Está en la página 1de 206

Analysis and Reduction of Moire Patterns in

Scanned Halftone Pictures


by
Xiangdong Liu
Dissertation submitted to the faculty of
Virginia Polytechnic Institute and State University
in partial fulllment of the requirements for the degree of
DOCTOR OF PHILOSOPHY
in
Computer Science
APPROVED:
Dr. Roger Ehrich, Chairman
Dr. Lynn Abbott
Dr. Lenwood Heath
Dr. James Campbell
Dr. Cliord Shaer
May, 1996
Blacksburg, Virginia
Key Words: Moire Pattern, Halftone, Scanning, Image Processing, Fourier Analysis
Analysis and Reduction of Moire Patterns in Scanned
Halftone Pictures
by
Xiangdong Liu
Committee Chairman: Dr. Roger Ehrich
Department of Computer Science
(ABSTRACT)
ACKNOWLEDGEMENTS
I am truly indebted to Dr. Roger W. Ehrich, my major professor, for his patience and
guidance throughout my career at Virginia Tech over the past ve years. He spent much
time discussing various problems with me, and also provided me with a wonderful working
environment which included his own computer equipment and library.
I want to express my gratitude to Dr. Abbott, Dr. Campbell, Dr. Heath, and Dr. Shaer,
for serving on my committee, and giving me advice and suggestions. Dr. Campbells satellite
imagery analysis project supported me nancially for almost two years.
My acknowledgment goes to Mr. Greg Degi at Hewlett-Packard for communicating with
me in the past year on scanning techniques. I also thank Dr. Lewis E. Franks at the
University of Massachusetts for enlightening me on spectrum analysis.
Randy Ribler and Siva Challa, two of my fellow graduate students who joined Virginia
Tech at about the same time as I did, and who are also graduating, gave me friendship.
They helped me on numerous occasions and on many matters.
Finally, this thesis is dedicated to my wife, Lihan, and my daughter, Heather, for their
understanding and support. To Lihan, I was away at school for half of her married life, and
to Heather, I was only to be seen once a month or less. The rst words Heather learned to
murmur were looking for daddy. I am also grateful to my parents, who shaped me as I
am, and always gave me encouragement.
iii
TABLE OF CONTENTS
1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Problem statement and contribution of this dissertation . . . . . . . . . . . 3
2 Moire patterns 4
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Moire patterns in sampled halftones . . . . . . . . . . . . . . . . . . . . . . 5
3 A Brief Review of Halftoning 9
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.1.1 The contrast sensitivity of the human visual system . . . . . . . . . 10
3.1.2 Overview of halftoning techniques . . . . . . . . . . . . . . . . . . . 11
3.2 Clustered-dot halftoning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2.1 Analog screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2.2 Clustered-dot ordered dither . . . . . . . . . . . . . . . . . . . . . . 16
3.2.3 Template-dot halftone . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.4 Comparison of clustered-dot halftoning methods . . . . . . . . . . . 19
4 Review of Previous Work 21
4.1 Research work underlying this research . . . . . . . . . . . . . . . . . . . . . 21
4.2 Research work on related problems . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 Research work addressing similar problems . . . . . . . . . . . . . . . . . . 23
5 Characterization of Moire in Sampled Halftones 26
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
iv
CONTENTS
5.2 Notation and denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.3 Interpretation of the Fourier spectrum of a picture . . . . . . . . . . . . . . 27
5.4 The Fourier spectrum of an ideally sampled uniform halftone . . . . . . . . 28
5.4.1 The Fourier spectrum of a uniform halftone . . . . . . . . . . . . . . 28
5.4.2 The Fourier spectrum of an ideally sampled uniform halftone . . . . 32
5.5 The Fourier spectrum of a practically sampled uniform halftone . . . . . . . 37
5.6 The Fourier spectrum of a practically sampled general halftone . . . . . . . 41
5.7 The geometric shape of a moire pattern . . . . . . . . . . . . . . . . . . . . 49
5.7.1 Sum of sinusoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.7.2 The geometric shape of a moire pattern in 1D . . . . . . . . . . . . . 55
5.7.3 The geometric shape of a moire pattern in 2D . . . . . . . . . . . . . 58
5.8 The half period phase reversal of a moire pattern . . . . . . . . . . . . . . . 61
5.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6 Case Study: Moire Patterns in Scanned Halftones by a Commercial Desk-
top Scanner 70
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.2 The image data processing procedure by the ScanJet IIc . . . . . . . . . . . 70
6.2.1 The optical sampling of the ScanJet IIc . . . . . . . . . . . . . . . . 71
6.2.2 Resolution reduction of an optically sampled row of pixels . . . . . . 72
6.3 The Fourier spectrum of scanned halftones from the ScanJet IIc . . . . . . . 75
6.3.1 The Fourier spectrum of one row of an image scanned at 150 dpi . . 78
6.3.2 The Fourier spectrum of an image scanned at 150 dpi . . . . . . . . 83
6.3.3 The Fourier spectrum of an image scanned at arbitrary resolutions . 89
6.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
7 Estimating the halftone lattice from a sampled halftone image 99
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7.2 The leakage phenomenon of the Fourier transform . . . . . . . . . . . . . . 101
v
CONTENTS
7.3 Halftone lattice estimation from a uniformly sampled halftone . . . . . . . . 106
7.3.1 Estimating the halftone frequencies . . . . . . . . . . . . . . . . . . . 107
7.3.2 Estimating the phases of the halftone lattice . . . . . . . . . . . . . 110
7.3.3 Measuring the accuracy of halftone lattice estimation . . . . . . . . 112
7.4 Halftone lattice estimation from a non-uniformly sampled halftone . . . . . 116
7.5 Summary and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8 Restoration of Sampled Halftones with Moire Patterns 121
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.2 Measurement of the quality of moire pattern suppression . . . . . . . . . . . 123
8.2.1 Subjective measures . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.2.2 A quantitative measure . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.2.3 Comparison with the dual-sampling method . . . . . . . . . . . . . . 126
8.3 Notch ltering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.4 Simulating the halftoning and the sampling processes . . . . . . . . . . . . . 153
8.4.1 Overview of the method . . . . . . . . . . . . . . . . . . . . . . . . . 153
8.4.2 Generating the simulated sampled halftone image . . . . . . . . . . . 155
8.4.3 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
8.5 The Relaxation method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
8.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
9 Prevention of Moire Patterns When Sampling a Halftone 180
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
9.2 Partial inverse Fourier transform . . . . . . . . . . . . . . . . . . . . . . . . 180
10 Concluding Remarks 187
10.1 The nature of a moire pattern . . . . . . . . . . . . . . . . . . . . . . . . . . 187
10.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
vi
LIST OF FIGURES
1.1 Moire pattern in a scanned halftone from Mercedes . . . . . . . . . . . . . . 2
2.1 A moire pattern formed by overlapping two bar patterns . . . . . . . . . . . 5
2.2 The formation of moire pattern . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1 Normalized human visual system contrast sensitivity against spatial frequency 11
3.2 Mathematical model of analog screening . . . . . . . . . . . . . . . . . . . . 15
3.3 Examples of digital screens . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4 Graphical description of template-dot halftoning . . . . . . . . . . . . . . . 19
5.1 Halftone representation of a uniform gray tone . . . . . . . . . . . . . . . . 29
5.2 The layout of the Fourier spectrum of a uniform halftone . . . . . . . . . . 30
5.3 A circular dot, a square dot, and their Fourier spectra . . . . . . . . . . . . 31
5.4 Two diamond-shaped dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.5 The Nyquist band for sampling frequencies s and t . . . . . . . . . . . . . . 33
5.6 The layout of the Fourier spectrum of a sampled uniform halftone . . . . . 35
5.7 The layout of the Fourier spectrum of a typical, practically sampled uniform
halftone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.8 House, a test halftone picture . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.9 A 512 512 pixel portion of an image scanned from the House halftone
picture at 400 dpi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.10 The Fourier spectrum of the scanned halftone House . . . . . . . . . . . . 46
5.11 Resampled scanned halftone House and its Fourier spectrum . . . . . . . 48
5.12 The beat pattern of two sinusoids . . . . . . . . . . . . . . . . . . . . . . . . 50
5.13 The bar pattern formed by the sum of two sinusoids . . . . . . . . . . . . . 51
vii
LIST OF FIGURES
5.14 Vector sum of amplitudes of sinusoids . . . . . . . . . . . . . . . . . . . . . 52
5.15 The Fourier spectrum of the sum of four sinusoids . . . . . . . . . . . . . . 56
5.16 1D moire: density pattern, contrast pattern, combined pattern . . . . . . . 57
5.17 Frequency component distribution of (a) an aliasing and (b) a non-aliasing
frequency clusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.18 2D moire: density patterns and contrast patterns (part (d) enlarged to show
detail) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.19 The phase reversal of a moire pattern . . . . . . . . . . . . . . . . . . . . . 62
5.20 1D halftone representing a uniform gray . . . . . . . . . . . . . . . . . . . . 63
5.21 The Fourier spectra of two uniform halftones of density 0.35 and 0.75 . . . 63
5.22 The Fourier spectra of two sampled uniform halftones: (a) density 0.75 and
(b) density 0.35 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.23 A halftone of two levels of gray and its Fourier spectrum . . . . . . . . . . . 67
6.1 The image data processing procedure of ScanJet IIc in horizontal direction
(a) and in vertical direction (b) . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.2 IBRM steps to produce an output row at 150 dpi . . . . . . . . . . . . . . . 74
6.3 House halftone scanned by ScanJet IIc at 150 dpi . . . . . . . . . . . . . 76
6.4 The Fourier power spectrum of the scanned halftone in Figure 6.3 . . . . . 77
6.5 Equivalent low-pass ltering and sampling models for obtaining the 800 dpi
interpolated row of pixels . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.6 Non-uniform sampling and uniform display . . . . . . . . . . . . . . . . . . 81
6.7 Magnitude of envelopes for sampling replicas in G
vs
(u) . . . . . . . . . . . . 83
6.8 Distribution of signicant frequency components of (a) the low-pass ltered
input halftone and (b) the scanned halftone . . . . . . . . . . . . . . . . . . 85
6.9 The low frequency moire pattern in the scanned halftone in Figure 6.3 . . . 86
6.10 The image in Figure 6.3 high-pass ltered to retain only frequency compo-
nents at the halftone frequency or higher . . . . . . . . . . . . . . . . . . . . 87
viii
LIST OF FIGURES
6.11 A portion of the high-pass ltered image in Figure 6.10 enlarged eight times 88
6.12 The distribution and strength of replicas created by IBRM sampling . . . . 91
6.13 A portion of the 85 dpi House test halftone scanned at 200 dpi by a ScanJet
IIc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.14 A portion of the 85 dpi house test halftone scanned at 250 dpi by a ScanJet
IIc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.1 The halftone lattice relative to the sampling lattice (a) and the halftone
frequencies in the frequency domain (b) . . . . . . . . . . . . . . . . . . . . 100
7.2 Leakage of DFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3 Leakage of Fourier transform in 2D . . . . . . . . . . . . . . . . . . . . . . . 105
