properties of a scene. They classify images with respect to “verticalness” (vertical
vs. horizontal), “naturalness” (vs. man-made), “openness” (presence of a horizon
line), “roughness” (fractal complexity), “busyness” (sense of clutter in man-made
scenes), “expansion” (perspective in man-made scenes), “ruggedness” (deviation
from the horizon in natural scenes), and “depth range”.

Images are then projected into this 8-dimensional space in which the dimen-
sions correspond to the spatial envelope features. They measure their success first
on individual dimensions through a ranking experiment. They then claim that
20

their features are highly correlated with the semantic categories of the images
(e.g., “highway” scenes are “open” and exhibit high “expansion”), demonstrating
some success on their set of images. It is unclear how their results generalize.

One observation they make is that their scene-centered approach is comple-


mentary to an “object-centered” approach like ours.

2.2.4 Semantic Features without Spatial Relationships

Semantic Features for Indoor Vs. Outdoor Classification

Luo and Savakis extended the method of [65] by incorporating semantic mate-
rial detection [37]. A Bayesian Network was trained for inference, with evidence
coming from low-level (color, texture) features and semantic (sky, grass) features.
Detected semantic features (which are not completely accurate) produced a gain
of over 2% and “best-case” (100% accurate) semantics gave a gain of almost 8%
over low-level features alone. The network used conditional probabilities of the
form P (sky present|outdoor). While this work showed the advantage of using
semantic material detection for certain types of scene classification, it stopped
short of using spatial relationships.

Semantic Features for Image Retrieval

Song and Zhang investigate the use of semantic features within the context of
image retrieval [70]. Their results are impressive, showing that semantic features
greatly outperform typical low-level features, including color histograms, color
coherence vectors, and wavelet texture for retrieval.

They use the illumination topology of images (using a variant of contour trees)
to identify image regions and combine this with other features to classify the
regions into the semantic categories such as sky, water, trees, waves, placid water,
lawn, and snow.
21

While they do not apply their work directly to scene classification, their success
with semantic features confirms our hypothesis that they help bridge the semantic
gap between pixel-representations and high-level understanding.

2.2.5 Semantic Features and Explicit Spatial Relationships

Mulhem, Leow, and Lee [44] present a novel variation of fuzzy conceptual graphs
for use in scene classification. Conceptual graphs are used for representing knowl-
edge in logic-based applications, since they can be converted to expressions of
first-order logic. Fuzzy conceptual graphs extend this by adding a method of
handling uncertainty.

A fuzzy conceptual graph is composed of three elements: a set of concepts


(e.g., ”mountain” or ”tree”), a set of relations (e.g., ”smaller than” or ”above”),
and a set of relation attributes (e.g., ”ratio” of two sizes). Any of these elements
which contain multiple possibilities is called fuzzy, while one which does not is
called crisp.

Model graphs for prototypical scenes are hand-crafted, and contain crisp con-
cepts and fuzzy relations and attributes. For example, a ”mountain-over-lake”
scene must contain a mountain and water, but the spatial relations are not guar-
anteed to hold. A fuzzy relation such as ”smaller than” may hold most of the
time, but not always.

Image graphs contain fuzzy concepts and crisp relations and attributes. This
is intuitive: while a material detector calculates the boundaries of objects and
can therefore calculate relations (e.g. ”to the left of”) between them, they can be
uncertain as to the actual classification of the material (consider the difficulty of
distinguishing between cloudy sky and snow, or of rock and sand). The ability to
handle uncertainty on the part of the material detectors is an advantage of this
framework.
22

Two subgraphs are matched using graph projection, a mapping such that each
part of a subgraph of the model graph exists in the image graph, and a metric
for linearly combining the strength of match between concepts, relations, and
attributes. A subgraph isomorphism algorithm is used to find the subgraph of the
model that matches best the image.

The basic idea of the algorithms is to decompose the model and image into
arches (two concepts connected by a relation), seed a subgraph with the best
matching pair of arches, and incrementally add other model arches that match
well.

They found that the image matching metric worked well on a small database of
two hundred images and four scene models (of mountain/lake scenes) generated by
hand. Fuzzy classification of materials was done using color histograms and Gabor
texture features. The method of generating confidence levels of the classification
is not specified.

While the results look promising for mountain/lake scenes, it remains to be


seen how well this approach will scale to a larger number of scene types.

2.2.6 Summary of Scene Classification Systems

Referring back to the summary of prior work in semantic scene classification given
in Table 2.3, we see that our work is closest to that of Mulhem, et al., but differs in
one key aspect: while theirs is logic-based, our proposed method is founded upon
probability theory, leading to principled methods of handling variability in scene
configurations. Our proposed method also learns the model parameters from a
set of training data, while theirs are fixed.
23

2.3 Options for Computing Spatial Relationships

If we are to utilize spatial relationships between regions, we need a method for


computing and encoding these relationships. We discuss both qualitative and
quantitative spatial relationships.

2.3.1 Computing Qualitative Spatial Relationships

We start by considering three models of computing qualitative spatial relation-


ships: Attentional Vector Sum, a biologically-inspired model; Weighted Walk-
throughs, a model developed for occluded regions; and a hybrid model produced
for efficient computation.

Attentional Vector Sum

Regier and Carlson [55] propose a computational model of spatial relations based
on human perception. They consider up, down, left, and right, to be symmetric,
and so focus their work on the above relation.

They call the reference object the landmark and the located object the tra-
jector. For example, ”the ball (trajector) is above the table (landmark)”. The
model is designed to handle 2D landmarks, but only point trajectors. However,
the researchers state that they are in the process of extending the model.

Four models are compared:

1. Bounding box (BB). A is above B if it is higher than the landmark’s highest


point and between its leftmost and rightmost points. The strength of the
match varies depending on the height and how centered it is above the ob-
ject; three parameters govern how quickly the match drops off, via sigmoidal
functions.
24

2. Proximal and Center-of-Mass (PC). Here, the projection is defined based


on the angle formed by the y-axis and the line connecting the trajector to
the landmark. Connecting it to the closest point on the landmark gives
the proximal angle and to the centroid gives the center-of-mass angle. This
model has four parameters: a gain, slope, and y-intercept of the piecewise
function for the goodness of the angle, and the relative weight given to the
components corresponding to the two angles.

3. Hybrid Model (PC-BB). This model extends the PC model by adding the
BB model’s height term. The height term gives the presence of a ”grazing
line” at the top of the landmark, an effect that was observed experimentally.
The model has four parameters: the slope, y-intercept, and relative weight
of the PC model plus the gain on the height function’s sigmoid.

4. Attentional Vector Sum (AVS). This model incorporates two human percep-
tual elements:

• Attention. Visual search for a target in a field of distractors is slow


when targets differ from distractors only in the spatial relation among
their elements (i.e. they do not exhibit ”pop-up”). Therefore, they
require attention.

• Vector sum. Studies of orientation cells in the monkey cortex show


that directions were modeled by a vector sum of the cells.

In the AVS model, the angle between the landmark and the trajector is
calculated as the weighted sum of angles between the points in the landmark
area and the trajector. The weights in the sum are related to attention. The
center of attention on the landmark is the point closest to the trajector; its
angle receives the most weight. As the landmark points get further from
the center of attention, they are weighted less, dropping off exponentially.
25

Lastly, the BB model’s height function is used again (for which they can
give no physiological or perceptual basis, but only because it was observed
experimentally). The model has four parameters: the width of the beam of
attention, the slope and y-intercept of the linear function relating angle to
match strength (as in the PC model) and the gain on the height function’s
sigmoid.

Optimal parameters for each model were found by fitting the model with an-
other researcher’s data set. A series of experiments was then performed to dis-
tinguish between the models. The AVS model fit each experiment’s data most
consistently.

The AVS method also gives a measure of ”how above” one region is compared
to another. This measure may potentially be used to our advantage.

However, as given, AVS may be too computationally expensive. Where ni


is # points in each region i, finding points on perimeter of each is O(n1 + n2 ),
giving p perimeter points. Finding closest point between landmark and trajector,
yields O(p) distances. Integrating over each region yields O(n1 × n2) distances.
However, if using a larger step size would not substantially reduce accuracy, we
could reduce the computation significantly.

Weighted Walkthroughs

Berretti et al. [5] developed a technique named ”weighted walkthroughs” to cal-


culate spatial relations. The method is designed to compare segmented regions
created by color backpropagation, and therefore has the advantage of handling
landmarks or trajectors that are made of multiple regions. This may be impor-
tant in natural images, where large regions are sometimes occluded.

The method is straightforward: consider two regions A and B. All pairs of


points (a, b) in the set S = {(a, b)|a ∈ A, b ∈ B} are compared (a ”walkthrough”
26

each of the regions). For some pairs, A will lie northeast of B, for others A will
lie SE, etc. for the four quadrants. The fraction of pairs contained in each of the
four quadrants are computed: four ”weights”: wN E , wN W , wSE , and wSW .

Finally, these can be converted to above/below, left/right, and diagonality by


computing four features: above = wN E +wN W , right = wN E +wSE , and diagonality
= wN E + wSW .

One advantage of this method is its ability to handle 2D, occluded ((i.e.,
disconnected) landmarks and trajectors.

Hybrid Approach

In their research, Luo and Zhu [40] use a hybrid approach, combining the bounding
box and weighted walkthrough methods. The method was designed for modeling
spatial relations between materials in natural scenes, and so favors above/below
calculations. It skips weighted walkthroughs when object bounding boxes do not
overlap. It does not handle some obscure cases correctly, but is fast and correct
when used in practice.

The final decision of the spatial relationship model would be most appropriate
for my work depends in large part on whether the AVS method can be extended
to 2D trajectors while being made computationally tractable. One answer may
be to work on a higher conceptual level than individual pixels.

2.3.2 Computing Quantitative Spatial Relationships

While computationally more expensive and possibly too sensitive, more detailed
spatial information may be necessary to distinguish some scene types. Rather
than just encoding the direction (such as “above”) in our knowledge framework,
we could incorporate a direction and a distance. For example, Rimey encoded
spatial relationships using an expected area net [56].
27

In the limit, we may wish to model the class-conditional spatial distributions as


continuous variables. There is a body of literature addressing the issue of efficient
inference of continuous variables in graphical models (e.g. Bayes Nets), if the
distributions are assumed Gaussian (e.g., [33]). Felzenszwalb and Huttenlocher
[19] used Gaussian models in their object recognition system, which we will review
in the next section.

Another option, if a lattice-structured Markov Random Field framework is


used (also discussed in the next section), would be to use the spatial relationships
that arise from the lattice structure.

2.4 Probabilistic Graphical Models

Early research involving spatial relationships between objects used logic [2]. This
approach was not particularly successful on natural scenes: while logic is certain,
life is uncertain. In an attempt to overcome this limitation, more recent work has
extended the framework to use fuzzy logic [44].

However, Pearl [50] argues that logic cannot be extended to the uncertainty
of life, where many rules have exceptions. The rules of Boolean logic contain no
method of combining exceptions. Furthermore, logic interactions occur in stages,
allowing for efficient computation. We would like to handle uncertain evidence
incrementally as well. But unless one makes strict independence assumptions, this
is impossible with logic–and computing the effect of evidence in one global step
is impossible.

Logic-based (syntactic or rule-based) systems combine beliefs numerically. The


uncertainty of a formula is calculated as a combination of the uncertainties of the
sub-formulas. Computationally, this approach mirrors the process of logical infer-
ence, leading to an efficient, modular scheme. Rules can be combined regardless
of other rules and regardless of how the rule was derived. Semantically, these
28

assumptions are too strong, except under the strongest of independence assump-
tions. They cause the following problems semantically.

1. Bidirectional inferences. Semantically, if we have the two statements, fire→smoke


and smoke, then fire should be more plausible. However, in a rule-based sys-
tem, this would introduce a loop.

2. Limits of modularity. Consider the rules alarm → burglar and alarm →


earthquake. Due to the modular nature of logic, if alarm becomes more
plausible, then burgular should become more plausible as well. However,
using plausible reasoning, if we add evidence for earthquake, then alarm
becomes more plausible and burglar becomes less plausible, which corre-
sponds with human intuition.

3. Correlated evidence. The rules of logic cannot handle multiple pieces of


evidence originating from a single source. As an example, one should not
independently increase the belief of an event based on many local news
stories that merely echo the Associated Press.

Some attempts have been made to overcome this last limitation, such as bounds
propagation or user-defined combination; however, each approach introduces fur-
ther difficulties.

We are fully aware that there is not universal agreement with Pearl philo-
sophically regarding the superiority of probability over logic. (Witness the heated
rebuttals to Cheeseman’s argument for probability [14] by the logic community!)
Still, we think his arguments are sound.

Specifically, Pearl argues for a graphical model-based approach founded on


probability calculus. While he elaborated on Bayesian Networks in [50], we also
consider Markov Random Fields (MRF), another probabilistic graphical model
29

that has been used primarily for low-level vision problems (finding boundaries,
growing regions), but has recently been used for object detection.