7.4 Frequencies used in the test images for frequency and phase estimation . . . 114
8.1 The 400 dpi scanned House halftone by a ScanJet IIc subsampled uniformly
at 150 dpi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
8.2 The image produced by the dual-sampling method using the images in Fig-
ure 5.11(a) and Figure 8.1 as input . . . . . . . . . . . . . . . . . . . . . . . 128
8.3 The Fourier spectrum of the image in Figure 8.2 . . . . . . . . . . . . . . . 128
8.4 The dierence between the image in Figure 5.11(a) and the image in Figure 8.2129
8.5 The 400 dpi uniformly scanned House halftone subsampled at 150 dpi using
the ScanJet IIc subsampling procedure . . . . . . . . . . . . . . . . . . . . . 130
8.6 The 400 dpi uniformly scanned House halftone subsampled at 160 dpi using
the ScanJet IIc subsampling procedure . . . . . . . . . . . . . . . . . . . . . 131
8.7 The image produced by the dual-sampling method using the images in Fig-
ure 8.5 and Figure 8.6 as input . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.8 The dierence between the images in Figure 8.5 and Figure 8.7 . . . . . . . 133
8.9 Two ideal notch ltered images of the sampled halftone in Figure 5.11(a) . 135
8.10 The dierence between (a) the image in Figure 5.11(a) and the image in
Figure 8.9(a) and (b) the image in Figure 5.11(a) and the image in Figure 8.9(b)136
ix
LIST OF FIGURES
8.11 The Fourier spectrum of the image in Figure 8.9(a) . . . . . . . . . . . . . . 137
8.12 The sampled halftone in Figure 5.11(a) notch ltered using blending inter-
polation in 3 3 pixel neighborhoods . . . . . . . . . . . . . . . . . . . . . . 138
8.13 The dierence between the image in Figure 5.11(a) and the image in Figure 8.12139
8.14 An ideal notch ltered image of the scanned halftone in Figure 6.3 . . . . . 140
8.15 The Fourier spectrum of the image in Figure 8.14 . . . . . . . . . . . . . . . 142
8.16 The dierence between the images in Figure 6.3 and Figure 8.14 . . . . . . 143
8.17 A 512512 pixel portion of the Mercedes image in Figure 1.1 . . . . . . . . 144
8.18 The Fourier spectrum of the Mercedes image in Figure 8.17 . . . . . . . . . 145
8.19 The Mercedes image ideal notch ltered using 11 pixel neighborhoods . . 146
8.20 The Mercedes image ideal notch ltered using 33 pixel neighborhoods . . 147
8.21 The Mercedes image notch ltered using blending smoothing in 33 pixel
neighborhoods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
8.22 The dierence between the images in Figure 8.17 and Figure 8.19 . . . . . . 149
8.23 The dierence between the images in Figure 8.17 and Figure 8.20 . . . . . . 150
8.24 The dierence between the images in Figure 8.17 and Figure 8.21 . . . . . . 151
8.25 The Fourier spectrum of the ideal-notch ltered Mercedes image in Figure 8.19152
8.26 The relative position of a CCD sensor and the halftone dots in its vicinity . 157
8.27 The position of a sampling point expressed in terms of the halftone lattice
vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
8.28 Two sets of parameters to dene a halftone dot lattice . . . . . . . . . . . . 160
8.29 A large halftone dot in a parallelogram cell . . . . . . . . . . . . . . . . . . 160
8.30 The pixel value to dot radius correspondence curve for a square halftone cell
that has a unit side length . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8.31 The CCD motion path(s) and the nominal sampling point for one output
pixel of the ScanJet IIc for (a) 300400 dpi resolution and (b) 150297 dpi
resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
8.32 The region that inuences the value of the pixel at a sampling point . . . . 164
x
LIST OF FIGURES
8.33 The mathematical model for the low-pass ltering by the optics of the ScanJet
IIc: in frequency domain (a) and in spatial domain (b) . . . . . . . . . . . . 166
8.34 The spatial domain low-pass ltering model by the optics and moving carriage
of the ScanJet IIc at 400 dpi scan resolution: (a) 3D plot, (b) contour plot . 166
8.35 The spatial domain low-pass ltering model by the optics and moving carriage
of the ScanJet IIc at 150 dpi scan resolution: (a) 3D plot, (b) contour plot . 167
8.36 The output image obtained by applying the simulation method on the image
in Figure 5.11(a) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
8.37 The dierence between the images in Figure 5.11(a) and Figure 8.36 . . . . 169
8.38 The output image obtained by applying the simulation method on the image
in Figure 6.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.39 The dierence between the images in Figure 6.3 and Figure 8.38 . . . . . . 172
8.40 A sampling point and the dots in or near its neighborhood . . . . . . . . . . 175
8.41 The overlap of the neighborhood (x
I
) of a halftone dot at x
I
and the neigh-
borhoods of the sampling points that fall in (x
I
) . . . . . . . . . . . . . . 176
9.1 Distribution of signicant frequency components of (a) a halftone scanned
uniformly with sucient low-pass ltering and (b) an output image from the
partial inverse Fourier transform method taking the scanned halftone as input182
9.2 The 342 342 pixel center portion of the DFT of the 400 dpi, 1024 1024
pixel master image used in Chapter 8 . . . . . . . . . . . . . . . . . . . . . 183
9.3 The 133
1
3
dpi output image obtained by inverse transforming the DFT in
Figure 9.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
9.4 The 133
1
3
dpi output image obtained by low-pass ltering the image in Fig-
ure 9.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
9.5 The 150 dpi output image obtained by low-pass ltering the 384 384 pixel
center portion of the master image DFT followed by inverse transform . . . 186
xi
LIST OF TABLES
7.1 Errors of the frequency and phase estimates on test images that are sums of
random noise and two sinusoids . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.2 Errors of the frequency and phase estimates on test images that are sums of
the house image and two sinusoids . . . . . . . . . . . . . . . . . . . . . . . 115
8.1 Quality metric values for the moire-reduced image in Figure 8.2 . . . . . . . 127
8.2 Quality metric values for the moire-reduced image in Figure 8.7 . . . . . . . 133
8.3 Quality metric values for the moire-reduced images in Figure 8.9 and Figure 8.12138
8.4 Neighborhood sizes used in notch ltering to obtain the image in Figure 8.14 141
8.5 Quality metric values for the moire-reduced image in Figure 8.36 . . . . . . 168
8.6 Quality metric values for the moire-reduced image in Figure 8.38 . . . . . . 170
xii
Chapter 1
Introduction
1.1 Motivation
As digital multimedia publishing has increased in importance, the inclusion of scanned
graphics has become widespread. All pictures from printed materials, such as newspapers,
magazines, and journals, are halftoned; when such a halftone is scanned, a noisy pattern
called a moire pattern often appears superimposed on the scanned image. For example,
Figure 1.1 shows a scanned halftone from an auto parts inventory of Mercedes ([MB93]),
one they have been unable to restore. Most available viewing software causes a stronger
moire pattern when a scanned halftone image is reduced in size.
There are two simple but not necessarily practical ways to avoid or reduce such noisy
patterns: scan the original graphics instead of their halftoned prints, or scan and store the
halftoned graphics at very high resolution. In reality, the original photos may not be easy
to obtain or may no longer exist. In fact, there are also large image archives of scanned
halftone pictures, such as the Mercedes inventory, for which neither the originals nor the
halftones are available.
High scanning resolution may result in unacceptably large data volume, which increases
the cost of transmitting, processing, and storing the data. Even at a normal scanning
resolution of 300 dots per inch (dpi), a large amount of data is already generated. The
human eye has a resolving capability of no more than about 250 dpi at normal viewing
distance of 12 inches [PN92, RL94]. A halftone picture does not contain ne image detail
that is beyond this resolution. Most often a halftone picture carries only information no
ner than about 75 dpi. High scanning resolution therefore does not capture more image
1
CHAPTER 1. INTRODUCTION
Figure 1.1: Moire pattern in a scanned halftone from Mercedes
information. For example, as will be discussed in Chapter 3, a 3 3 inch picture printed
with a 133 dpi halftone (magazine quality) contains graphical content that can largely be
represented by 400 400 gray pixels. When scanned at 300 dpi, however, the picture
yields an image of 900 900 pixels. Furthermore, when an image so scanned is displayed,
unnecessary detail of the halftone dot structure may dominate the graphic content. Lastly,
even though there exist fairly mature algorithms that achieve high compression ratios on
images whose neighboring pixel dependencies are high, they do not perform well on high
resolution scanned halftone pictures because the latter lack pixel-level correlation. Thus,
while the amount of data scanned is large, compression is also dicult.
2
CHAPTER 1. INTRODUCTION
1.2 Problem statement and contribution of this dissertation
From the discussion of Section 1.1 three questions that are both of practical and of
intellectual value can be posed:
1. How is a moire pattern formed when a halftone picture is scanned or the scanned
image is resampled?
2. How can a moire pattern in a scanned halftone image be removed or reduced?
3. How should a halftone picture be scanned so that the resultant image does not have a
moire pattern and its resolution is comparable to the bandwidth of the image content?
This dissertation attempts to answer these three questions. Specically, a unied theory
of the formation of moire patterns in sampled halftones is established. Part of the theory is
expanded from previous researchers work. The theory developed is applied to and veried
by scanned halftones from a commercial desktop scanner. Algorithms are developed to
explore the feasibility of reducing moire patterns in scanned halftones (image restoration).
Also developed is a technique to prevent moire patterns when a halftone picture is scanned
(moire pattern prevention).
This dissertation also attempts to lay the foundation for further research on the subject.
A moire pattern in a sampled halftone is a complex phenomenon, even under ideal sampling
models. The fact that most scanners are not ideal samplers adds complexity to the moire
patterns thus formed. This dissertation only provides a partial answer to the second question
posed above. Further research is still necessary. Some potential research directions that are
not explored in this dissertation are discussed.
3
Chapter 2
Moire patterns
2.1 Introduction
Moire is a French word; it is not a persons name. Websters New Collegiate Dictio-
nary, 1978 edition, explains that in addition to referring to an irregular wavy nish on a
fabric, moire or moire, also means an independent usu. shimmering pattern seen when
two geometrically regular patterns (as two sets of parallel lines or two halftone screens) are
superimposed esp. at an acute angle. The independent pattern is the result of the interfer-
ence between the two regular patterns. Figure 2.1 shows an example of a moire formed by
overlapping two bar patterns. Bryngdahl [Bry74] discusses the formation of a moire using
various numbers of gratings of dierent geometry. Moire phenomena can be seen in daily
life. For example, we often see a moire pattern when we look through two layers of the
mesh safety barrier covering a pedestrian bridge over a street. We also see a moire when
we overlap two pieces of window screen. In this dissertation, the spelling moire is adopted.
Techniques that create a moire pattern and use it to measure the deformation or surface
topology of an object are often used in optics and mechanics ([RR93, Section 2.4] and
[Mal92, Chapter 16]). For example, to test the deformation of an object under certain
conditions (e.g. strain), a periodic pattern or grating is printed on the surface of the object
before the prescribed conditions are applied. Another grating is then superimposed on the
rst grating. The superposition can be achieved by direct contact of the two gratings or by
overlapping the second grating with an image of the rst grating or by photographing both
gratings on the same lm. The deformation of the object also deforms the grating printed
on it. An analysis of the resultant, usually irregular, moire reveals the displacement of the
4
CHAPTER 2. MOIRE PATTERNS
Figure 2.1: A moire pattern formed by overlapping two bar patterns
grating on the object. A similar technique does not print a grating on the object. Rather
it shines a beam of light through a grating screen. The screen casts a periodic shadow on
the object. When an observer looks at the object through the screen from another angle,
the grating on the screen and its shadow on the object create a moire. This moire carries
information of the surface topology of the object being examined.
The above techniques are similar in nature to the traditional method of measuring the
atness of an optical surface by Newton rings [Mal92, Chapter 1]. However, in the Newton
ring method the fringes or Newton rings are the result of the interference of two beams
of light, rather than the interference of two explicit geometric patterns. Newton rings are
therefore not considered to be a moire pattern.
2.2 Moire patterns in sampled halftones
A moire pattern may occur when a halftone picture is sampled because the sampling
process overlaps two periodic entities, the halftone lattice and the sampling lattice. A
halftone picture consists of a lattice of halftone dots whose sizes reect the local density of
the graphical content. The sampling process itself is usually periodic because such sampling
is easy to implement in practice and captures the source signal most eciently on average.
In this dissertation, the terms moire and moire pattern are used interchangeably. When
5
CHAPTER 2. MOIRE PATTERNS
used without qualication they refer to the moire pattern in a sampled halftone.
A moire pattern often looks like a coarse grid of intersecting fringes superimposed on a
sampled image. Sometimes the moire fringes are more pronounced in one direction than in
the other. The fringe grid may be oriented in any direction.
In the image domain, a moire pattern can be understood as the interference between the
halftone lattice and the sampling lattice. To illustrate this interference, Figure 2.2 shows
a simplied example of a bar pattern (shaded rectangles) being sampled by a 2D array
of square sensors. The bar pattern represents a halftone that consists of a 1D array of
dark lines instead of a 2D lattice of dark dots to carry image information. Line h alftones
are indeed used in practice sometimes for special eect. Figure 2.2 (a) shows the relative
position of the sensor array to the halftone lines, and Figure 2.2 (b) shows the output of
the sensor array. In Figure 2.2 (b) a dark square means a sensor with a low intensity
reading. Notice the periodic moire pattern in Figure 2.2 (b). The dierent periods of the
halftone lines and the sensor array make the sensors periodically pick up dierent amounts
of black, even though the halftone actually represents a uniform of gray tone. Due to
the line halftone being used and the relative orientation between the halftone lines and the
sensor array, the moire pattern is one set of vertical fringes.
A moire pattern can also be studied in the frequency domain. In Chapter 5 we will
explain that a moire pattern results from the interference among strong frequency compo-
nents that are close to one another. Very often, this interference is due in part to aliasing.
Since a halftone picture theoretically has innitely wide bandwidth, sampling such a pic-
ture without prior low-pass ltering always results in aliasing. Since most of the energy
of a halftone picture concentrates at lower frequencies, as a rule of thumb, if the sampling
rate is eight times the frequency of the halftone dot lattice or higher, moire patterns are
negligible [Hua74].
As will be discussed in Chapter 6, commercial scanners usually sample an input picture
twice in a cascading manner. An input picture is rst scanned optoelectronically and
mechanically by a moving CCD array at one rate, and the scanned raw data is subsequently
6
CHAPTER 2. MOIRE PATTERNS
(b)
(a)
Figure 2.2: The formation of moire pattern
digitally subsampled to produce an image of a user-requested resolution. The scanner optics
can be so designed that it lters out frequencies higher than half the CCD scanning rate,
and there is virtually no aliasing in the scanned raw data. However, the digital subsampling
may introduce aliasing if it is not preceded by sucient low-pass ltering.
We need to address a fundamental question regarding the sampling th eorem and the
restoration of a sampled halftone bearing a moire pattern due in part to aliasing: Why is
it possible at all to correct the aliasing error in the form of a moire pattern, even if only
partially, while the sampling theorem states that an aliasing error cannot be corrected?
The answer is that the sampling theorem assumes that we do not have any additional
information about the sampled signal, other than the samples themselves. Hence given the
sampled signal we cannot recover the original when aliasing occurs. In our case, we know
more: the signal being sampled has a strong periodic component in the form of the halftone
lattice. Further, the spectrum of this signal contains not only the halftone frequency but
7
CHAPTER 2. MOIRE PATTERNS
also its harmonics. As long as the sampling frequency is higher than twice that of the
halftone lattice, we can determine the latter from the sampled signal and then deduce how
it and its harmonics are distributed and aliased. In other words, we can determine the
frequency of this type of aliasing error.
When a picture containing a periodic content is halftoned, a moire pattern may also
occur because the halftone lattice may interfere with the periodic content in the picture.
An example is the re-halftoning of a halftone picture. Although this is an interesting
problem, it is outside our current discussion. Much work has been done to reduce this kind
of moire ([Roe76], [All78], and [MSO90]).
Finally we compare the moire patterns used in metrology in optics and mechanics with
moire patterns in sampled halftones. In moire metrology, the researcher has control of what
gratings to use to create a moire that is easy to analyze. The two gratings are often chosen
to have similar frequencies to maximize the strength of moire so that the latter is easy to
measure. The geometry of the gratings may be a set of parallel lines, a lattice, or a set
of concentric circles. The width of the bars in a grating either is a constant or varies
in some simple manner (e.g. always one half of the bar spacing). The researcher can also
decide the relative oset between the two gratings. The analysis of a moire focuses on the
spatial distribution of the fringes. The objective is to deduce the deformation of one of
the two interfering gratings. In a sampled halftone the two interfering patterns are always
lattices. The researcher has no choice of the frequencies for the two lattices or their relative
oset. To determine the halftone lattice frequency, the sampling lattice frequency must be
at least twice as high as the halftone lattice frequency. The size of the dots on the halftone
lattice varies with the image content. To remove or reduce a moire pattern, both its spatial
position and its strength need to be taken into account. In lieu of the above comparison,
major eort is needed to make use of the knowledge of the moire pattern analysis from
metrology to address the moire pattern problem in sampled halftones.
8
Chapter 3
A Brief Review of Halftoning
3.1 Introduction
Halftoning refers to techniques that create the visual illusion of gray scale using a dot
pattern that has only two levels of gray. A normal printing process is binary in nature in
that it cannot adjust the density of ink for every point on the paper. Rather, it can only
either print ink on a spot or leave it blank. For white paper and black ink, the process
makes a spot either black or white. To print pictures with gray tones, halftoning must be
used. Halftone techniques can also be used in other binary output devices as well as devices
that can produce only a few gray levels, to display gray tone images.
Color pictures must also be printed in halftone. A color picture is decomposed into three
primary color bands, and each band halftoned. The three single color halftones are printed
in such a way that their graphical features are accurately aligned and superimposed. The
integration of the three colors by the human eye creates the perception of color. In practice,
a color picture is usually printed using four halftones, the three primary color halftones and
a gray halftone. The gray halftone provides the gray scale (or brightness) of a color graphic
more eciently than using a combination of the three primary colors.
In this research we focus on monochromatic halftones in order to develop the basic
theory and algorithms. Before discussing halftoning techniques in more detail, we need to
discuss relevant properties of the human visual system rst.
9
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
3.1.1 The contrast sensitivity of the human visual system
The contrast sensitivity of the human visual system is the capability of the latter to
detect the dierence in brightness between neighboring regions in a scene. A high sensitiv-
ity means the ability to distinguish small dierences in brightness. Human visual contrast
sensitivity is largely dependent upon the sizes of the neighboring regions in question. That
is, the sensitivity is a function of spatial frequency. This sensitivity also varies with the
background illumination; it achieves the highest values in a moderately illuminated en-
vironment such as normal sunshine.
Many psychophysical experiments have been conducted to determine how the human
visual contrast sensitivity varies with spatial frequency [MS74]. Most often used as test
scenes are bar patterns or gratings with dierent spatial frequencies and contrast. For
each frequency, gratings of dierent contrast are shown to human subjects to determine
the lowest contrast discernible. It turns out that the human visual contrast sensitivity also
varies with the orientation of the grating; it achieves the highest value when a grating is
horizontally or vertically oriented and achieves the lowest value when a grating is oriented
at 45 degrees from horizontal. Dierent results are obtained by dierent experimenters due
in part to dierent experimental conditions and assumptions. However, all the results show
that the human visual contrast sensitivity, as a function of spatial frequency, varies in a
curve similar to the one shown in Figure 3.1. The curve shows the normalized sensitivity
and is based on the data obtained by several experimenters that are cited in [MS74]. In
most experimental results the spatial frequency is expressed in terms of cycles per degree of
a subjects eld of view. This unit is translated to cycles per inch (cpi) at a normal viewing
distance of 12 inches in Figure 3.1. The peaks of the curves from dierent experimenters
range from about 10 cycles per inch to 50 cycles per inch with an average of about 20 cycles
per inch. The sensitivity drops rapidly at frequencies away from the peak frequency. At
the lower frequency end of the curve, the sensitivity is low due to a technicality: If the
frequency is zero, there is no grating and thus no contrast to observe. Figure 3.1 is by no
10
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
50 100 150 200 250
0.2
0.4
0.6
0.8
1
spatial frequency (in cpi)
c
o
n
t
r
a
s
t