In general, graphical probability models provide a distinct advantage in prob-


lems of inference and learning, that of statistical independence assumptions. In
a graphical model, nodes represent random variables and edges represent depen-
dencies between those variables. Ideally, nodes are connected by an edge if and
only if their variables are directly dependent; however, many models only capture
one direction.

Sparse graphs, in particular, benefit from the message-passing algorithms used


to propagate evidence around the network. While the calculation of a joint proba-
bility distribution takes exponential space (and marginals are difficult to calculate)
in general, these calculations are much cheaper in certain types of graphs, as we
will see.

2.4.1 Bayesian Networks

Bayesian (or belief ) networks are used to model causal probabilistic relationships
[13] between a system of random variables. The causal relationships are repre-
sented by a directed acyclic graph (DAG) in which each link connects a cause (the
“parent” node) to an effect (the “child” node). The strength of the link between
the two is represented as the conditional probability of the child given the parent.
The directed nature of the graph allows conditional independence to be specified;
in particular, a node is conditionally independent of all of its non-successors, given
its parent(s).

The independence assumptions allow the joint probability distribution of all


of the variables in the system to be specified in a simplified manner, particularly
if the graph is sparse.

Specifically, the network consists of four parts, as follows [66]:


30

• Prior probabilities are the initial beliefs about the root node(s) in the net-
work when no evidence is presented.

• Each node has a conditional probability matrix (CPM) associated with it,
representing the causality between the node and its parents. These can be
assigned by an expert or learned from data.

• Evidence is the input presented to the network. Nodes can be instantiated


(by setting the belief in one of its hypotheses to 1) or set to fractional
(uncertain) beliefs (via virtual evidence [50]).

• Posteriors are the output of the network. Their value is calculated from
the product of priors and likelihoods arising from the evidence (as in Bayes’
Rule).

The expressive power, inference schemes and associated computational com-


plexity all depend greatly on the density and topology of the graph. We discuss
three categories: tree, poly-tree, and general DAG.

Trees

If the graph is tree-structured, with each node having exactly one parent node,
each node’s exact posterior belief can be calculated quickly and in a distributed
fashion using a simple message-passing scheme. Feedback is avoided by separating
causal and diagnostic (evidential) support for each variable using top-down and
bottom-up propagation of messages, respectively.

The message-passing algorithm for tree-structured Bayesian networks is simple


and allows for inference in polynomial time. However, its expressive power is
somewhat limited because each effect can have only a single cause. In human
reasoning, effects can have multiple potential causes that are weighed against one
another as independent variables [50].
31

Causal Polytrees

A polytree is a singly-connected graph (one whose underlying undirected graph


is acyclic). Polytrees are a generalization of trees that allow for effects to have
multiple causes.

The message-passing schemes for trees generalize to polytrees, and exact pos-
terior beliefs can be calculated. One drawback is that each variable is conditioned
on the combination of its parents’ values. Estimating the values in the condi-
tional probability matrix may be difficult because its size is exponential in the
number of parent nodes. Large numbers of parents for a node can induce con-
siderable computational complexity, since the message involves a summation over
each combination of parent values.

Models for multicausal interactions, such as the noisy-OR gate, have been
developed to solve this problem. They are modeled after human reasoning and
reduce the complexity of the messages from a node to O(p), linear in the number
of its parents. The messages in the noisy-OR gate model can be computed in
closed form (see [50]).

A nice summary of the inference processes for trees and polytrees given in [50]
can be found in [66].

General Directed Acyclic Graphs

The most general case is a DAG that contains undirected loops. While a DAG
cannot contain a directed cycle, its underlying undirected graph may contain a
cycle, as shown in Figure 2.1.

Loops cause problems for Bayesian networks, both architectural and semantic.
First, the message passing algorithm fails, since messages may cycle around the
loop. Second, the posterior probabilities may not be correct, since the conditional
independence assumption is violated. In Figure 2.1, variables B and C may be
32

B C

Figure 2.1: A Bayes Net with a loop.

conditionally independent given their common parent A, but messages passed


through D from B will also (incorrectly) affect the belief in C.

There exist a number of methods for coping with loops [50]. Two methods,
clustering and conditioning, are tractable only for sparse graphs. Another method,
stochastic simulation, involves sampling the Bayesian network. We use a simple
top-down version, called logic sampling, as a generative model and describe it in
Section 3.2.2. However, it is inefficient in the face of instantiated evidence, since
it involves rejecting each sample that does not agree with the evidence.

Finally, the methods of belief propagation and generalized belief propagation, in


which the loops are simply ignored, has been applied with success in many cases
[82] and is worth further investigation. We discuss these methods in the context
of MRFs in the next section.

Applications of Bayesian Networks

In computer vision, Bayesian networks have been used in many applications in-
cluding indoor vs. outdoor image classification [37; 48], main subject detection
[66], and control of selective perception [57]. An advantage of Bayesian networks
33

is that they are able to fuse different types of sensory data (e.g. low-level and
semantic features) in a well-founded manner.

2.4.2 Markov Random Fields

Markov Random Fields (MRFs), or Markov Networks, model a set of random


variables as nodes in a graph. Dependencies between variables are represented
by undirected arcs between the corresponding nodes in the graph. The topology
of the network explicitly identifies independence assumptions – absence of an
arc between two nodes indicates that the nodes are assumed to be conditionally
independent given their nbrs. MRFs are used extensively for problems in low-
level computer vision and statistical physics. MRFs provide a framework to infer
underlying global structure from local observations.

We now discuss the basic concepts of MRFs, drawing from our in-house review
[7] of the typical treatments in the literature [30; 16; 22; 15].

Random Field A set of random variables X = {xi }.

Graphical Model A random field X may be represented as a graphical model


G = (Q, E) composed of a set of nodes Q and edges E connecting pairs
of nodes. A node i ∈ Q represents the random variable xi ∈ X. An
edge (i, j) ∈ E connecting nodes i and j indicates a statistical dependency
between random variables xi and xj . More importantly, the lack of an edge
between two graph nodes indicates an assumption of independence between
the nodes given their neighbors.

Configuration For a random field X of size n, a configuration ω of X assigns


a value (x1 = ω1 , x2 = ω2 , . . . xn = ωn ) to each random variable xi ∈ X.
P (Ω) is the probability density function over the set Ω = {ω} of all possible
configurations of X.
34

Neighborhood Relationship We define a neighborhood relationship N on a


field X as follows. Let E be a set of ordered pairs representing connections
(typically probabilistic dependencies) between elements of X. Then for any
xi , xj ∈ X, xj ∈ Ni ⇔ (xi , xj ) ∈ E.

Markov Property [30]

For a variable xi in a random field X, if xi satisfies the Markov Property,

P (xi = ωi |xj = ωj , j 6= i) = P (xi = ωi |xj = ωj , j ∈ Ni )

The probabilities in Equation 2.4.2 are called local characteristics and intu-
itively describe a locality condition, namely that the value of any variable
in the field depends only on its neighbors.

Positivity Condition The positivity condition states that for every configura-
tion ω ∈ Ω, P (x = ω) > 0.

Markov Random Field Any random field satisfying both the Markov property
and the positivity condition. Also called a Markov Network.

Two-Layer MRF [23; 16; 22]

“Two-Layer” describes the network topology of the MRF. The top layer
represents the input, or evidence, while the bottom layer represents the
relationships between neighboring nodes (Figure 2.2).

In typical computer vision problems, inter level links between the top and
bottom layers enforce compatibility between image evidence and the un-
derlying scene. Intra-level links in the top layer of the MRF leverage a
prioriknowledge about relationships between parts of the underlying scene
to enforce consistency between neighboring nodes in the underlying scene
[16]..
35

Figure 2.2: Portion of a typical two-layer MRF. In low-level computer vision prob-
lems, the top layer (black) represents the external evidence of the observed image
while the bottom layer (white) expresses the a prioriknowledge about relationships
between parts of the underlying scene.

Pairwise MRF [23; 16; 82]

In a pairwise MRF, the joint distribution over the MRF is captured by a set
of compatibility functions that describe the statistical relationships between
pairs of random variables in the MRF. For inferential purposes, this means
that the graphical model representing the MRF has no cliques larger than
size two.

Compatibility Functions The statistical dependency between the two random


variables xi , xj in a random field is characterized by a compatibility function
ψi,j (ωi , ωj ) that scores every possible pair of hypotheses (xi = ωi , xj = ωj ).

As an example, consider a link (i, j) in a graphical model G connecting


nodes i and j. If there are three possible outcomes for xi and two possible
outcomes for xj , the compatibility function relating i and j is a 3×2 matrix,
M = [mij ].

Depending upon the problem, compatibilities may be characterized by either


the joint distribution of the two variables2 . For some problems for which
2
Or equivalently by both conditional distributions (p(x,y) may be obtained from p(x|y) and
p(y|x).
36

the joint is unobtainable, a single conditional distribution suffices (e.g.for a


problem for which p(x|y) is known, but p(y|x) cannot be computed).

Inference

In typical low-level computer vision applications of MRFs, what is desired from


the inference procedure is the MAP estimate of the true scene (the labeling), given
the observed data (the image). We have identified two complementary approaches
in the literature for calculating the MAP estimate: deterministic techniques and
Monte Carlo techniques (described later in this section).

We start by reviewing two deterministic techniques: Belief Propagation and


Highest Confidence First. The Highest Confidence First algorithm finds local
maxima of the posterior distribution by using the principle of least commitment
[43], while belief propagation is an inexact inference procedure using message-
passing algorithms successfully in loopy networks by simply ignoring the loops.

Highest Confidence First (HCF) [16]

The HCF algorithm is used for MAP estimation, finding local maxima of
the posterior distribution. It is a deterministic procedure founded on the
principle of least commitment. Scene nodes connected to image nodes with
the strongest external evidence (i.e. a hypothesis with a large ratio of the
maximum-likelihood hypothesis to the others) are “committed” first, since
they are unlikely to change (based on compatibility with neighbors). Nodes
with weak evidence commit later and are based primarily on their compat-
ibility with their “committed” neighbors.

Using edge-modeling MRF as an example, large intensity gradients might


constitute strong evidence in some of the network’s nodes. The nodes with
strong evidence should influence scene nodes with weaker evidence (via edge
continuity constraints) more than the other way around.
37

Belief Propagation [82; 22; 29]

The Belief Propagation (BP) algorithm is a message-passing algorithm for


probabilistic networks. It is a generalization of a number of inference al-
gorithms such as the forward-backward algorithm, the Viterbi algorithm,
Pearl’s algorithm for Bayesian polytrees, and the Kalman filter.

At each iteration, each node computes a belief, which is a marginal of the


joint probability. The belief is a function of local compatibilities, φ i (xi )
(e.g. local evidence nodes, which are constant, can subsumed) and incoming
messages, mji from neighboring nodes:

X
bi (xi ) = kφi (xi ) mji (xi )
j∈N (i)

The messages, mji , are computed from a function of compatibilities of


the message’s sender and recipient nodes and previous messages from the
sender’s other neighbors:

X Y
mij (xj ) = φ(xi , yi )ψ(xi , xj ) mki (xi )
xi k∈N (i)j

Intuitively, the incoming messages represent combined evidence that has


already propagated through the network.

In the rare case that the graph contains no loops, it can be shown that
the marginals are exact. However, some experimental work suggests that
at least for certain problems, that the approximations are good even in the
typical “loopy” networks, as the evidence is “double-counted” [81].

One can calculate the MAP estimate at each node by replacing the summa-
tions in the messages with max.

Generalized belief propagation [82]


38

Generalized belief propagation (GBP) uses messages from clusters of nodes


to other clusters of nodes. Since these messages are expected to be more
informative, the performance is also expected to increase.

GBP is theoretically justified: in fact, GBP is related to Kikuchi approxi-


mations in a manner analogous to BP and Bethe free energy: a set of beliefs
gives a GBP fixed point in a graph if and only if the beliefs are local station-
ary points of the Kikuchi free energy (for details of free energy minimization
techniques, see [82]).

GBP has been found to perform much better than BP on graphs with short
loops. The drawback is that the complexity is exponential in the cluster
size, but again, if the graph has short loops (and thus necessitates only
small clusters), the increased complexity can be minimal.

Pearl’s clustering algorithm is a special case of GBP, with clusters chosen


to overlap in a fixed manner that are usually large. They obtain increased
accuracy, but at increased complexity.

An advantage of GBP is that it can be used to vary the cluster size in order
to make a trade-off between accuracy and complexity.

Inference on Tree-Structured MRFs

Felzenszwalb and Huttenlocher [19] use tree-structured MRFs for recognition of


objects such as faces and people. They model their objects as a collection of
parts appearing in a particular spatial arrangement. Their premise is that in a
part-based approach, recognition of individual parts is difficult without context,
and needs spatial context for more accurate performance.

They model the expected part locations using a tree-structured two-layer


MRF. In the scene layer, the nodes represent parts and the connections repre-
39

sent general spatial relationships between the parts. However, rather than using
the typical square lattice, they use a tree.