s
e
n
s
i
t
i
v
i
t
y
Figure 3.1: Normalized human visual system contrast sensitivity against spatial frequency
means precise; it only shows the general behavior of the resolving capability of the human
visual system.
3.1.2 Overview of halftoning techniques
A halftoning method is a binary encoding method. The basic idea is to print black
points or groups of black points in such a way that the local point density is roughly equal
to the average gray value in the corresponding regions of the source picture. The printing is
controlled in such a ne fashion that the human eye cannot completely resolve the individual
printed points or individual groups of points. The printed picture then appears to have
continuous gray tone because of the spatial integration performed by the eye. The high
resolution of a printer that cannot be fully perceived by the human eye is used to create an
illusion of gray scale. This is in analogy to holograms which utilize the high resolution of
the recording medium to encode a depth dimension.
There are two broad categories of halftoning techniques: analog screening, which has
evolved since the 1850s, and digital halftoning, which started in the 1960s. Through one
and a half centuries of improvement, especially with the invention of the contact screen
in the 1940s, analog screening has matured and is the most widely used technique in
the printing industry. A halftone picture produced through analog screening consists of a
11
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
periodic lattice of h alftone dots of ink color. Each halftone dot can be considered to reside
in an area of the same shape and size of the printed picture. The area is called a h alftone
cell, and the size of the dot is proportional to the density of the input picture in the cell
area.
In the digital halftoning category, many algorithms have been developed. These algo-
rithms take as input a gray level image and produce a binary output image of the same
dimensions. Each binary pixel is reproduced as a point on the output device. The mapping
of the input gray level image to the binary output image is called a h alftone function. Usu-
ally, a normal gray level image is rst interpolated to a much larger size before being fed to
a halftoning algorithm to avoid losing too much information in the halftoning process. In
doing so the size of the printed picture can also be controlled. For example, to print a 3 by
3 inch halftone picture out of a 500 500 pixel gray level image on a 600 dpi laser printer,
the image is rst interpolated to 1800 1800 pixels. The interpolated gray level image is
then mapped, or halftoned, to an 18001800 pixel binary image. The binary image is then
reproduced on the printer.
In this dissertation, a h alftone (picture) means a halftone that is printed on paper. A
digital h alftone (image) means the binary digital image produced by a digital halftoning
algorithm from a gray scale digital image. A digital printer, such as a laser printer, takes
as input a digit halftone image and reproduces it as a halftone picture on paper. A scanned
h alftone (image) means the gray scale digital image scanned from a halftone picture by a
scanner. A sampled h alftone (image) means a gray scale digital image that is sampled, by
digital or analog means or a combination of both, from a halftone picture or from a digital
halftone image or from a scanned halftone image. A scanned halftone image is a sampled
halftone image.
Digital halftoning algorithms can be further classied into point-based algorithms and
block-based algorithms. A point-based algorithm determines the value of each pixel in the
output halftone image individually using the same procedure. To determine an output pixel,
the algorithm examines the corresponding pixel in the input gray level image and probably
12
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
also its neighboring pixels. This category includes the random dith er meth od and the error
diusion meth od. A block-based algorithm partitions an input gray level image and the
output halftone image into blocks. Each block is called a h alftone cell. The algorithm
examines the pixels in each cell of the input image collectively to determine the values of
pixels in the corresponding cell of the output image using the same procedure. A cell in
the output image may have all the 1-valued (black) pixels clustered together or have them
spread out in the cell. When all the 1-valued pixels are clustered together, they form a
clustered dot similar to that produced by the analog screening method. Otherwise the cell
contains dispersed dots. Two clustered-dot algorithms are most prominent: the clustered-dot
ordered dith er algorithm and the template-dot algorithm. Dispersed-dot algorithms include
the dispersed-dot ordered dith er algorithm and the blue noise mask algorithm.
Most digital printing devices (including most laser printers) have the adverse property
of non-linear reectance response, which means the reectance of a small printed region
varies with the distribution of the black pixels in the region, and so the density also varies
and is usually not equal to the ratio of the number of black pixels to the total number of
pixels in the region. If the black pixels are clustered, the density is more stable and is closer
to the correct ratio ([RL94, Section IV] and [Uli87, Sections 2.2-2.3]). For this reason the
clustered-dot ordered dither algorithm and the template-dot algorithm are better suited to
most available digital printing devices.
Because the template-dot algorithm is inferior to the clustered-dot ordered dither algo-
rithm in terms of picture quality and run-time eciency, it is rarely used in printing. Hence,
the clustered-dot ordered dither is the most widely used digital halftoning algorithm. We
will see in the next section that the clustered-dot ordered dither algorithm is actually the
digital simulation of the analog screening method. In the literature, the term h alftone
used with no restriction usually refers to the analog screening halftone or the clustered-dot
ordered dither halftone. Nevertheless, because the template-dot halftone is simple to an-
alyze and can sometimes be considered to be an approximation to the other two types of
halftones, it is often used to model the other two types when the recovery of image infor-
13
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
mation from a sampled halftone is considered ([OYT
+
86, FJ94]). In this dissertation we
use the term clustered-dot h alftone to mean the analog screening halftone, the clustered-dot
ordered dither halftone, or the template-dot halftone.
In the next section we describe the three clustered-dot halftone techniques which are
suitable to mass printing. Roetling [RL94], Ulichney [Uli87], and Jarvis, Judice, and Ninke
[JJN76] give more information about halftoning methods in general.
3.2 Clustered-dot halftoning
The analog screening, clustered-dot ordered dither, and template-dot techniques all
partition a halftone picture into a lattice of halftone cells and place a halftone dot in each
cell. To exploit the characteristics of the human visual contrast sensitivity, a lattice is
usually oriented at 45 degrees. The frequency of the lattice is called the h alftone frequency.
The highest frequency graphical information that can be rmly carried by such a halftone
is at about one half the halftone frequency. A halftone produced by the analog screening
method or the clustered-dot ordered dither method is able to carry some information at
frequencies higher than one half the halftone frequency, but there is no quantitative measure
in the literature. All three techniques share the same property of inherent periodicity. That
is, a picture printed through one of these techniques has fundamental periodic content that
is not present in the source graphic. This periodicity is the ultimate cause of a moire
pattern when the halftoned picture is sampled. This research targets halftone pictures
printed through these three techniques.
3.2.1 Analog screening
In this method, a h alftone screen that is a transparency (or lm) with periodically vary-
ing transmittance is placed in between a source picture and a high gamma photosensitive
medium. After exposure the sum of the optical density of the source picture and that of the
screen is recorded in essentially binary values on the medium. The photosensitive medium
14
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
-8 -6 -4 -2 2 4 6 8
Figure 3.2: Mathematical model of analog screening
is then developed and subsequently used to prepare the printing plate of a press printer.
The frequency of the screen is high (usually greater than 50 dpi) relative to the resolving
capability of the human eye.
The halftoning process can be mathematically modeled. The model is graphically illus-
trated in Figure 3.2 in one dimension. In the upper half of the gure is a unit amplitude
sinusoidal source picture added with a unit amplitude sawtooth screen. The vertical axis
represents density. Each tooth of the screen corresponds to one halftone cell. The sum
is thresholded at unit height, resulting in the binary output in the lower half of the gure.
Each rectangle in the gure represents a halftone dot. The halftone cells have a size of
two units in the horizontal direction, and the tick marks show the centers of the cells. The
halftone dots are generally not centered in their cells, conveying sub-cell information about
the source picture. For example, the source picture has a peak near 7. The dots in the
two nearest cells with centers at 8 and 6 are located close to each other to reect this
fact. Such dots are called partial dots, in contrast to full dots that are centered at their cells
and occur when the picture content is a constant.
In two dimensions, we may think of the halftoning as adding a halftone screen sur-
face consisting of an array of hills (or pyramids) to the source picture surface and then
15
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
thresholding the sum. Mathematically, let 0 f(x, y ) 1 be the source picture. Let
0 screen(x, y ) 1 be the halftone screen satisfying
screen(x, y ) =