Inference in the MRF is both exact and efficient, due to the tree structure.
Their MAP estimation algorithm is based on dynamic programming and is very
similar in flavor to the Viterbi algorithm for Hidden Markov Models. In fact, the
brief literature in the field on using Hidden Markov Models for object and people
detection [20] might be better cast in an MRF framework.

Monte Carlo Methods

Our treatment is taken in large part from [7; 45; 41; 23].

Monte Carlo methods are used for sampling. The goal is to characterize a
distribution using a set of well-chosen samples. These can be used for approximate
MAP estimation, computing expectations, etc., and are especially helpful when
the expectations cannot be calculated analytically.

How the representative samples are drawn and weighted depends on the Monte
Carlo method used. One must keep in mind that the number of iterations of the
various algorithms that are needed to obtain independent samples may be large.

Monte Carlo Integration Monte Carlo integration is used to compute expec-


tations of functions over probability distributions. Let p(x) be a probability
R
distribution and a(x) be a function of interest. We assume that a(x)p(x)
cannot be evaluated analytically, but p(x) is easy to sample from (e.g., Gaus-
R
sian). To compute a Monte Carlo estimate of a(x)p(x), we first create a
representative sample of Xs from p(x). There will be many Xs from the
regions of high probability density for p(x) (intuitively, the Xs that should
be common in the real world). We then calculate a(x) for each x in the set.
The average value of a(x) closely approximates the expectation.

A key insight into this concept can be stated as follows:


40

The expectation of a function of random variables depends not


only on the function’s value, but on how often the random vari-
ables take certain values!

The drawback is that our assumption that p(x) is “easy” is often not valid;
p(x) is often not easy to sample from, and so we need a search mechanism
to draw good samples. Furthermore, we must be careful that this search
mechanism does not bias the results.

Monte Carlo Markov Chain (MCMC) Methods [45; 41]

In Monte Carlo Markov Chain (MCMC) methods, the samples are drawn
from the end of a random walk.

A Markov chain is a series of random variables, X 0 , X 1 , . . . , X t in which a


locality condition is satisfied, that is,

P (X ( t + 1)|X t , X ( t − 1), . . . , X 0 ) = [P (X ( t + 1)|X t )

The chain can be specified using the initial probability, p0 (x) = P (X 0 ) and
the transition probabilities, p(X ( t + 1)|X t ). The transition probability of
moving from state x to state y at time t is denoted Tt (x, y) , which can be
summarized in a transition matrix, Tt .

We consider homogenous Markov chains, those in which the transition prob-


abilities are constant, i.e. Tt = T for all T .

We take an initial distribution across the state space (which, in the case of
MRFs, is the set of possible configurations of the individual variables).

This distribution is multiplied by a matrix of transition probabilities re-


peatedly, each iteration yielding a new distribution. The theory of Markov
chains gaurantees that the distribution will converge to the true distribution
if the chain is ergodic (always converging to the same distribution).
41

In practice, one generates a sample from the initial probability distribution


(e.g., uniform) and then moves through the state space stochastically; a
random walk guided by the values in the transition matrix. Since the distri-
bution is guaranteed to converge, at the end of the random walk, the sample
should be from the actual distribution.

The number of steps needed before convergence is reached is bounded by


theoretical results, but varies in practice.

There are a number MCMC algorithms:

• Gibbs Sampler. [23; 45; 41]

In the Gibbs sampling algorithm, each step of the random walk is taken
along one dimension, conditioned on the present values of the other
dimensions. In a MRF problem, it is assumed that the conditional
probabilities are known, since they are local (by the Markov property).

This method was developed in the field of physics, but was first applied
to low-level computer vision problems by Geman and Geman on the
problem of image restoration [23]. Geman and Geman furthermore
combined Gibbs sampling with simulated annealing to obtain not just
a sample, but the MAP estimate of their distribution.

Their application of Gibbs sampling is also called stochastic relaxation


(so as to differentiate it from deterministic relation techniques).

• Metropolis Sampler. [45; 41]

In each iteration of the Metropolis algorithm, one makes a small change


from the current state, and accepts the change based on how good
(probabilistically) the new state is compared to the old one.

• Metropolis-Hastings Sampler. [41]

Generalization of the Metropolis algorithm.


42

Importance Sampling [45; 41]

To sample from a distribution f (x), sample from a simpler distribution,


g(x), and weight based on the ratio between the original distribution and
g(x). One caveat is that g(x) should have heavy tails, e.g. Cauchy, because
g(x) should not equal zero where the original distribution has non-zero prob-
ability.

Rejection Sampling [45; 41]

To sample from a distribution f (x), sample from another, similar distribu-


tion g(x), which is bounded by a constant multiple, c, of the true distribu-
tion. Generate a point x̃ from g(x) and accept the point with probability
f (x)/cg(x), repeating until a point is accepted.

The efficiency of the method depends on how tight c is and how close the
functions f (x) and g(x) are, and is not useful in practice for “difficult”
problems [45].

2.4.3 Relative Merits

The literature is divided regarding the potential equivalence of Bayesian networks


and Markov Random Fields. Pearl argues that only a subset of each model (de-
composable models) are equivalent, due to the differing semantics provided by
directed and undirected links [50]. However, Yedidia [82] argues that they are
equivalent, providing algorithms for converting each to and from factor graphs, a
third graphical model.

While we need to settle this issue in our minds, we argue that even if the models
are technically equivalent, their ease of use is not necessarily; each definitely has
3
its particular merits. Bayesian networks model causal dependencies and allow
3
Consider the theoretical equivalence of Turing machines with modern computers; which is
easier to program for practical tasks?
43

for efficient inference in sparse graphs. Their utility in high-level vision problems
has been proven. In contrast, Markov Random Fields have been used primarily
for low-level vision problems in which the spatial constraints can be specified
directly and only recently have been applied to object recognition (in which case
a tree-structured MRF was used). We believe that each model warrants further
investigation.
44

3 Methodology

In this chapter, we discuss our experimental setup, the statistics calculated on


natural images, the advantages of working in simulation, and the details of the
simulator.

For our prototype scene classifier using spatial information, we have focused
on the problem of classifying outdoor scenery (i.e., natural ) scenes. Many other
attempts have been made to classify this type of scene (e.g., [4; 9; 27; 35; 44; 46;
77]), all with some degree of success. One reason is that in natural scenes, low-
level features tend to be more correlated with semantics (than scenes of human
activity). Another is that there is a limited range of object type present in these
scenes [70]. Some research suggests that natural scene databases are less complex
(by a certain metric) than those of people [52]. We do note that these findings
are by no means conclusive because the range of features that were used in the
experiments was extremely narrow.

Furthermore, outdoor scenery images tend to contain specific, albeit loose,


spatial arrangements. For instance, in a beach scene, we would expect to see sky
appearing above water, in turn appearing above sand. One reason for this seems
to be the large field-of-view in these images: perspective is much more constrained
than in short-range, human-activity dominated scenes.
45

Finally, the ultimate test for our research is to solve a practical, yet very
interesting problem: classifying real consumer-type images. In much of the scene
classification work, as we have seen, experimental databases are usually limited
to professional stock photo libraries. If we can classify consumer images, which
are much less constrained in color and composition, successfully, our work could
potentially generalize to other domains. Robot vision, for instance, needs to
operate in non-structured environments with low-quality sensors.

When the time comes to experiment with real images and real detectors, we
would be well-equipped to detect the materials in these scenes. We have access
to state-of-the-art detectors for sky (blue [38] and mixed [67]), as well as grass,
water, and snow [39].

We have also started to gather training data, labeling approximately 4000


images classified as beach, sunset, fall foliage, field, mountain, urban, night, lake
and ocean, skiing, and desert. The materials and locations of approximately 300
of these images have been hand-labeled as well, as we describe in the following
section.

3.1 Statistics of Natural Images

For a small set of these natural images, we have marked the location and extent of
each material in each image. While gathering data about the presence or absence
of materials is fairly straightforward (and can be represented by text), learning the
spatial relationships between them requires labeling the materials in each image
in such a way that the spatial information can be extracted. Because we want to
compare the effects of qualitative and quantitative spatial relationships, it is not
enough to collect data in the form “the sky is above the sand”. We need to outline
each material’s region in each image in some manner, so that the relationships
can be calculated. A sample of the desired information is captured in Figure 3.1.
46

Figure 3.1: Ground-truth labeling of a beach scene. Sky, water, and sand regions
are clearly shown.

3.1.1 Ground Truth Collection Process

Ground truth collection involved two stages, collecting images of each class and
labeling the material regions in each image.

Scene Class Definitions

As stated, we have collected approximately 400 images of each of the following


types of scenes: beach, sunset, fall foliage, field, mountain, and urban. These
scene types are like those used in the literature. These are chosen such material
detection of various types should be helpful. The scene definitions are given in
Table 3.1.

In general, images chosen did not have a dominant main subject. This is
legitimate; for example, a close-up image containing a horse in a field would be
considered a “horse” image, not a “field” image. We do plan to investigate later
how to adapt the framework for such images.
47

Table 3.1: Scene class definitions.


Class Definition

beach At least 5% each of water, sand, and sky


sunset Illuminant source in front of camera
fall foliage Detectable color in turned leaves on tree
field No aerial view, not cluttered with trees (“open”)
mountain Open, whole mountains, mid-range view
urban At least one building, no extreme perspectives

Material Region Labeling

Our region labeling was done in a principled fashion, using the following method-
ology:

• Identify key materials. We chose a number of materials to label in each


image: sand/gravel, sky (blue and mixed/cloudy), water, snow, buildings
(skyscrapers), roads, grass, foliage (evergreen, green and turned deciduous),
rock, crowds of people, and crops.

• Define each material precisely. For instance, how is a crowd of people de-
fined? Is wet sand on a beach considered part of a sand or water region? We
defined “sand/gravel” as dry sand or gravel, and ignored wet sand, some-
times causing wide borders between the sand and water regions as a result.
For a detailed list of definitions, see [39].

• Label each image. We select unobstructed (continuous) areas of each mate-


rial; for instance, we do not select the bits of sky visible between the leaves
of a tree. While our labeling tool’s user interface allows us to create and
label only polygonal regions, this is sufficient.
48

The labeled regions correspond to a semantic-based coarse segmentation of


each image, which is ideal for training. The labels correspond to semantic re-
gions, even when changes in appearance (e.g., deep shadows) would fool many
of today’s segmentation algorithms. Then, in the classification stage, errors are
either concept failures, segmentation errors, or material detection errors.

For each image region, we selected a polygon contained strictly in the interior
of the region, yielding a “pure” sample of that region. Avoiding the boundary
altogether is simpler, yet leaves spaces between the regions (as shown in Figure
3.1). We could, of course, achieve a closer approximation using a minimization
refinement such as snakes [71]; however, it is doubtful when such a fine-grained
segmentation would help.

3.1.2 Scene Prototypes

Prototypical beach scenes, in order of increasing complexity, are given in Figure


3.2.

We labeled a total of 321 images, belonging to the beach, urban, and field
classes.

3.1.3 Spatial Relationships

We use Luo, et al.’s hybrid bounding-box/weighted walkthrough approach (see


Section 2.3.1) to computing spatial relationships. In [40], the relations were used
to solve the related problem of building holistic material detection using spatial
context-aware scene models 1 . They were interested in outdoor scenes and the
typical materials found there as well. The relations of interest are above, below,
beside, far above, and far below, but favoring the above and below relations over
1
Note that they did not build scene-specific models; we could apply our research to solve
this.
49

Figure 3.2: Prototypical beach scenes. (a) A simple beach scene without back-
ground objects. (b) Because we make no attempt to detect it, we consider the
sailboat to be “background”. (c) A more complicated scene: a developed beach-
front. (d) A scene from a more distant field-of-view. (e) A crowded beach.

the beside relation, because they were found to be more helpful for outdoor scenes.
We use their algorithm for its proven performance on similar scenes (and because
it is faster than the other methods discussed).

We make the simplifying assumption that multiple (potentially) disconnected


regions with the same semantic label should be considered as one region for the
purposes of computing spatial relations. This is not generally an issue; see, for
example, the sand regions in Figure 3.2 (e); calculating their relationship as one
entity with respect to the sky or water is equivalent to computing the individual
relationships and merging them.

This also solves the issue of conflicting relations between a material’s subre-
gions in an intuitive fashion. As a simple example, see Figure 3.4. We have two
regions (A1 and A2 ) of material MA and one region (B) of material MB . A1
is beside B and A2 is above B. However, by considering A1 and A2 as one re-
50

Figure 3.3: Prototypical urban scenes. (a) The most common urban scene, con-
taining sky, buildings, and roads. (b),(c) The sky is not simply above the buildings
in these images. (d) Roads are not necessary. (e) Perspective views induce varied
spatial relationships. (f) Close views can preclude the presence of sky.

gion, the resulting bounding box overlaps region B and the weighted-walkthrough
algorithm yields a relationship favoring the larger region.

A2

A1 B

Figure 3.4: An example of spatial relationships involving a split region.


51

3.1.4 Scene-specific Spatial Relationship Statistics

Using the relationships defined above, we extracted spatial relationships statistics


for two classes of natural images: beach and urban scenes. The statistics are given
in Appendix A.

3.2 Experimental Environment

One goal of our research is to develop and demonstrate representations and al-
gorithms for difficult semantic scene classification problems. We start with prop-
erties of image regions and their spatial relations. We exploit probabilistic con-
straints on interpretations (discrete labels) of regions arising from cultural and
physical regularity and predictability. A simulated environment is essential for us
to manipulate and explore the various relationships, properties, algorithms, and
assumptions available to us without being constrained by the “accidental” (in the
philosophical sense) nature of any particular domain.

However, another goal certainly is to solve practical problems in scene classifi-


cation, spurred on by the motivations given in the Introduction. That is, we want
our solutions to be specific as well as general, and we care about application as
well as abstraction.

These goals can stand in competition with each other: abstracting away too
much in a simulating environment can be totally unrealistic and give little hope
of solving the original problem. Our goal is to balance these two goals.

We can safely abstract away the bank of material and object detectors available
by representing them by their probabilities of correct and incorrect detection,
along with their expected beliefs. For some detectors, we learn these beliefs from
training data, and for others, we estimate them. We can also abstract away the
images (to some degree) by using a probabilistic generative model, as will be
52

described. We do not have to label a large number of images, and yet we learn
the probabilities from the images we have labeled, thus achieving some balance.

Using these abstractions, the simulator should allow us to experiment with


schemes using various degrees of spatial information: qualitative vs. quantita-
tive. Furthermore, we will want to track the effect on each of these schemes of
substituting full knowledge of the scene parts with potentially faulty detectors.

We have developed a prototype of such an environment, dubbed MASSES


(Material And Spatial Simulated Experimental Scenes). In MASSES, materi-
als and their relationships are abstracted for classification purposes. Each scene
contains “objects” arranged according to some loose structure.

MASSES is similar in spirit to Rimey’s T-world [56; 57], except T-world mod-
eled scenes with tight structure (e.g., a table setting), and should generalize to
scenes such as traffic scenes or medical diagnosis, where parts of the image follow
precise rules. It is doubtful that T-world could model natural scenes, which have
a much looser structure.

3.2.1 Advantages of Working in Simulation

We now discuss the specific advantages of idealizing the world with MASSES.

1. It is a standalone work, under our control and free from outside influences.
Therefore, we are not dependent upon others’ ground truth data collection
or upon their detectors. Intellectual property issues are avoided.

2. It can embody the ideas we hold about the world, namely the strength and
granularity of the spatial organization and the accuracy of the detectors.

3. We can experiment with various algorithms.

4. Perhaps most importantly, we can easily quantify the value of added infor-
mation. We can answer questions such as:
53

(a) What is the value of spatial information? Intuitively, we believe that


it should help, but we need to quantify the effect.

(b) How does performance decline when we replace true material knowledge
with faulty detectors?

(c) How does a better sky detector help? We can vary the performance
continuously and track the effects.

(d) How does adding another spatial constraint help? For instance, cliques
in a region adjacency graph (RAG) [71] may contain discriminatory
information.

(e) How strict do the spatial rules have to be before we can benefit from
them? How fine-grained do the measurements have to be?

(f) What is the value of added training data?

3.2.2 MASSES Prototype

The expected output of the simulator is a series of virtual scenes, which can be
used to train a classifier or to test its performance. In this section, we describe
our scene generative model.

For our prototype simulator, we have chosen to use a tree-structured Bayesian


network as an inference engine. Bayesian networks of this form are well-understood
and have efficient inference algorithms, as we discussed in Section 2.4.1. More ex-
pressive formalisms, such as Bayesian networks (for general graphs), or Markov
Random Fields, may better capture the spatial relationship distributions, but we
have not yet explored these options.

The flow of information in MASSES is depicted in Figure 3.5.


54

Material Detection

Material List Spatial Rels.

Sky
Water MASSES
Sand Above(Sk, Wa)
Above(Sky, Sa)
Beside(Wa, Sa)

Inference Engine

Beach Class

Figure 3.5: The MASSES environment. Statistics from labeled scenes are used to
bootstrap the generative model, which can then produce new virtual scenes for
training or testing the inference module.

Bayesian Network

We designed a Bayesian network to classify scenes. We place the conditional


independence assumption on each material, given the scene class. We also assume
that all of the spatial relationships are independent, given the scene class. Our
single-level network is shown in Figure 3.6.

Class

Sky Water ... Material_n Rel(Sky, Wat) ... Rel(M_i, M_j)

Figure 3.6: Single-level Bayesian network used for MASSES


55

The network has been designed with the following assumptions in mind:

1. We model regions as point sources, and model qualitative spatial relation-


ships between them. This means we ignore the shape and size of the region,
adjacency information about the regions, and occlusions causing regions to
be split. We also ignore the underlying color and texture features of the
region which were used to determine the classification of the region. While
any of these may be useful features in a full-scale system, we temporarily
ignore them.

2. We treat each material independently. This ignores the potential co-occurrence


relations between materials: this is a limitation if materials are dependent,
given the class. For example, in beach images, roads are more likely to be
present when a building is also present.

3. We model spatial relationships as binary, symmetric relations, such as A AboveB .


This raises interesting questions. In some cases, relationships are pairwise
consistent, yet globally inconsistent. For example, it is possible for region R 1
to be above region R2 , region R2 to be above region R3 , and region R1 to be
below region R3 . Of course, while this would be impossible when modeling
regions as point sources, it is certainty possible (though unlikely) with 2D
regions of varying shape (consider region R3 in the shape of the letter “C”,
enclosing the other two regions).

4. We assume flat priors on the scene classes, thus eliminating their effect.
While this assumption does not hold in practice, the analysis becomes sim-
pler and we lose no generality.

Finally, we base the values of the conditional probability matrices (CPMs)


on the material and spatial relationship links on the statistics of natural scenes
presented in Section 3.1. Specifically, we round the probabilities and ignore rare,
56

apparently less important materials (like the rarely-occurring grass region on a


beach).

Sampling the Bayesian Network to Generate Scenes

Our first experiment is to determine if best-case (100% accurate) material de-


tection alone (without spatial relations) can give reliable scene classification. If
not, we want to determine the effects of adding spatial information. To this end,
we ignore detector characteristics and create scenes directly from the material
distributions learned from the outdoor scenes.

With this in mind, the method of sampling the Bayes Net is straightforward,
using its priors and the conditional probability matrices of its links, because the
network is a directed and acyclic graph (DAG). The algorithm called probabilistic
logic sampling for DAGs was proposed by Henrion [25] (and extended by Cowell
[17]) 2 .

We apply Henrion’s algorithm as follows:

1. Generate the class, C, of the scene by randomly sampling the priors, P (class =
C).

2. Generate the list of materials present in the scene. Each of the materials has
as its parent the scene’s class, so we sample from P (materiali |class = C)
(one row in the CPM on the link) for each material i.

3. Generate spatial relationships for each pair of materials < i, j > present in
the scene by sampling P (Mi relij Mj |class = C).

Figure 3.7 shows an example. Note that there is a separate node in the network
for each material and for each pair of materials in the image, a total of (m C2 + m)
2
If the graph is not a DAG, a more complicated method such as rejection or a different
framework such as a nested junction tree [17] can be used.
57

Class

1 2 3 4 ... Rel(1,2) Rel(1,3) Rel(1,4) Rel(2,3) Rel(2,4) Rel(3,4) ...

Figure 3.7: Sampling the scene type yields class C. Then we sample to find the
materials present in the image, in this case, M1 , M3 , and M4 . Finally, we sample
to find the relationships between each pair of these material regions.

nodes for m materials. However, when generating scenes or doing inference, we


need only instantiate those nodes corresponding to materials already determined
to be present in the image. If d materials are detected in a given image, there will
be d C2 instantiated relationship nodes. In practice, this is a large savings because
for typical scenes, d < m.

The alternative to Henrion’s algorithm is to use a guess-and-check method:


generate a random placement of material regions and then use the Bayes Net
to classify them. In a non-forced choice classification problem, this would give
the ability to generate scenes in an “other” category, but it would also generate
non-sensical scenes. It is also extremely inefficient.

Inference in the Bayesian Network for Classifying Scenes

To classify scenes generated in this fashion is straightforward. A scene is rep-


resented by a list of materials and binary relations. The material leaves are
instantiated as either present or not present. Only the relations corresponding
to materials present in the scene are instantiated; the others cannot be because
two materials not present have no relationship to instantiate. This also causes
inference to be much more efficient. The likelihoods propagate to the root. Since
the priors are flat, the scene is classified as the one with the maximum likelihood.
58

3.2.3 Simulating Faulty Detectors

If we were to have full knowledge of the materials in a scene, then many scenes
could be distinguished by the mere presence or absence of materials only. For
instance, consider the outdoor scenes described above. An image containing only
sky, a skyscraper, and roads is almost guaranteed to be an urban scene.

However, to be practical, we must allow for faulty detection of objects and


materials in a scene to occur during the classification stage. Our hope is that if
a number of materials are falsely detected, their spatial relationships will yield
low beliefs in any scene type, and can therefore be rejected as no known scene, or
as one that cannot be classified from the materials detected. We can do this by
incorporating a reject option into our inference mechanism (e.g., by thresholding
class beliefs). In the more general case, a region may be classified as multiple
materials, each with corresponding beliefs. The correct material may be detected
with a belief smaller than the maximum, but spatial relationships may boost the
overall belief in the configuration with this material to become the maximum.

A method of simulating faulty detectors will involve a causal relationship be-


tween the true materials and the detectors. Define the characteristics of detector
D on a set of materials M to be the set of probabilities {P (D fires |mi ) : mi ∈ M }.

Take the sand detector as an example:

Sand_Detector
0.05 0.95 Sky
0.95 0.05 Sand
0.20 0.80 Road
0.20 0.80 Water
0.25 0.75 Building
0.30 0.70 Background
59

The first column gives P (Ds |M ), the probability that the sand detector Ds
fires, given the true material M . The second column gives the probability 1 −
(P (Ds |M )) that the detector does not fire. In this example, the sand detector has
95% recall of true sand, and detects sand falsely on 20% of the water, for instance
(perhaps due to a brown reflection in the water or shallow sand). It may also fire
falsely on 30% of the background (unmodeled) regions in the images because they
have similar colors.

Assume we have a detector for each material of interest. Each of these detec-
tors will be connected to any given region in the image and will have a certain
probability of firing given the true material in that region. The evidence consists
of the detectors that actually fired and the corresponding beliefs. Take a sand
region; each of the detectors’ characteristics may be as follows:

True Sand
0.03 0.97 Sky_detector
0.95 0.05 Sand_detector
0.20 0.80 Road_detector
0.20 0.80 Water_detector
0.10 0.90 Building_detector

In this case, the first column gives for each material detector Dm , P (Dm |S),
the probability that the detector fires on a true sand region.

Each detector is linked to each material in the image, as shown in Figure 3.8.
Note that the subgraph’s “root” node corresponds not to a specific material, but
to a specific region.

Inference on this subgraph allows the combined likelihood of each true material
to be calculated, yielding a soft belief in each material. Let Mi be a material and
D be the set of detectors, D = {D1 , D2 , . . . Dn }.

We calculate the likelihood of each detector firing as follows.


60

Region_label

Sky_det Water_det ... Sand_det ... ... Mat_n_det

Figure 3.8: Bayesian network subgraph showing relationship between regions and
detectors.

Y
λ(R) = α λDi (R) (3.1)
i
Y
= α MD|Ri λ(Di ) (3.2)
i

where α is a normalizing constant and the matrix notation for M is defined by


Pearl [50] as:

My|x , P (y|x)

More specifically, the (i, j)th position in My|x , P (Y = yj |X = xi )

Equation 3.1 follows from the conditional independence assumption of detec-


tors and 3.2 follows by conditioning and summing over the values of each detector.

At this point, we have two options. The likelihoods, λ(R) can be passed on to
the remainder of the network (i.e. attaching the subgraph to each material leaf
in the original network) or only the material with the maximum likelihood (ML)
can be passed on by taking arg maxi P (Mi |D). This second option corresponds to
performing inference on the subgraph off-line. Each option can advantages and
disadvantages.

1. Pass up the belief in each material, given which detectors fired on the re-
gion. Maintaining uncertainty concerning the true material is desirable. Our
61

single-level, tree-structured network described above cannot handle beliefs.


The main limitation is that there is no way of describing which material spa-
tial relationships should be instantiated. Consider regions A and B, with
likelihood vectors ΛA = (0.5, 0.3, 0.2) and ΛB = (0.1, 0.2, 0.7) for materials
M1 , M2 , and M3 , respectively. Which relationship node should be instan-
tiated? The combination Rel(A = M1 , B = M3 ) is most likely based on
materials alone, but may not be more likely once the relationship probabil-
ities are included. If we consider all combinations of materials, this leads
to an explosion in number of relationship nodes to be considered during the
inference stage, since each potential material combination must be explored.