m,n
screen
cell
(x mp, y np),
where p is the period of the screen in the x and the y directions, and screen
cell
(x, y) is one
period of the screen:
screen
cell
(x, y ) =
_
hill(x, y ), [x[, [y [ p/2;
0, otherwise.
hill(x, y ) represents the density distribution of the halftone screen in one cell. Dierent
screens have dierent distribution. The halftone picture h(x, y ) then is
h(x, y ) =
_
1, f(x, y ) +screen(x, y ) 1;
0, otherwise.
(3.1)
Eq. 3.1 is the halftone function. In each halftone cell a spot near the center (mp, np) is
blackened, forming a dot while the rest of the cell remains blank. The size of the dot is
roughly proportional to the average density of the input picture in the cell. The shape and
position of the dot in a cell depends upon the detail of the source picture in the cell and is
not xed.
3.2.2 Clustered-dot ordered dither
This method is the digital simulation of analog screening. Both the source picture and
the halftone screen are arrays of numbers, or digital images, of the same dimensions. The
output digital halftone image can be obtained by adding pixel-wise the digital source image
with the digital screen and thresholding the sum. In the ctional 1D case, the halftoning
process is still illustrated by Figure 3.2, and a screen is a 1D array of numbers that has
a periodic distribution pattern resembling a sawtooth function. In 2D, a screen array is
also periodic and consists of numbers that form a lattice of hills, each hill corresponding
to a halftone cell. Note that even though the method is conceptually block based, in
implementation it is a pixel-to-pixel operation.
16
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
8 7 9 11 12 10
2 1 6 17 18 13
3 4 5 16 15 14
8 7 9 11 12 10
2 1 6 17 18 13
3 4 5 16 15 14
8 7 9 11 12 10
(a)
2 1 6 17 18 13
3 4 5 16 15 14
8 7 9 11 12 10
2 1 6 17 18 13
3 4 5 16 15 14
2 12 5 3
7 14 15 9
10 16 13 6
4 8 11 1
(b)
Figure 3.3: Examples of digital screens
Eq. 3.1 can be written as
h(x, y ) =
_
1, f(x, y ) 1 screen(x, y );
0, otherwise,
and 1 screen(x, y ) is simply another screen that is shifted by half a cell from screen(x, y ).
Therefore, a simpler way to generate a halftone image is to threshold each source image pixel
against the corresponding halftone screen pixel, and thus leading to the following halftone
function:
h(x, y ) =
_
1, f(x, y ) screen(x, y );
0, otherwise,
(3.2)
The most common way to create a digital screen is to start with a small threshold
array of carefully chosen and arranged numbers. Then the threshold array is used to tile
the source image. The thresholding in Eq. 3.2 can be performed while the source image
is tiled. Figure 3.3 shows two examples of threshold arrays. In Figure 3.3(a) is a 4 4
pixel threshold array for a screen that is horizontally and vertically oriented. The threshold
array corresponds to a halftone cell and provides 17 levels of gray. If a binary image thus
obtained is printed on a 600 dpi printer, the resultant halftone picture is said to have a 150
dpi screen. The array elements shown are integers. For a source image whose pixel values
17
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
are between 0 and 1, the array elements should be scaled by a factor of
1
16
. In Figure 3.3(b)
is a 3 6 pixel threshold array replicated four times to show how it is used to tile an image
and form a 45 degree screen. The array corresponds to a screen whose halftone cell size is
3

2 3

2 pixels, and the array itself covers portions of four halftone cells. One halftone
cell is marked by the dashed square. If a binary image thus obtained is printed on a 600
dpi printer, the resultant halftone picture is said to have a
600
3

2
141 dpi, 45 degree screen
and has 19 levels of gray.
Since Eq. 3.2 can be performed in a tile by tile fashion, some authors call the mapping
of the source image to the binary halftone image in one tile of the threshold array or even
the threshold array itself the halftone function. Much research has been done on designing
the threshold arrays.
A halftone image can be obtained by thresholding, pixel by pixel, a gray level source
image against a uniformly distributed random noise or dith er. This method is called the
random dither method. Eq. 3.2 thresholds, pixel by pixel, the source image against a xed
distribution of values, hence the name ordered dith er method. If a threshold array does not
have all the elements with large values clustered together, the resultant halftone may have
more than one black spot in a halftone cell. Such halftoning is called the dispersed-dot
ordered dither method.
3.2.3 Template-dot halftone
This method is conceptually simple. A source image is partitioned into cells. The
average value of all pixels in a cell is used to match a predetermined dot whose size is
proportional to the average. The dot is copied into the corresponding cell of the output
image. Figure 3.4 shows in one dimension the same sinusoid in Figure 3.2 halftoned by
the template-dot method. In the gure the vertical dashed lines indicate boundaries of
the halftone cells. The size (width) of a dot is equal to the average of the sinusoid in
the corresponding cell. In a template-dot halftone, the shape and position of a dot is
predetermined. Such a dot is always a full dot. The method is simply an area sampling
18
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
-8 -6 -4 -2 2 4 6 8
Figure 3.4: Graphical description of template-dot halftoning
method and is also known as surface area modulation (SAM). Note that the predetermined
dots need not be clustered. However, in the literature the term template-dot usually implies
so. Also note that a template dot cannot be formed piece by piece as in the ordered dither
case. Therefore, the template-dot halftoning is not a pixel-to-pixel operation.
3.2.4 Comparison of clustered-dot halftoning methods
Each of analog screening, clustered-dot ordered dithering, and the template-dot method
produces a halftone picture that consists of a lattice of halftone dots whose size reects the
local density of the source image. When a single gray level source picture is halftoned by
any of these methods, the result is an array of full dots of the same size, shape, and relative
position. The halftone is periodic.
For a source image with ne detail (high frequency features), the template-dot method
ignores detail ner than a halftone cell and places full dots of the same size, shape, and
relative position in all cells having the same average density. In areas that the source image
has high frequency features (where density changes quickly) the periodicity of the resultant
halftone is weak because adjacent dots tend to have dierent sizes.
For the same source image with high frequency features, the analog screening method
and the clustered-dot ordered dither method place dots of not only dierent sizes but
also dierent shapes and relative positions in dierent cells. The within-cell detail of the
source are retained to some degree. Thus the resultant halftone looks sharper than the
19
CHAPTER 3. A BRIEF REVIEW OF HALFTONING
one produced by the template-dot. In high frequency feature areas, the periodicity of the
halftone is weaker than that of the template-dot halftone.
In summary, all three clustered-dot halftoning methods produce very similar, periodic
halftone dot structures in areas that a source image has relatively uniform density. The
three methods produce less periodic dots in areas that the source image has ne detail.
A scanning process is usually lossy. High frequency information in an input picture
is easily lost or distorted due to ltering and aliasing. Thus, a scanned analog screening
halftone or clustered-dot ordered dither halftone is more similar to a scanned template-dot
halftone than the halftones themselves.
20
Chapter 4
Review of Previous Work
In this chapter we describe previous research work. The work is classied into three
categories: 1) work whose results underly this research, 2) work that solves related problems,
and 3) work that addresses problems similar to the ones we are concerned with.
4.1 Research work underlying this research
Kermisch and Roetling [KR75] and Roetling [Roe77] derive and discuss the continuous
Fourier spectrum of a halftone picture. Such a spectrum is the sum of the spectrum of
the source picture and its higher order non-linear distortions. The nth order distortion is
centered at the nth harmonic of the halftone frequency and is scaled by a factor of
1
n
. They
consider the result to be applicable both to halftones produced by the analog screening
technique and to halftones printed via digital means, and in the latter case, ordered dither
halftones are assumed.
Allebach and Liu [AL77] give the discrete Fourier spectrum of a digital halftone im-
age that is produced by an ordered dither method. The result is similar to Kermisch and
Roetlings, just as most discrete Fourier analysis results are to their continuous space coun-
terparts.
Huang [Hua74] studies the Fourier spectrum of an ideally sampled halftone picture of
a single gray tone. He uses the spectrum to explain the formation of a moire pattern and
attributes the occurrence of a moire pattern to the low frequency components at aliasing
harmonics of the halftone frequency.
Steinbach and Wong [SW82] extend Huangs work by taking into account additional
21
CHAPTER 4. REVIEW OF PREVIOUS WORK
factors such as sampling aperture, unequal sampling rate in the two directions of sampling,
and non-perpendicular sampling lattice. They note that a moire pattern may also be at-
tributed to the beat of two high frequencies that are close to each other. They claim that
the moire fringe behavior remains the same for halftones with dierent content, and use this
claim to justify their use of a uniform tone image as the analysis subject. We will show in
Chapter 5 that their claim is correct in terms of the frequency of a moire pattern. However,
at dierent density levels of the picture content, a moire pattern may exhibit a half period
phase shift.
4.2 Research work on related problems
Wong [Won77] considers scanning a halftone picture into a gray scale image and then
converting the gray scale image to a two-level image by thresholding. He assumes that a
halftone picture is accurately scanned with a high resolution and explains why a simple
thresholding of the scanned image may cause moire in the binary output. He suggests an
error feedback method to compensate for the quantization error introduced in the thresh-
olding. He points out that the method is simply an application of the error diusion method
by Floyd and Steinberg [FS75] which is a digital halftoning method.
Allebach [All78] proposes a random nucleated h alftoning meth od to reduce the moire pat-
tern in a halftone picture whose graphical content includes periodic features. This method
randomizes the position of the dot in a halftone cell. The attempt is to reduce the periodic-
ity that the halftone dots may exhibit. In fact, the halftone dots in a halftone picture thus
obtained no longer form a lattice. One can expect that when such a picture is sampled,
the moire problem should not be as serious as with a clustered-dot halftone. However, this
method produces very grainy pictures.
Stoel [Sto80] and Stoel and Moreland [SM81] describe a pictorial reproduction system,
part of which re-halftones scanned halftone images. An input halftone is scanned at a xed
rate. They note that if the screen frequency of the input halftone is higher than one-eighth
22
CHAPTER 4. REVIEW OF PREVIOUS WORK
of the scanning rate, re-halftoning the scanned halftone by either xed level or adaptive
thresholding may result in an output image that has a moire artifact. Therefore, their
system uses autocorrelation to determine if the halftone frequency of a scanned halftone
image is higher than one-eighth of the scanning rate. If so, the system low-pass lters the
scanned image. The cut-o frequency of the lter is a xed value lower than one-eighth of
the scanning rate. Thus, the halftone structure in the scanned image is suppressed, and a
gray level image is obtained. The gray level image is then digitally halftoned. The authors
do not address directly the moire pattern problem that may arise from the scanning process.
Analoui and Allebach [AA92] consider reconstructing a gray scale image from a sam-
pled halftone image by rst estimating the halftone function that was used to generate the
input halftone picture. However, they assume that the sampled halftone image is binary
and that a period of the halftone screen covers an integer number of pixels. Thus, a sam-
pled halftone image under their assumption can only be obtained from a digital halftoning
algorithm. In other words, the sampled halftone image is really a digital halftone image.
In the literature, the reconstruction of a gray level image from a digital halftone image is
called inverse h alftoning. A digital halftone image can be considered to be ideally sam-
pled from a halftone picture with a sampling period that is an exact multiple of a halftone
cell. Consequently, the sampling rate is a multiple of the halftone frequency, and no moire
pattern occurs.
Miceli and Parker [MP92] propose three inverse halftoning algorithms that also require
a digital halftone image as input. They use the blue noise mask ([MP91]) type digital
halftone images in their experiments. They note that their algorithms do not work well on
clustered-dot halftone images.
4.3 Research work addressing similar problems
Shu, Springer, and Yeh [SSY89] categorize techniques for reducing moire patterns into
four categories: low-pass pre-ltering, post-scan processing, high frequency scanning, and
23
CHAPTER 4. REVIEW OF PREVIOUS WORK
direct manipulation of the factors that cause a moire. They suggest using a scanning rate
that is k +
1
2
times the halftone frequency, where k is an integer, to conne a moire to
high frequency (about one half the halftone frequency) so it is not very objectionable to
the human eye. They do not address the determination of the halftone frequency. They
also perform subjective visual experiments and claim that a moire pattern is most visible
in areas of a source halftone picture where the density is 60% to 80% (black). One practical
issue that may prevent the sampling rate selection approach from being useful is that a
scanner may not allow a user to select the true scanning rate at will, and the samples may
not be uniformly spaced.
Ohyama, Yamaguchi, Tsujiuchi, Honda, and Hiratsuka [OYT
+
86] approximate a
clustered-dot ordered dither halftone by a template-dot halftone (which they call pulse
width modulation or PWM) to devise a moire pattern suppression technique. They con-
sider, in the frequency domain, a simplied model of the Fourier spectrum for a template-dot
halftone and assume that a moire pattern is caused by the aliasing rst order harmonics of
the halftone frequency. They note that in the spectrum of a sampled halftone, the spectrum
of the source picture itself is independent of the sampling frequency, but that an aliasing
rst order harmonic of the halftone frequency from a rst order replica changes with the
sampling frequency. They then propose a dual sampling meth od that samples an input
halftone picture twice with two dierent sampling rates. The Fourier spectra of the two
sampled images are compared frequency by frequency, and for each frequency, the frequency
component with the smaller amplitude is retained. Since the frequency component of an
aliasing rst order harmonic of the halftone frequency appears as a spike and occurs at
dierent positions in the two spectra, the selection process eectively retains the correct
frequency components. They note the dierence between a template-dot halftone and a
clustered-dot ordered dither halftone but believe that the two types of halftones can be
treated the same way.
This method has the advantage of being logically simple. It also has two drawbacks.
First, since the comparison process can only be performed up to half the lower sampling
24
CHAPTER 4. REVIEW OF PREVIOUS WORK
rate, the high frequency image information carried by the image sampled with the higher
rate is wasted. Second, a practical and critical issue with this method is how to make two
scans of th e same scene at two resolutions. If the two scanned images contain dierent
graphical content, the combined spectrum would yield an incorrect image because of the
mismatch of phase information from the two spectra. A consequence is that the method
can only be used to prevent a moire pattern when a halftone is scanned because it is not
likely that an image archive would carry two images sampled at dierent rates from the
same scene with exactly the same boundary. In a later paper [YOH
+
89], the authors do
indicate that the method is meant to be a prevention method and suggest using two sensor
arrays of dierent resolutions on a scanner. However, as demonstrated in Chapter 9, moire
prevention can be achieved more eectively and more eciently.
In addition, the method assumes that the input picture is uniformly sampled. The
method may not work on two images scanned from exactly the same scene at two resolutions
using a commercial scanner, because the two images may be subsampled non-uniformly at
dierent average rates from the same optoelectronically scanned image. An aliasing halftone
frequency harmonic may appear at the same frequency in the spectra of both images. In
Chapter 6 we discuss non-uniform sampling in detail, and in Chapter 8 we compare the
dual sampling method with our methods.
Krause [Kra92] gives some empirical, ad h oc methods for reducing moire patterns that
manually manipulate a scanned image. These methods operate on the pixel values that are
scanned in; they require patience and experience and are more art than science.
The HP ScanJet IIcx Scanner Users Guide [Hew93] also suggests some empirical meth-
ods such as changing the scanning rate and comparing the results or placing a transparent
plastic sheet between the scanner and the source halftone picture (low-pass pre-ltering).
Low-pass pre-ltering the input halftone is meant to reduce the signicance of the halftone
structure, but the amount of ltering is dicult to control manually.
25
Chapter 5
Characterization of Moire in Sampled Halftones
5.1 Introduction
In this chapter we develop the theory for characterizing moire patterns in sampled
halftones. We show how the frequency, initial phase, and geometric shape of a moire
pattern are dependent upon the halftone frequency, the sampling frequency, and the oset
between the halftone lattice and the sampling lattice.
We can analyze a moire pattern either in the spatial (image) domain or in the frequency
(Fourier) domain. It appears to be easier to study the dierent components of a moire in
the frequency domain. The Fourier spectrum of a sampled halftone provides much insight
into what frequencies contribute to a moire and to what extent they contribute to it. On
the other hand, it is also obvious that only a spatial domain synthesis of the contributing
frequencies can ultimately demonstrate the appearance of a moire and its visual eect. In
this chapter we take this hybrid approach.
5.2 Notation and denitions
In this section we summarize the mathematical notation and some other relevant de-
nitions used in this dissertation.
A point in a 2D plane is almost always written as a vector. Vectors are denoted by bold
face letters. In particular, x = (x, y ) is a variable point in image space (also referred to as
the spatial domain). A function in image space is denoted by one or more lower case letters.
For example, a real-valued function, f(x), called a picture function [Ros69], represents a
picture. In this dissertation, all picture functions represent image density and have the
26
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
range [0,1]. If f(x
0
) = 0, the picture has maximum brightness at x
0
. A picture function
representing a halftone picture has binary values 0 and 1. In this chapter, we deal entirely
with picture functions in continuous space. A function f(x) is even if f(x) = f(x), that
is, f(x) is symmetric with respect to the origin. The Fourier transform of an image space
function is denoted by the same letter(s) denoting the function but in upper case. For
example F(u) is the Fourier transform of f(x), where u = (u, v) is a variable vector in the
frequency space (also referred to as the frequency domain), and a point in the frequency
space is also almost always represented as a vector, such as u above. When we plot F(u),
we always plot only the amplitude [F(u)[. In the literature, [F(u)[ is called the periodogram
of f(x). The value F(u
0
) is the (frequency) component of f(x) at frequency u
0
. [F(u
0
)[
2
is the energy of f(x) at u
0
. (x) is the impulse or Dirac function. The sum of an array
of innitely many uniformly spaced impulse functions is called an impulse train. The sinc
function is dened as sinc(x) =
sin x
x
. In an equation, is the dot-product operator,
is the cross-product operator, is the convolution operator, and the multiplication
operator between two scalars is usually implied. The complex imaginary quantity