2. Use a Maximum Likelihood approach for detecting one material per region.
Even though we lose the uncertainty in the detection, we decided to choose
this approach initially, since incorporating it into the network was straight-
forward.

In the generative stage for test images, we first generate the scene as described
earlier in this section, including true materials and spatial relations. Consider a
single region with known material. We can approximate, by sampling, the proba-
bility of each configuration of detectors firing, and the material chosen using the
ML approach. Doing this offline once can save time and lend insight into the
true quality of the detectors. The result is a distribution of ML detections given
the true material; sampling from this distribution yields a perturbed (possibly
faulty) material that we substitute for the true one. We keep the spatial relation-
ships originally generated; the composition of the region was mis-detected, not its
location.

We describe this approach in the following section.


62

Offline ML Approach to Faulty Detectors using Detector Beliefs

We start by describing how to perturb a single material region. The idea is that
we can use each detector’s belief in the region to determine the material with the
maximum likelihood.

Material Perturbation Algorithm

1. Determine which detectors fire by sampling the detector characteristics.

2. For each detector that fires, sample the belief distribution to determine the
confidence in the detection.

3. Propagate the beliefs in the Bayesian network to determine the overall like-
lihood of each material.

4. Replace the true material with the (possibly different) material with the
maximum likelihood.

We work an example using the detector characteristics given in Appendix B.


Consider the sand region example in the previous section. Given the large (95%)
probability that the sand detector fires, and smaller probabilities that the road and
water detectors fire, it is most likely that only the sand detector fires, obviously
giving no perturbation. However, it is also somewhat likely that the road detector
fires as well. Say that when we sample, we find the sand detector firing with belief
= 0.6 and the road detector firing with belief = 0.7. (That the false detector would
give a larger belief than the true detector is unlikely, yet possible, and we use it
to illustrate a point.) Then calculating the combined likelihood of each material,
we get:

Y
λ(R) = α λDi (R) (3.3)
i∈{KD,SD,RD,W D,BD}
63

where, for example,

 
0.05 0.95
 
0.95 0.05
 
 
  
0.20 0.80  0.6
 

λDSD (R) =    (3.4)
0.20 0.80  0.4
 

 
0.25 0.75
 
 
 
0.30 0.70
= (0.41, 0.59, 0.44, 0.44, 0.45, 0.46)T (3.5)

Completing similar calculations for the other four detectors yields:

λDKD (R) = (0.05, 0.97, 0.95, 0.82, 0.95, 0.99)T (3.6)

λDRD (R) = (0.32, 0.38, 0.64, 0.36, 0.46, 0.40)T (3.7)

λDW D (R) = (0.52, 0.80, 0.80, 0.15, 0.85, 0.85)T (3.8)

λDBD (R) = (0.95, 0.90, 0.85, 0.95, 0.20, 0.90)T (3.9)

Multiplying yields:

λ(R) = α(0.0032, 0.1566, 0.1819, 0.0185, 0.0334, 0.1394)T

In this somewhat unlikely case, because the road material has the greater
likelihood, the ML approach would choose it.

This sampling method can be used on-line to perturb materials. However,


assuming then that we are not interested in the underlying belief in each mate-
rial (only which has the maximum likelihood), we can create a new distribution
from which we can sample directly to perturb the material. The idea is that we
determine for each material, the probability of another material being detected
64

Table 3.2: Distribution resulting from offline sampling procedure.


True Detected Material
Material Accuracy Sky Sand Road Water Bldg Bkgrd
Sky 78% 0.78204 0.00951 0.008457 0.060073 0.005972 0.133948
sand 73% 0.005474 0.726594 0.030825 0.034847 0.015689 0.186571
road 62% 0.011089 0.04848 0.620218 0.041971 0.039691 0.238551
water 62% 0.080188 0.048161 0.034062 0.617168 0.008486 0.211935
building 57% 0.012262 0.058007 0.073957 0.034174 0.573262 0.248338

in its place. We achieve this off-line by simply running the Material Perturbation
Algorithm repeatedly, keeping tallies of each material detected and normalizing.

Using a given set of five detector characteristics (see Appendix B), we find the
distributions in Table 3.2.

Note that using an ML framework, the detectors are not as good as first
appeared. The most common detector error is for none of the detectors to fire (or
to fire with low belief), causing the region not to be detected at all.

Because we assume to know the orientation of the image, we could also incor-
porate an orientation model into the material fusion module, which prior work
suggests yields a substantial gain in accuracy (11% in [40]).

We note our underlying assumption that each detector’s CPM is independent,


which may not be true in practice: what may cause the one detector to fire may
be correlated with what may cause another detector to fire.

2) Not using orientation info in the detectors gives lousy detection overall, so
we can expect better when we use it. However, we can just interpret these results
as what we would get in the case of bad detectors!

3.2.4 Background Regions

We define a background region as any substantial region not detectable or not of


interest to us (e.g., people in a beach scene). We do not include the small borders
65

between materials. Background regions can be treated like other material regions,
their probability of occurrence and their spatial relationships with other regions
can be calculated. It is not yet clear whether spatial relationships involving the
background would be helpful.

When our framework includes faulty detectors, there is always a chance that
true materials of interest will not be detected by any detector (missed regions),
and would therefore be considered background, or that background regions could
falsely be detected as materials of interest (false positives). We currently account
for misses by ignoring (by not instantiating) the background and any relationships
involving it.

We require two additional extensions to the initial generative model in MASSES:

1. Model the probability of background regions and their relationships with


the materials of interest.

2. Generate background regions (even if we don’t use them during inference).

3. Allow them to be perturbed to become materials of interest (false positives).

4. Modify the conditional probabilities to make the model more robust to er-
rors. For instance, 0s in the matrix should be changed to a small .

A possible side-effect of modeling the background regions is that once we hy-


pothesize a scene type, we can attempt to infer the identity of unknown back-
ground regions.
66

4 Experimental Results

We have hypothesized that spatial relationships can be used to distinguish certain


scene types and that they can mitigate the effects of faulty detectors. We present
in this section our experiments and results.

4.1 Best-case Detection in MASSES

As described, the generative model in MASSES is a Bayesian network with con-


ditional probabilities based on our training data. We used MASSES to generate
a number of scenes and their relationships. We then classify them using two
Bayesian networks, the one with spatial information used to generate the scenes
and another in which the spatial relationship nodes are removed.

1. B -scenes always contain water, sky, and sand/gravel. The sky is above or
far above the other materials, and the water tends to lie above the sand,
although it occasionally can lie to the side or below the sand (accounting
for certain less-likely perspectives).

2. L-scenes always contain water and sky, and often contain sand/gravel. How-
ever, the sand tends to lie at the horizon line, above the water.
67

3. U -scenes contain buildings and sky, and often contain roads.

We have modeled these three simulated scenes after beach, lake/ocean, and
urban scenes. They are similar to their real-life counterparts, except that we
model the land at the horizon as sand/gravel (since the horizon may be confused
with these in images). The adjectives “occasionally”, and “often” in the definitions
above are quantified by the conditional probability distributions in the Bayesian
network from which we are sampling.

We based the CPMs for the the B- and U-scenes on those learned from data
(120 beach scenes and 100 urban scenes), but only using a subset of the materials
present: sky, water, sand, building, and road. We hand-modeled the L-scene
CPMs ourselves. The scene priors were chosen to be uniform.

We expect that the B- and L-scenes will be difficult to discriminate using


materials alone, since they often contain the same materials. However, the location
of the sand is usually different, causing differences in their spatial layouts. The
U-scene should be classified correctly with or without spatial information. We
have designed it for use in later experiments involving faulty detectors.

Accuracy on 100,000 scenes was 89.94% without spatial information and 96.99%
with spatial information, a gain of 7%. The absolute performance gain is not
important; what is important is that the hypothesis was verified under the as-
sumption of best-case material detection.

Data Size Material Only Spatial Information Added


100,000 89.94% 96.99%

Table 4.1: MASSES with best-case material detection: Accuracy with and without
spatial information.

Success Spatial relationships were most helpful in discriminating L-scenes con-


taining sand, water, and sky from B-scenes. Scenes of this type were classified
68

correctly (and with high confidence) if the sand region was between the sky and
the water regions.

Baseline Success The urban scenes could be discriminated from the other
scene types, using material presence alone. In our simplified model, buildings were
present if and only if the image was an U-scene, so this was expected. Furthermore,
L-scenes not containing sand were correctly classified as well, without spatial
relationships.

Continued failure What scenes escaped even spatial relations? First, with
small probability, the sand region was below the water region in some L-scenes.
These were mistaken as B-scenes. Second, when the sand and water regions
were next to each other, the two scenes were often confused. In these cases,
the final likelihoods (and hence the scene type) were determined by the spatial
relations between each region and the sky region. In B-scenes, sand tends to lie
below the sky and water far below the sky, whereas in L-scenes, the opposite
is true. However, the small fraction of images not holding to this relationship
were confused. Third, some simulated scenes contained seeming (at first sight)
conflicts, primarily due to the “far” modifier. Take, for instance, a scene with
water below the sky, and sand FAR below the sky. We could reason from this
that the sand would lie below the water, but often the sand would lie above the
water. Because we assumed each pairwise relationship to be independent of each
other relationship, this situation occurred.

4.2 Best-case Detection on Beach Photographs

We classified the 120 material-marked, beach scenes in our collection using the
same single-level Bayesian networks used in MASSES. As stated earlier, the
69

MASSES net is designed to distinguish between the B-, L-, and U-scenes.

Dataset Size Material Only Spatial Information Added


116
120 100% 120
= 96.67%

Table 4.2: MASSES with best-case material detection: Accuracy with and without
spatial information.

These results are not surprising, given our definitions. We had defined beach
scenes (see Table 3.1) as ones with at least 5% of the image occupied by each
of water, sand, and sky regions, with no regard for the spatial relationships be-
tween these regions. Therefore, we expect accuracy to be 100% without spatial
information.

Furthermore, we are not surprised that accuracy went down when spatial in-
formation was added to the inference process. Upon observation of the images,
three contained sand above water, due to camera location (e.g., on a boat), and
could be classified as L-scenes. They all contain large water regions and small
sand regions.

Figure 4.1: Images incorrectly classified due to spatial relationships.

The last image on the right, contains sand beside the water (which alone is
ambiguous), but the sky closer to the sand causes a higher belief in lake. The
rationale for calling it a beach scene is that the sand region dominates the water
region, having nothing to do with spatial relationships.
70

4.3 Faulty Detection on Beach Photographs

Before the inference stage, we perturb the results according to the method de-
scribed in 3.2.3. Note that the detector model is weak, so we can expect a large
percentage of the materials to be misclassified. Results are shown in Figure 4.3.

Dataset Size Material Only Spatial Information Added


110 108
120 120
= 91.67% 120
= 90.00%

Table 4.3: MASSES with faulty material detection: Accuracy with and without
spatial information.

The 10 images missed when the material-only Bayesian network was used all
contained undetected sand. These fall into two categories.

1. Eight images contain sky, water, and background, and are thus classified as
L-scenes (Figure 4.2).

Figure 4.2: Images incorrectly classified using faulty detectors and no spatial
relationships. The actual materials are shown the top row; the detected materials
are shown below each.

2. The other two images mis-detected a building as well and were thus classified
as U-scenes (Figure 4.3).
71

Figure 4.3: Images incorrectly classified using faulty detectors.

The 12 images missed when the spatial relationships were added to the Bayesian
network included the 10 missed above. In addition, spatial relationships caused
an incorrect classification in two cases.

1. Spatial relationships were unable to rescue the two images above in which
the sand was missed and a building was falsely detected; they were still
classified as U-scenes. In the first image, spatial relationships should be
able to catch the mis-detected regions: the building and road regions are
besides each other, and the building is far below the sky regions, both of
which are unlikely. However, the belief in the U-scene still dominated that in
the other scenes. Mis-detection in the second image was such that it would
be indistinguishable from a normal beach scene.

2. Spatial relationships were unable to rescue the eight images containing sky,
water, and background: they are still classified as L-scenes.

3. Two additional images were misclassified (Figure 4.4). Both contain a region
misclassified as sand present above the water region, causing them to be
classified as L-scenes. In the first image, a large sand region was undetected,
and the small remaining sand region was above the water, leading to a higher
belief in an L-scene than a B-scene. The second image continued to be
misclassified as an L-scene, due to camera location (as above).
72

Figure 4.4: Images incorrectly classified using faulty detectors.

We reiterate that in the current model of faulty detectors, we threshold the


combined material beliefs and replace the true material with the material most
likely to be detected. That the underlying material beliefs are not used in inference
is a severe limitation that must be addressed. Take, for example, a true sand region
that has 49% belief in sand and 51% belief in background. By only allowing binary
belief, we call it background and eliminate all chance that it could still be sand,
even if the spatial relationships imply that would fit the model better as a sand
region. We do plan to overcome this and other limitations in the thesis, as we
now describe in Chapter 5.
73

5 Proposed Research

Automatically determining the semantic classification of images is useful, among


other things, for context-sensitive image enhancement and image cataloging. As
we reviewed in Chapter 2, most current systems use a standard pattern recognition
approach, treating images as high-dimensional vectors of low-level features. Due
to advances in robust classifiers such as Support Vector Machines, these systems
have achieved some success in limited domains.