1 is
represented by j. Occasionally j is also used as a summation subscript. The meaning of
j should be clear from the context. A general sinusoid is normally expressed as a cosine
function because the latter is even and may lead to simpler mathematical expressions.
5.3 Interpretation of the Fourier spectrum of a picture
The continuous Fourier transform F(u) of a real picture f(x) is h ermitian. That is,
F(u
0
) is the complex conjugate of F(u
0
) for any point u
0
. Loosely speaking, the pair of
frequency components F(u
0
) and F(u
0
) correspond to a sinusoid with frequency u
0
in the
picture, and the amplitude of the sinusoid is proportional to the magnitude of F(u
0
). f(x)
is the integral sum of sinusoids whose frequencies and amplitudes are specied by F(u).
If F(u) has a component (+i)(uu
0
), it must also have a component (i)(u+
u
0
). The pair of impulses correspond to a sinusoid cos(2u
0
x +
0
) in f(x) where
27
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
= 2
_

2
+
2
and
0
= arctan

. The sinusoid is a wave traveling in the u


0
direction
with frequency [u
0
[ (or period
1
|u
0
|
) and initial phase
0
.
Many textbooks, such as Bracewell [Bra86], have general discussions of Fourier trans-
forms.
5.4 The Fourier spectrum of an ideally sampled uniform halftone
A sampled halftone of a single gray level, called a sampled uniform h alftone, exhibits the
most pronounced moire. Such a sampled halftone is also relatively easy to analyze. In this
section, we discuss the Fourier spectrum of a uniform halftone and the Fourier spectrum of
a sampled uniform halftone.
5.4.1 The Fourier spectrum of a uniform halftone
A uniform halftone consists of a lattice of dots of the same shape and size. It can be
expressed as the convolution of a 2D impulse train that reects the periodicity of the lattice
and a function that denes one halftone cell. The convolution replicates the halftone cell
at each impulse. Formally, let h(x) be a uniform halftone. Then
h(x) =
_
_

m,n=
(x ma nb)
_
_
c(x) =

m,n=
c(x ma nb),
where the vectors a and b dene the shape of a halftone cell as well as the shape of the
halftone lattice. The origin of the x coordinate system is at the center of a halftone cell.
The function c(x) denes a halftone cell and satises
c(x) =
_
1, if x in the dot area;
0, otherwise.
We assume c(x) is even. Figures 5.1 (a) and (b) illustrate, respectively, the layout of h(x)
and c(x) in the image space. In the gures dark spots represent the halftone dots, and the
dot shape is drawn as circular. h(x) and c(x) have value 1 in dark area(s) and have value
0 elsewhere. The parallelograms show the halftone cells. In the literature, the halftone
28
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
y
a
b
(b) (a)
x
y
x
Figure 5.1: Halftone representation of a uniform gray tone
lattice is usually assumed to be square, which is indeed usually the case in reality. A square
halftone lattice corresponds to a b and [a[ = [b[. The halftone lattice is often oriented
at 45 degrees, corresponding to a and b pointing 45 degrees to the lower right and upper
right, respectively.
The Fourier transform H(u) of h(x) is the product of two Fourier transforms: that of
the impulse train that denes the halftone lattice, which is another impulse train, and that
of one halftone cell. We have
H(u) =
_
_

m,n=
(u mp nq)
_
_
C(u), (5.1)
where p =
b (a b)
[a b[
2
and q =
a (b a)
[a b[
2
dene the impulse train in the frequency do-
main that corresponds to the impulse train in the spatial domain. C(u) is the Fourier
transform of c(x) and has the general property of decreasing in magnitude as [u[ increases.
Thus, C(u) attenuates the magnitude of impulses that are away from the origin. C(u) is
real because c(x) is even.
p and q are the halftone frequency vectors. They represent both the halftone lattice
density and the lattice shape and orientation. (The lattice density should not be confused
with the density of a picture.) For brevity, we call p and q the halftone frequencies, even
though [p[ and [q[ are more commonly called the halftone frequency in the literature. We
use the plural term h alftone frequencies to refer to p and q when we want to emphasize
29
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
p
b
a
p
u
v
q
q
Figure 5.2: The layout of the Fourier spectrum of a uniform halftone
that they have dierent vector values. We use the singular term h alftone frequency to refer
collectively to both p and q or both [p[ and [q[ when we only want to mean the density of
the halftone lattice, especially when [p[ = [q[.
Figure 5.2 depicts the layout of H(u) as well as the relationship between the directions of
a,b and the directions of p and q. Note that p is perpendicular to b, and q is perpendicular
to a. In the gure, each spot represents an impulse that is a frequency component of
h(x). The position of the spot represents the frequency of the impulse, and the size of
the spot indicates the magnitude of the impulse. Such an impulse corresponds to a term
(u mp nq)C(u) in Eq. 5.1 and is located at a frequency that is a harmonic of the
halftone frequencies p and q. The impulse is referenced by its harmonic order (m, n) of
the halftone frequencies and is also loosely said to be the ith order component, where
i = [m[ +[n[. The strongest impulse is always the zeroth harmonic component because the
envelope C(u) achieves maximum magnitude at u = 0. In other words, the zero frequency
component is always the strongest.
In this dissertation we mostly use a layout plot as in Figure 5.2 to illustrate the Fourier
spectrum of a picture. Ulichney ([Uli87]) calls this type of plot the exposure plot. Oc-
30
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Circular dot
u
v
C(u,v)
u
Square dot
u
v
C(u,v)
u
Figure 5.3: A circular dot, a square dot, and their Fourier spectra
casionally we also show a 3D plot of the magnitude of a Fourier spectrum to provide more
perspective. Of the two types of plot, the layout plot provides more precise (albeit more
abstract) information of the spectrum.
The function C(u) is determined by the shape of a halftone dot (Allebach and Liu
[AL77] call the shape the dot prole). For a circular shaped dot, C(u) is of the form
C(u) =
1
J
1
(
1

u
2
+v
2
)

u
2
+v
2
,
where J
1
(u) is the rst order Bessel function and
1
and
1
are constants dependent upon
the size of the dot. For a square dot, C(u) is a two-dimensional sinc function of the form
C(u) =
2
sin
2
u