However, the “semantic gap” between low-level pixel representations of images


and the high-level understanding of them we seek is still wide. We believe that
more expressive features, such as the output of semantic object and material
detectors, are needed in order to bridge this gap and thus advance the state of
the art in scene classification.

While semantic detectors are not perfect, they continue to increase in speed
and accuracy; they are mature enough now to be reliably useful. We have access
to state-of-the-art detectors for blue and mixed sky, grass, skin, and faces, as well
as less-reliable detectors for water and snow fields, still under development [39;
38; 67]. Current research, still in its infancy, into related frameworks that rely
on high-level features like these is promising [37; 44; 48; 70]. However, no scene
classification system makes use of the spatial relationships between the objects
74

detected.

We have chosen to use a probabilistic graphical model to perform inference,


because our input and desired output is probabilistic and we can learning the
model parameters from real data.

Experience developing the Bayesian tree that served as our initial inference
mechanism was encouraging, but we found the Bayesian tree to be not expressive
enough to handle evidence from faulty detectors when that evidence corresponds
to individual regions and their spatial relationships.

With this in mind, our central claim is as follows: Spatial modeling of semantic
objects and materials can be used to disambiguate certain scene types as well as
mitigate the effects of faulty detectors. Furthermore, an appropriate probabilistic
inference mechanism must be developed to handle the loose spatial structure found
in real images.

5.1 Research Plan

Our research plan can be described succinctly:

• Identify the probabilistic inference mechanism, either Markov Random Fields


or Bayesian networks, most appropriate for modeling scenes for semantic
classification.

• Extend the chosen framework, because neither fits our needs exactly.

• Measure the efficacy both of our features and of our inference mechanism
by conducting experiments both on simulated and real images.

We define the input and output of our system, then elaborate on our research
plan in the following sections.
75

System Input: Fused detector belief maps We assume the existence of a


bank of relevant object and material detectors (e.g., sky, water, sand) each of
which produces a probabilistic belief map. In the belief map, each material region
is labeled with the corresponding belief in that material: 1 = highest confidence
that region is the material down to 0 = no confidence that region is the material.

We also assume the existence of a mechanism for fusing the material beliefs
of each region and for adjudicating between discrepancies in the segmentation
induced by each map. Material detectors and the associated fusion mechanism
are documented in the literature [40; 39].

A simplified input to the system is shown in Figure 5.1

Bel(sky) = 0.78
Bel(background) = 0.12
Bel(water) = 0.07
...

BEACH
Bel(Beach) = 0.73
Bel(sand) = 0.68 Bel(Urban) = 0.04
Bel(rock) = 0.22 ...
Bel(bkgrd)
... = 0.10
Bel(water)= 0.53
Bel(sky) = 0.36
...

Figure 5.1: Classifier input (labeled regions) and output (classification and confi-
dence).

System Output: Semantic label and confidence We seek first to output


one of a small (e.g., less than 12 ) set of discrete labels for the scene classification.
As two examples, in our simulated experiments, we used B-, L-, and U-scenes, and
in an outdoor scene classifier, we might use beach, lake, urban, field, mountain,
picnic, and forest.

We would also like to output a belief in at least the most likely scene type
(if not each scene type). The probabilistic framework allows us to output these
beliefs, which can be thresholded to label a scene as “unknown”.
76

5.1.1 Why a New Inference Mechanism?

As we have seen, the generative model (single-level Bayes net) of our MASSES
simulator generates materials, not individual regions. Under best-case material
detection, this is not problematic. However, when materials are falsely detected,
it is limited. Consider the case shown in Figure 5.2. Initially, we have the re-
lationships (1) sky Abovewater , (2) sand Abovewater , and (3) sky Abovesand . Consider
the case where the water is mis-detected as sky. Now the relationships are (1)
sky Abovesky , (2) sand Abovesky , and (3) sky Abovesand . Relation (1) can be safely ig-
nored if we consider it was a sky region broken into two pieces. However, relations
(2) and (3) are contradictory.

Sky Sky

Sand Sand
Water (Sky)

Figure 5.2: An image in which a region is mis-detected, creating contradictory


spatial relationships in the material-based inference scheme.

It is not clear how to handle conflicting evidence. Splitting belief within a single
relationship node is not correct, and adding multiple nodes to capture multiple
binary relationships between two materials is ad hoc.

The limitation seems to stem from the current system’s architecture, which
only encodes materials and not individual regions in the nodes. It is true that
some non-adjacent regions of the same material could be merged using model-
based or belief-based grouping [67] (and creating occlusion relationships in the
process). However, since multiple regions may have different associated material
beliefs, information is lost.
77

Because the underlying material beliefs should be useful in inference, it appears


better to encode each region (and the corresponding relations between them)
distinctly.

5.1.2 Bayes Nets Vs. MRFs

It is clear that the single-level Bayesian network is not expressive enough to encode
our knowledge of the world. We have identified two potential frameworks to
explore: general-topology Bayesian networks and Markov Random Fields.

General Bayesian Networks

If the network topology is allowed to the deviate from a strict tree-structure,


Bayesian networks are a natural framework to encode our knowledge. Specifically,
we must model individual regions to incorporate material beliefs. Figure 5.3 shows
an example of such a framework. Note that we cannot simply replace the material
nodes with region nodes; doing so would preclude the expression of knowledge such
as ”there is no sky present in the image”.

Class

Materials Material
Spatial Rels

Regions Region
Spatial Rels

Figure 5.3: Proposed DAG Bayesian network. Note that separate material and
region layers are needed.

The network shown uses a two-level Bayesian network, in which a level corre-
sponds to the materials present and a level corresponds to the individual regions.
78

The two levels are combined using a fuzzy-OR logical relation: for example, if any
region is detected as sky, then the sky node is activated.

However, we have some concerns with this network. First, while the approx-
imate inference algorithms discussed in Section 2.4.1 are designed for networks
with loops, it is unclear how well they would perform on dense graphs such as
this. Pearl hints that both the clustering and conditioning algorithms are tractable
only for sparse graphs. We need to investigate the stochastic simulation and belief
propagation algorithms further.

Second, it is not clear how to quantify the links connecting the region and
material nodes, because they represent logical relationships of variables with mul-
tiple hypotheses. Formal mechanisms such as the noisy-OR gate are designed for
binary variables.

Furthermore, given the debate between the relative merits of logic and proba-
bility [14; 50], mixing the two by adding logical constraints to a probability model
raises red flags in our minds due to technical details.

Markov Random Fields

Markov Random Fields are an attractive alternative to Bayesian networks. The


lattice structure typically used models spatial relationships directly. Furthermore,
two-layer MRFs (Figure 2.2) were designed to infer interpretations of underlying
scenes from uncertain evidence. For this reason, they have been applied with suc-
cess on low-level image-processing problems such as restoration [23], segmentation,
and depth map integration [16].

One approach to using MRFs for scene classification would be to create an


array of scene-specific MRFs and run them all in parallel. (While seemingly inef-
ficient, the number of scenes of interest is fairly small). We then classify the image
with the scene corresponding to the MRF producing the region configuration with
79

the highest posterior probability.

For each scene-specific MRF, the top level of the MRF corresponds to the
combined material beliefs in each region. The bottom-level corresponds to the
scene model, which is given by a distribution of ”typical” material configurations
for that scene.

Some reasons why the MRF is attractive include:

1. An available code base. We are nearing completion of a basic implemen-


tation of MRFs with inference using the belief propagation algorithm [7].
I could then extend this framework (as discussed in the next section) to
accommodate the scene classification problem.

2. Ease of representing multiple regions of the same material.

3. Direct use of detector belief values.

4. Solution to a related problem, that of spatial- and scene-context-aware ma-


terial detection, because the configuration with the smallest energy corre-
sponds to the most likely material labeling.

5.1.3 Extend the Chosen Framework

Neither MRFs nor Bayesian networks fit our needs exactly. We briefly describe
shortcomings in each framework and our initial ideas for extending them.

Extensions to Bayesian Networks

1. We argued that material regions must be encoded directly as nodes in the


Bayesian network. However, the number of regions is unknown in advance.
This would necessitate the ability to build the network dynamically during
the inference process. The alternative, that of creating more nodes than
80

should be necessary and only instantiating those actually present in the


image, is possible, yet seems ad hoc.

2. Our network requires the use of logical functions of multi-valued hypotheses.


However, the canonical models, such as noisy-OR gates, are designed for
binary-value hypotheses.

3. In cases where the detector evidence is weak (or evidence is conflicting, the
configurations are unlikely, or the scene is unmodeled), we would like to
label the image as “unknown”. This necessitates our modeling an “other”
category, which would require a large training set. Thie problem is also
much more difficult than forced-choice classification.

Extensions to Markov Random Fields

MRFs are typically used on a single scene model and generate a MAP estimate
of the scene given the observed image. To approximate the MAP, one does not
need to calculate the actual probabilities:

P (I|S)P (S)
argi max P (S|I) = (5.1)
P (I)
∝ P (I|S)P (S) (5.2)

We are proposing a novel use of MRFs: classification problems. We are moti-


vated by the successful use of Hidden Markov Models for speech recognition. In
speech recognition applications, HMM models for each word are run in parallel
and the probabilities are compared to decide which word is most likely.

Inspired by this successful use of HMMs, we envision a series of scene-specific


MRF models that can be run in parallel. Each MRF produces a MAP estimate
of the scene labeling and the probability of occurrence within that model (e.g.,
81

the probability of the image of that scene). We label the image with the class
of the MRF model producing the “Maximum of the MAP estimates” (MMAP).
Of course, cross-model comparison necessitates normalization, which is our first
challenge.

1. Because we are comparing multiple models, the probabilities produced by


each model must be normalized to the same scale. Computing the normal-
ization constant is intractable in general. We have two ideas for overcoming
this.

First, using a small number of nodes allows direct computation of the nor-
malization constant. Having a node correspond to each region or using a
small (e.g. 4 × 4) square lattice should suffice.

Second, we may be able to approximate the normalization constant by sam-


pling the scene models using a Monte Carlo method like Gibbs Sampling
2.4.2. This may prove to be too computationally expensive, but is worth
investigation.

2. While MRFs encode spatial relationships explicitly, there is no work in the


literature demonstrating their use in high level vision. Whether compati-
bility functions can be generalized beyond non-trivial, stationary, isotropic,
universal pixel-or edgel-level relationships is an interesting open problem
and an opportunity to make an intellectual contribution.

3. Specifically, the compatibility functions in MRFs are typically stationary;


ours depend on the location in the image. For example, in a beach scene, sky
next to sky may be very compatible on the top of the image, but completely
incompatible at the bottom of the image.

4. Because they operate on the pixel level, typical MRFs for vision are con-
structed on regular lattices. Our high-level vision problem starts with seg-
82

mented regions of irregular shape. Two options seem possible:

(a) Encode irregular regions and spatial relationships using a general graph
structure. The literature shows few departures from the lattice struc-
ture, except for use in hypertext link modeling [11] and object recog-
nition [19]

Our approach would be similar to the latter usage, since Felzenszwalb


and Huttenlocher (Section 2.4.2) were modeling spatial relationships
between object parts. However, scene configurations are less con-
strained than object configurations, so the spatial probability model
must be more general than the multivariate Gaussians they used.

(b) Overlay a lattice on the regions, but constrain belief propagation so


as to remain consistent with the segmentation given by the detectors.
The open question is how this will affect the theoretical properties of
probabilities or energy minimization. One option might be to propagate
as usual, and then reconcile the lowest-energy configuration with the
segmentation map at the end.

5.1.4 Analyze the Effect of Detector Quality

As we argued in Chapter 3, one benefit of working in simulation is that we are not


dependent on the actual detectors used. We plan to investigate the monotonic
function between classification performance and material detector performance.

5.1.5 Evaluate Explicit Spatial Relationships

In our experiments, we demonstrated that modeling spatial relationships can help


distinguish between certain types of scenes. Our goal here is to isolate the value
83

added by explicitly modeling the relationships. We compare the proposed prob-


abilistic framework with a framework which uses semantic features, but models
spatial information implicitly. We will implement a pattern recognition classifier
(e.g., a Support Vector Machine) and comparing to chosen inference mechanism.
For example, the SVM classifier with block-based belief features, encodes spatial
relationships implicitly (as position within the feature vector).

5.1.6 Evaluate Semantic Features

In order to validate our claims about semantic features and spatial relationships,
we must compare the efficacy of our system with a system using low-level features
and pattern classification techniques. We plan to use a benchmark set of outdoor
scene images.

5.1.7 Generalize to Other Scene Classes

Our proposed framework is not specific to outdoor scenes; the probabilities of the
presence of materials and the spatial relationships between them are learned from
data. We plan to experiment with other object-centric scene types which have
spatial structure.

One example is outdoor scenes dominated by a main-subject (e.g., picnic,


backyards). Since the main subject in typical consumer images is a person, adding
a person detector could handle these cases. The only extension to the scene
model would be to include a model of background (non-people, non-object) regions
including statistics of their spatial relationships with the modeled objects.

A second, more challenging example is indoor scenes. Office environments are


dominated by desks and people. Living room scenes tend to contain people on
sofas. Conference rooms contain an arrangement of a number of people around a
84

table. While the spatial relationships vary more with scenes in which the image
is captured at close range, this will test the generalizability of our methods.

5.1.8 Explore Potential Long-term Directions

We have identified a number of potential dimensions along which we can expand


our thesis.

• Temporal modeling: photographs do not come from a vacuum. Scenes occur


in loosely-connected groupings of events, like a comic strip or “storyboard”.
Exploring these groupings probabilistically gives a third, temporal dimen-
sion to our modeling.

• Exploring the use of camera metadata (timestamps, focal distance, flash


firing) to improve classification. This evidence is probabilistic in nature and
must be incorporated appropriately.

• Exploring the combination of low- or mid-level (e.g. color, texture, depth


cue, other elements of the “spatial envelope” [47]) and semantic features:
framework, control, cost.

Some researchers, in particular Oliva and Torralba [47], argue that human
observers can recover scene identity even when objects are unrecognizable
in isolation and thus propose the “scene-centered” features discussed in Sec-
tion 2.2.3. This appears to be true in certain cases, but I would agree with
Oliva and Torralba’s assessment that their approach is complementary to
approaches like the proposed one which are based on semantic objects. An
interesting direction would be to begin the process of uniting the two com-
plementary approaches.
85

• Comparison of the two schemes in Section 2.3.1 (AVS and Weighted Walk-
through) for computing qualitative relations in order to achieve a balance
between computational cost and perceptual integrity.

5.2 Issues Not Addressed in This Thesis

Our experiments revealed some ambiguity in the distinction between real beach
and lake photos. There is more to semantic understanding than materials and
spatial relations. For example, if I take a picture while standing on the beach
looking out over the water, and no sand is in the image, we think of it as a lake
scene. If a large sand region dominates the foreground, then we consider it to be
a beach scene. If the camera is positioned above the water and its field of view
contains the beach, it may be classified as either: if I am standing ten feet from
the shore, then it would be a beach semantically; if I am on a boat three miles
from the shore, then it would be a lake semantically. In this case, our system
classifies each as a lake, because sand appears above water in the image.

One limitation of our system in real classification is that we are considering


material presence and spatial relationships only. Relative size of regions also
appears to be helpful [44].

5.3 Research Schedule

The proposed research schedule includes the following:

• Compare Bayes Net and Markov Random Fields (2 months).

• Implement chosen framework (2 months).

• Performance analysis on natural images (1 semester).


86

• Evaluate semantic features by implementing low-level framework and com-


paring to proposed system (2 months).

• Evaluate explicit modeling of spatial relationships by implementing pattern


recognition classifier and comparing to proposed framework (2 months).

• Temporal context (or other appropriate) extension (1 semester).

• Write dissertation (1 semester).

Accounting for one summer working in industry, the anticipated defense date
is May, 2005.
87

6 Acknowledgments

This research was supported by a grant from Eastman Kodak Company, by the
NSF under grant number EIA-0080124, and by the Department of Education
(GAANN) under grant number P200A000306.
88

Bibliography

[1] Dana H Ballard. An Introduction to Natural Computation. MIT Press, Cam-


bridge, MA, 1997.

[2] Dana H Ballard and Christopher M Brown. Computer Vision. Prentice-Hall,


Englewood Cliffs, NJ, 1982.

[3] J. Batlle, A Casals, J Freixenet, and J Mart. A review on strategies for


recognizing natural objects in colour images of outdoor scenes. Image and
Vision Computing, 18(6-7):515–530, May 2000.

[4] D.C. Becalick. Natural scene classification that avoids a segmentation stage.
Technical report, ECE Department, Imperial College of Science, Technology,
and Medicine, 1996.

[5] S. Berretti, A. Del Bimbo, and E. Vicario. Spatial arrangement of color in


retrieval by visual similarity. Pattern Recognition, 35:1661–1674, 2001.

[6] Matthew Boutell, Jiebo Luo, and Christopher Brown. Review of the state of
the art in semantic scene classification. Technical Report 799, University of
Rochester, Rochester, NY, December 2002.

[7] Matthew Boutell and Brandon Sanders. Markov random fields: An overview.
Technical report, University of Rochester, Rochester, NY, 2003. in prepara-
tion.
89

[8] Christopher J. C. Burges. A tutorial on support vector machines for pattern


recognition. Data Minging and Knowledge Discovery, 2(2):121–167, 1998.

[9] C. Carson, S. Belongie, H Greenspan, and J. Malik. Recognition of images in


large databases using a learning framework. Technical Report 97-939, U.C.
Berkeley, 1997.

[10] C. Carson, S. Thomas, M. Belongie, J.M. Hellerstein, and J. Malik. Blob-


world: a system for region-based image indexing and retrieval. In Third Intl.
Conf. on Visual Information Systems. Springer-Verlag, June 1999.

[11] S. Chakrabarti, B. Dom, and P Indyk. Enhanced hypertext categorization


using hyperlinks. In ACM SIGMOD, 1998.

[12] Peng Chang and John Krumm. Object recognition woth color cooccurance
histograms. In IEEE Conference on Computer Vision and Pattern Recogni-
tion, Fort Collins, CO, June 23-25 1999.

[13] Eugene Charniak. Bayesian networks without tears. AI Magazine, 12(4):50–


63, April 1991.

[14] P. Cheeseman. An inquiry into computer understanding (with discussion).


Computational Intelligence, 4:57–142, 1988. Further discussion appears in
6:179–192.

[15] Rama Chellappa and Anil Jain, editors. Markov Random Fields: Theory and
Application. Academic Press, San Diego, CA, 1993.

[16] Paul Chou. The Theory and Practice of Bayesian Image Labeling. PhD thesis,
University of Rochester, Rochester, NY, 1988.

[17] Robert Cowell. Advanced inference in bayesian networks. In Michael I.


Jordan, editor, Learning in Graphical Models, volume 89 of NATO Science
Series D. Kluwer Academic Publishers, 1998.
90

[18] R. Duda, R. Hart, and D. Stork. Pattern Classification. John Wiley and
Sons, Inc., New York, 2nd edition, 2001.

[19] Pedro F. Felzenszwalb and Daniel P. Huttenlocher. Efficient matching of


pictorial structures. In IEEE Computer Vision and Pattern Recognition,
pages 66–73, 2000.

[20] Forsyth and Ponce. Computer Vision: A Modern Approach. Prentice Hall,
2002.

[21] D. Forsyth, J. L. Mundy, A. Zisserman, C. Coelho, A. Heller, and C. Rothwell.


Invariant descriptors for 3-d object recognition and pose. IEEE Transactions
on Pattern Analysis and Machine Intelligence, 13:971–991, October 1991.

[22] W.T. Freeman, E.C. Pasztor, and O.T. Carmichael. Learning low-level vision.
International Journal of Computer Vision, 40(1):24–57, October 2000.

[23] Stuart Geman and Donald Geman. Stochastic relaxation, gibbs distribu-
tions, and the bayesian restoration of images. IEEE Transactions on Pattern
Analysis and Machine Intelligence, 6(6):721–741, November 1984.

[24] A. Hauptmann and M. Smith. Text, speech, and vision for video segmenta-
tion: The informedia project. In AAAI Symposium on Computational Models
for Integrating Language and Vision, Fall 1995.

[25] M. Henrion. Propagation of uncertainty by probabilistic logic sampling in


bayes’ networks. In J. F. Lemmer and L. N. Kanal, editors, Uncertainty in
Artificial Intelligence, pages 149–164. Elsevier Science Publishers B.V., 1988.

[26] P. Hong, T. Huang, and R. Wang. Learning patterns from images by com-
bining soft decisions and hard decisions. In IEEE Conference on Computer
Vision and Pattern Recognition, volume 1, pages 78–83, Hilton Head, SC,
June 13-15 2000.
91

[27] Q. Iqbal and J. K. Aggarwal. Combining structure, color and texture for
image retrieval: A performance evaluation. In International Conference on
Pattern Recognition (ICPR), August 2002.

[28] A.K. Jain, R.P.W. Duin, and Jianchang Mao. Statistical pattern recognition:
a review. IEEE PAMI, 22(1):4–37, January 2000.

[29] Michael I. Jordan, editor. Learning in graphical models. NATO Advanced


Study Institute on Learning in Graphical Models. Kluwer Academic Pub-
lisher, Boston, 1998.

[30] Ross Kindermann and J. Laurie Snell. Markov Random Fields and Their
Applications, volume 1. American Mathematical Society, Providence, RI,
1980.

[31] T. Kohonen. Improved versions of learning vector quantization. In Proc. In-


ternational Joint Conference on Neural Networks, pages 545–550, San Diego,
June 1990.

[32] T. Kohonen, J. Kangas, J. Laaksonen, and K. Torkkola. Lvq pak: A program


package for the correct application of learning vector quantization algorithms.
In Proc. International Joint Conference on Neural Networks, volume 1, pages
725–730, Baltimore, June 1992.

[33] S. Lauritzen. Propagation of probabilities, means, and variances in mixed


graphical association models. Journal of the American Statistical Association,
87(420):1098–1108, 1992.

[34] P. Lipson. Context and Configuration-Based Scene Classification. PhD thesis,


MIT, Cambridge, MA, 1996.

[35] P. Lipson, E. Grimson, and P. Sinha. Configuration based scene classification


and image indexing, 1997.
92

[36] Y. Lu, C. Hu, X. Zhu, H. J. Zhang, and Q. Yang. A unified framework for
semantics and feature based relevance feedback in image retrieval systems.
In ACM Multimedia Conference, Los Angeles, October 2000.

[37] J. Luo and A. Savakis. Indoor vs. outdoor classification of consumer pho-
tographs using low-level and semantic features. In IEEE International Con-
ference on Image Processing, Thessaloniki, Greece, October 2001.

[38] Jiebo Luo and Stephen Etz. A physics-motivated approach to detecting sky in
photographs. In International Conference on Pattern Recognition, volume 1,
Quebec City, QC, Canada, August 11 - 15 2002.

[39] Jiebo Luo, Amit Singhal, and Weiyu Zhu. Natural object detection in outdoor
scenes based on probabilistic spatial context models. In ICME, Baltimore,
MD, July 2003.

[40] Jiebo Luo, Amit Singhal, and Weiyu Zhu. Towards holistic scene content
classification using spatial context-aware scene models. In IEEE Conference
on Computer Vision and Pattern Recognition, Madison, WI, June 2003.

[41] D. J. C. Mackay. Introduction to monte carlo methods. In Michael I. Jordan,


editor, Learning in Graphical Models, volume 89 of NATO Science Series.
Kluwer Academic Publishers, 1998.

[42] Stephane Mallat. A theory for multiresolution signal decomposition: The


wavelet representation. IEEE Transactions on Pattern Analysis and Machine
Intelligence, 11(7):674–692, July 1989.

[43] David Marr. Vision - A Computational Investigation into the Human Rep-
resentation and Processing of VIsual Information. Freeman, San Francisco,
1982.
93

[44] Philippe Mulhem, Wee Kheng Leow, and Yoong Keok Lee. Fuzzy conceptual
graphs for matching images of natural scenes. In IJCAI, pages 1397–1404,
2001.

[45] R. M. Neal. Probabilistic inference using markov chain monte carlo methods.
Technical Report CRG-TR-93-1, Dept. of Computer Science, University of
Toronto, 1993.

[46] A. Oliva and A. Torralba. Modeling the shape of the scene: A holistic repre-
sentation of the spatial envelope. International Journal of Computer Vision,
42(3):145–175, 2001.

[47] A. Oliva and A Torralba. Scene-centered description from spatial envelope


properties. In 2nd Workshop on Biologically Motivated Computer Vision,
Lecture Notes in Computer Science, Tuebingen, Germany, 2002.

[48] S. Paek and S.-F. Chang. A knowledge engineering approach for image clas-
sification based on probabilistic reasoning systems. In IEEE International
Conference on Multimedia and Expo. (ICME-2000), volume II, pages 1133–
1136, New York City, NY, Jul 30-Aug 2 2000.

[49] G. Pass, R. Zabih, and J. Miller. Comparing images using color coherence
vectors. In Proceedings of the 4th ACM International Conference on Multi-
media, pages 65–73, Boston, Massachusetts, November 1996.

[50] J. Pearl. Probabilistic Reasoning in Intelligent Systems. Morgan Kaufmann


Publishers, 1988.

[51] T. Randen and J.M. Husoy. Filtering for texture classification. IEEE Trans-
actions on Pattern Analysis and Machine Intelligence, 21(4):291–310, April
1999.
94

[52] A. Rao, R. Srihari, L. Zhu, and A. Zhang. A theory for measuring the
complexity of image databases. IEEE Transactions on Multimedia, 4(2):160–
173, June 2002.

[53] Rajesh Rao and Dana Ballard. Efficient encoding of natural time varying
images produces oriented space-time receptive fields. Technical Report 97.4,
University of Rochester, Rochester, NY, August 1997.

[54] A. Ratan and W.E.L. Grimson. Training templates for scene classification
using a few examples. In Proceedings of IEEE Content Based Access of Image
and Video Libraries, San Juan, 1997.

[55] Terry Regier and Laura Carlson. Grounding spatial language in perception:
An empirical and computational investigation. Journal of Experimental Psy-
chology: General, 130(2):273–298, 2001.

[56] R. D. Rimey and C. M. Brown. Control of selective perception using bayes


nets and decision theory. International Journal of Computer Vision, Special
Issue on Active Vision, 1994.

[57] Raymond D. Rimey. Control of Selective Perception using Bayes Nets and De-
cision Theory. PhD thesis, Computer Science Dept., U. Rochester, Rochester,
NY, December 1993.

[58] S. Roweis and L. Saul. Nonlinear dimensionality reduction by locally linear


embedding. Science, 290(5500):2323–2326, December 22 2000.

[59] Cordelia Schmid. Constructing models for content-based image retrieval. In


Computer Vision and Pattern Recognition, Kauai, Hawaii USA, December
2001.
95

[60] Henry Schneiderman and Takeo Kanade. A statistical method for 3d object
detection applied to faces and cars. In IEEE Conference on Computer Vision
and Pattern Recognition, 2000.

[61] B. Scholkopf, C. Burges, and A. Smola. Advances in Kernel Methods: Support


Vector Learning. MIT Press, Cambridge, MA, 1999.

[62] Andrea Selinger. Analysis and Applications of Feature-Based Object Recog-


nition. PhD thesis, University of Rochester, Rochester, NY, 2001.

[63] Andrea Selinger and Randal C. Nelson. A perceptual grouping hierarchy


for appearance-based 3d object recognition. Computer Vision and Image
Understanding, 76(1):83–92, October 1999.

[64] Satoshi Semba, Masayoshi Shimizu, and Shoji Suzuki. Skin color based light-
ness correction method for digital color images. In PICS, pages 399–402,
2001.

[65] Navid Serrano, Andreas Savakis, and Jiebo Luo. A computationally efficient
approach to indoor/outdoor scene classification. In International Conference
on Pattern Recognition, September 2002.

[66] A. Singhal. Bayesian Evidence Combination for Region Labeling. PhD thesis,
University of Rochester, Rochester, NY, 2001.

[67] Amit Singhal and Jiebo Luo. Hybrid approach to classifying sky regions in
natural images. In Proceedings of the SPIE, volume 5022, July 2003.

[68] A.W.M. Smeulders, M. Worring, S. Santini, A. Gupta, and R. Jain. Content-


based image retrieval at the end of early years. IEEE Transactions on Pattern
Analysis and Machine Intelligence, 22(12):1349–1380, December 2000.
96

[69] J. R. Smith and C.-S. Li. Image classification and querying using composite
region templates. Computer Vision and Image Understanding, 75(1/2):165 –
174, July/August 1999.

[70] Y. Song and A. Zhang. Analyzing scenery images by monotonic tree. ACM
Multimedia Systems Journal, 2002.

[71] Milan Sonka, Vaclav Hlavac, and Roger Boyle. Image Processing, Analysis,
and Machine Vision. Brooks/Cole Publishing, Pacific Grove, CA, 2 edition,
1999.

[72] M. J. Swain and D. H. Ballard. Color indexing. International Journal of


Computer Vision, 7(1), 1991.

[73] Martin Szummer and Rosalind W. Picard. Indoor-outdoor image classifica-


tion. In IEEE International Workshop on Content-based Access of Image and
Video Databases, Bombay, India, 1998.

[74] A. Torralba and P. Sinha. Recognizing indoor scenes. Technical Report AI


Memo 2001-015, CBCL Memo 202, MIT, July 2001.

[75] A. Torralba and P. Sinha. Statistical context priming for object detection.
In Proceedings of the International Conference on Computer Vision, pages
763–770, Vancouver, Canada, 2001.

[76] A. Vailaya. Semantic Classification in Image Databases. PhD thesis, Michi-


gan State University, East Lansing, MI, 2000.

[77] A. Vailaya, M. Figueiredo, A. Jain, and H.J. Zhang. Content-based hierarchi-


cal classification of vacation images. In Proc. IEEE Multimedia Systems ’99
(International Conference on Multimedia Computing and Systems), Florence,
Italy, June 1999.
97

[78] A. Vailaya, A. K. Jain, and H.-J. Zhang. On image classification: City images
vs. landscapes. Pattern Recognition, 31:1921–1936, December 1998.

[79] A. Vailaya, H.J. Zhang, and A. Jain. Automatic image orientation detection.
In Proc. IEEE International Conference on Image Processing, Kobe, Japan,
October 1999.

[80] J. Wang, J. Li, and G Wiederhold. Simplicity: Semantics-sensitive integrated


matching for picture libraries. IEEE Transactions on Pattern Analysis and
Machine Intelligence, 23(9):947–963, 2001.

[81] Y. Weiss and W. T. Freeman. Belief propagation and revision in networks


with loops. Technical Report 1616, MIT, Cambridge, MA, 1998.

[82] J.S. Yedidia, W.T. Freeman, and Y. Weiss. Understanding belief propagation
and its generalizations. International Joint Conference on Artificial Intelli-
gence (IJCAI), Distinguished Presentations, August 2001.

[83] H. Yu and W.E.L. Grimson. Combining configurational and statistical ap-


proaches in image retrieval. In 2nd IEEE Pacific-Rim Conference on Multi-
media, Beijing, October 2001.
98

A Natural Scene Statistics

We give the spatial relationship counts calculated from the 220 labeled beach and
urban scenes.

The row is the landmark object and the column is the trajector object, so
for instance, Above(f oliage, sky) = 24 from the first table means that sky was
located above foliage in 24 beach images. Blue patch (of sky) was incorporated
into sky; we do not differentiate based on cloudy/clear sky.

BEACH

Relation above
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 1 0 0 0 0 0 0 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 1 0 0 0 0 0 0 0
foliage 24 0 0 0 2 0 8 0 0 1 1
water 97 0 0 12 0 0 5 8 1 7 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 16 0 0 5 51 0 0 9 0 4 10
rock 11 0 0 4 3 0 3 0 0 0 0
road 1 0 0 0 1 0 0 0 0 1 0
building 13 0 0 1 0 0 0 1 0 0 0
crowd 3 0 0 2 6 0 0 0 0 2 0
99

Relation far above


sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 1 0 0 0 0 0 0 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 3 0 0 0 2 0 1 0 0 0 0
foliage 9 0 0 0 6 0 0 0 0 3 1
water 23 0 1 17 0 0 0 2 0 6 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 104 0 1 24 32 0 0 7 1 11 1
rock 11 0 0 2 4 0 1 0 0 1 0
road 4 0 0 1 2 0 0 0 0 1 0
building 0 0 0 0 0 0 0 0 0 0 0
crowd 15 0 0 3 1 0 0 3 0 5 0
Relation beside
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 9 0 0 0 2 0 2 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 1 0 0 1 0 1 0 0
foliage 9 0 1 0 7 0 7 2 3 0 0
water 0 0 0 7 0 0 32 7 1 2 11
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 1 7 32 0 0 4 4 0 7
rock 2 0 0 2 7 0 4 0 1 0 0
road 0 0 1 3 1 0 4 1 0 0 0
building 2 0 0 0 2 0 0 0 0 0 0
crowd 0 0 0 0 11 0 7 0 0 0 0
Relation far below
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 3 9 23 0 104 11 4 0 15
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 0 1 0 1 0 0 0 0
foliage 1 0 0 0 17 0 24 2 1 0 3
water 0 0 2 6 0 0 32 4 2 0 1
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 1 0 0 0 0 1 0 0 0
rock 0 0 0 0 2 0 7 0 0 0 3
road 0 0 0 0 0 0 1 0 0 0 0
building 0 0 0 3 6 0 11 1 1 0 5
crowd 0 0 0 1 0 0 1 0 0 0 0
100

Relation below
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 24 97 0 16 11 1 13 3
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 0 0 0 0 0 0 0 0
foliage 1 0 1 0 12 0 5 4 0 1 2
water 0 0 0 2 0 0 51 3 1 0 6
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 0 8 5 0 0 3 0 0 0
rock 0 0 0 0 8 0 9 0 0 1 0
road 0 0 0 0 1 0 0 0 0 0 0
building 0 0 0 1 7 0 4 0 1 0 2
crowd 0 0 0 1 0 0 10 0 0 0 0

Individual material occurrence in 120 images:


sky 120
bluePatch 0
grass 3
foliage 44
water 120
snow 0
sand 120
rock 24
road 5
building 15
crowd 18

URBAN

Relation above
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 1 0 0 0 0 0 0 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 6 0 0 0 0 0 1 0
foliage 9 0 0 0 1 0 0 0 0 22 0
water 6 0 0 0 0 0 0 0 0 14 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 0 0 0 0 0 0 0 0 0
rock 0 0 0 0 0 0 0 0 0 0 0
road 1 0 1 4 2 0 0 0 0 19 0
building 77 0 0 2 1 0 0 1 0 0 0
crowd 0 0 0 0 0 0 0 0 0 0 0
101

Relation far above


sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 0 0 0 0 0 0 0 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 9 0 0 1 3 0 0 0 0 7 0
foliage 23 0 0 0 0 0 0 0 0 2 0
water 20 0 0 3 0 0 0 0 0 8 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 1 0 0 0 0 0 0 0 0 1 0
rock 3 0 0 1 0 0 0 0 0 2 0
road 33 0 0 12 2 0 0 0 0 16 0
building 1 0 0 1 0 0 0 0 0 0 0
crowd 0 0 0 0 0 0 0 0 0 0 0
Relation beside
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 5 1 0 0 0 0 19 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 0 0 0 1 0 3 1 0
foliage 5 0 0 0 1 0 1 0 4 11 0
water 1 0 0 1 0 0 1 2 0 5 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 1 1 1 0 0 0 1 0 0
rock 0 0 0 0 2 0 0 0 0 0 0
road 0 0 3 4 0 0 1 0 0 0 0
building 19 0 1 11 5 0 0 0 0 0 0
crowd 0 0 0 0 0 0 0 0 0 0 0
Relation far below
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 9 23 20 0 1 3 33 1 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 0 0 0 0 0 0 0 0
foliage 0 0 1 0 3 0 0 1 12 1 0
water 0 0 3 0 0 0 0 0 2 0 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 0 0 0 0 0 0 0 0 0
rock 0 0 0 0 0 0 0 0 0 0 0
road 0 0 0 0 0 0 0 0 0 0 0
building 0 0 7 2 8 0 1 2 16 0 0
crowd 0 0 0 0 0 0 0 0 0 0 0
102

Relation below
sky blu gra fol wat sno san roc roa bui cro
sky 0 0 0 9 6 0 0 0 1 77 0
bluePatch 0 0 0 0 0 0 0 0 0 0 0
grass 0 0 0 0 0 0 0 0 1 0 0
foliage 1 0 6 0 0 0 0 0 4 2 0
water 0 0 0 1 0 0 0 0 2 1 0
snow 0 0 0 0 0 0 0 0 0 0 0
sand 0 0 0 0 0 0 0 0 0 0 0
rock 0 0 0 0 0 0 0 0 0 1 0
road 0 0 0 0 0 0 0 0 0 0 0
building 0 0 1 22 14 0 0 0 19 0 0
crowd 0 0 0 0 0 0 0 0 0 0 0

Individual material occurrence in 100 images:


sky 97
bluePatch 0
grass 9
foliage 38
water 28
snow 0
sand 1
rock 3
road 35
building 100
crowd 0
103

B Detector Characteristics

Each row gives the probability of the detector firing and not firing on each true
material.

node-name Sky_Detector
0.95 0.05 Sky
0.03 0.97 Sand
0.05 0.95 Road
0.18 0.82 Water
0.05 0.95 Building
0.01 0.99 Background

node-name Sand_Detector
0.05 0.95 Sky
0.95 0.05 Sand
0.20 0.80 Road
0.20 0.80 Water
0.25 0.75 Building
0.30 0.70 Background

node-name Road_Detector
0.05 0.95 Sky
0.20 0.80 Sand
0.85 0.15 Road
0.15 0.85 Water
0.40 0.60 Building
0.25 0.75 Background

node-name Water_Detector
104

0.48 0.52 Sky


0.20 0.80 Sand
0.20 0.80 Road
0.85 0.15 Water
0.15 0.85 Building
0.15 0.85 Background

node-name Building_Detector
0.05 0.95 Sky
0.10 0.90 Sand
0.15 0.85 Road
0.05 0.95 Water
0.80 0.20 Building
0.10 0.90 Background

You might also like