2
u
sin
2
v

2
v
where, again,
2
and
2
are constants dependent upon the size of the dot. For a diamond-
shaped dot, C(u) is the same as for the square dot, except for a rotation by 45 degrees in the
frequency plane [KR75]. Figure 5.3 shows a circular dot in a square halftone cell, a square
dot in a square cell, and C(u) for the dot shapes. Figure 5.4 shows two diamond-shaped
dots of dierent sizes. For a convex, arbitrarily shaped clustered dot, C(u) is a surface that
31
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.4: Two diamond-shaped dots
achieves maximum at (0, 0) and decreases in magnitude with a wavy skirt as frequency
[u[ increases. C(u) may be negative for some values of u.
The Fourier transform of a uniform halftone is the sum of a set of impulses with dierent
magnitude. The impulses are located at harmonic frequencies of the halftone frequencies.
Each impulse also has a complex conjugate in the set. The pair of impulses correspond to
a sinusoid in the the spatial space. Therefore, a uniform halftone can also be represented
by a Fourier series.
5.4.2 The Fourier spectrum of an ideally sampled uniform halftone
Now we sample h(x) with a lattice dened by vectors d and e. That is, we take samples
at interval [d[ in the d direction and at interval [e[ in the e direction. We assume that a
sampling point coincides with the origin of our x coordinate system. The resultant sampled
halftone h
s
(x) can be expressed as
h
s
(x) = h(x)
_
_

k,l=
(x md ne)
_
_
. (5.2)
The Fourier transform H
s
(u) of h
s
(x) is
H
s
(u) = H(u)
_
_

k,l=
(u ks lt)
_
_
=

k,l=
H(u ks lt)
=

k,l=
_
_

m,n=
(u ks lt mp nq)
_
_
C(u ks lt), (5.3)
where s =
e (d e)
[d e[
2
and t =
d (e d)
[d e[
2
represent the sampling frequencies and their
directions. The Fourier transform of the sampled halftone is that of the original halftone
32
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
v
t
s u
Figure 5.5: The Nyquist band for sampling frequencies s and t
replicated in the s direction with spacing [s[ and then replicated in the t direction with
spacing [t[. Each copy is called a replica, and a replica is referenced by its sampling fre-
quency harmonic order (k, l). The (k, l)th replica is also referred to as an ith order replica,
with i = [k[ +[l[. The (m, n)th halftone frequency harmonic of the (k, l)th replica is referred
to as the (k, l, m, n)th halftone frequency harmonic, and the component at the (k, l, m, n)th
halftone frequency harmonic is referred to as the (k, l, m, n)th component. The (k, l, m, n)th
component is an alias of the (0, 0, m, n)th component, and is also called an aliased compo-
nent.
From sampling theory, we know that when a continuous signal is sampled, the sampled
signal can retain only those components of the continuous signal whose frequencies are
lower than half the sampling frequency. The half sampling frequency spectral band is called
the Nyquist band. The dotted parallelogram in Figure 5.5 shows the Nyquist band of the
sampling frequencies s and t. An aliased component that appears in the Nyquist band is
called an aliasing component. The discrete Fourier transform (DFT) of a sampled signal
also represents only frequencies in the Nyquist band. We use the term DFT spectral band
to refer to the Nyquist band when discussing the DFT of a sampled halftone.
Figure 5.6(a) illustrates the layout of the spectrum H
s
(u) of a sampled uniform halftone.
The original halftone is the one shown in Figure 5.1, with spectrum shown in Figure 5.2.
33
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Also shown in Figure 5.6(a) are the sampling frequencies s and t, indicating how the spec-
trum of the original halftone is replicated. To show the distribution pattern of frequency
components, more high order halftone frequency harmonics than in Figure 5.2 are drawn
so that the halftone frequency harmonics from dierent replicas can be explicitly shown
to overlap. Since each replica is enveloped by a decreasing surface, replicas other than the
nine nearest to the origin do not contribute signicant frequency components to the Nyquist
band. Thus, only the nine replicas nearest to the origin are shown, and they are the replicas
that need to be considered in moire analysis.
Due to replica overlapping, frequency components from dierent replicas may come
close to one another. Note that a frequency component here is an impulse, and together
with its complex conjugate, corresponds to a sinusoid in the spatial domain. Summing two
sinusoids with similar frequencies produces a low frequency beat pattern that resembles
neither sinusoid. The sum of multiple sinusoids of similar frequencies may also exhibit a
low frequency though more complex pattern. In Section 5.7 we discuss the surface formed
by the sum of multiple sinusoids in detail.
The frequency component distribution in Figure 5.6(a) is very regular, because the
component distribution pattern in one replica is regular, and the locations of the replicas
are also regular. We can partition the frequency plane with parallelograms formed using
the halftone frequencies p and q, as shown by the dashed parallelograms and the solid
parallelogram in Figure 5.6(a). Each parallelogram is centered at a halftone frequency
harmonic of the (0, 0)th replica, and there is a cluster of frequency components in the
parallelogram. In addition, let be the frequency of a component in the cluster at the
zero frequency, and let (m, n, l, k) be the order of the component. That is, = ks +
lt + mp + nq, and the frequency dierence between the (m, n, l, k)th component and the
(0, 0, 0, 0)th component is . Then for any component (k

, l

, m

, n

), there is a component
(k

+k, l

+l, m

+m, n

+n) such that the frequency dierence between the two components
is .
Figure 5.6(b) shows a close-up of the cluster at zero frequency. Let
1
and
2
be the
34
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
1

(b)
p
s
t
q
(a)
Figure 5.6: The layout of the Fourier spectrum of a sampled uniform halftone
35
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
frequency vectors as shown in the gure. If r is the frequency of a component in a cluster,
then any component in the cluster has a frequency of the form r + i
1
+ j
2
for integer
values of i and j. In other words, all clusters have the same grid structure. Each of these
clusters produces the same beat frequency.
Now we examine the frequency component distribution within the solid parallelogram in
Figure 5.6. We rst notice that each replica contributes one component to the parallelogram.
We then see that the four components contributed by the four rst order replicas form a
small parallelogram, as shown by the dotted parallelogram in Figure 5.6. So do the other
four components which are contributed by the four second order replicas (although that
parallelogramis not shown because the gure is too crowded). The two small parallelograms
and the solid parallelogramare concentric. When both the halftone lattice and the sampling
lattice are square, each of the three parallelograms becomes a square.
As the last part of this subsection, we introduce an oset between the halftone lattice
and the sampling lattice. We still let a sampling point coincide with the origin of the x
coordinate system but let the uniform halftone h(x) translate by x. The sampled halftone
and its Fourier transform then are, respectively,
h
s
(x) = h(x x)
_
_

k,l=
(x md ne)
_
_
, (5.4)
and
H
s
(u) =
_
H(u)e
j2xu
_

_
_

k,l=
(u ks lt)
_
_
=

k,l=
e
j2x(ukslt)
H(u ks lt)
=

k,l=
_
_

m,n=
e
j2x(ukslt)
(u ks lt mp nq)
_
_
C(u ks lt).
(5.5)
The shift of h(x) in the image space results in a linear phase shift to every frequency com-
ponent in H
s
(u). The phase shift does not aect the distribution of frequency components
36
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
shown in Figure 5.6 because it is a multiplicative factor of unit magnitude.
For the (l, k, m, n)th frequency component, the phase shift factor is e
j2x(mp+nq)
,
which is independent of the sampling frequency harmonic order. Note that, depending
on the size of the halftone dots, the function C(u ks lt) may be either positive or
negative. The initial phase of the frequency component is thus either 2x (mp + nq)
or 2x (mp + nq) + . The initial phase is determined by 1) the halftone lattice, the
sampling lattice, and their relative position and 2) the gray level that the halftone represents.
In Section 5.8 we will show that the change in polarity of C(u ks lt) due to change in
gray level of the image content may cause the moire to exhibit a phase reversal.
Let (l, k, m, n) again be the order of a component in the cluster at zero frequency, and let
be the phase shift to the component, = 2x (mp+nq). The (l, k, m, n)th component
then has a phase dierence of or + to the (0, 0, 0, 0)th component, because the latter is
always real and nonnegative. The phase shift dierence between the (k

, l

, m

, n

)th and the


(k

+k, l

+l, m

+m, n

+n)th components is 2x [(m

+m, n

+n)(m

, n

)] (p, q) = .
Thus, the (k

, l

, m

, n

)th and the (k

+k, l

+l, m

+m, n

+n)th components have a phase


dierence of either or + .
As a summary of this subsection, we have the following theorem:
Theorem 5.4.1 Th e Fourier transform of an ideally sampled uniform h alftone consists of
impulses of dierent magnitudes. Th ese impulses may form clusters in th e frequency plane.
All clusters h ave th e same frequency distribution grid pattern. In addition, if a pair of
impulses h as a frequency dierence th at is equal to th at of anoth er pair, th e two pairs of
impulses also h ave th e same ph ase dierence modulo .
5.5 The Fourier spectrum of a practically sampled uniform halftone
In this section, we add three practical considerations to the Fourier spectrum of a sam-
pled uniform halftone. These include low-pass ltering prior to sampling, typical geometry
of the halftone screen, and typical geometry of the sampling lattice.
37
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
A practical sampler is usually preceded by a low-pass lter. For a scanner, one low-pass
lter is its optical system. In such cases a low-pass ltered (blurred) version of the input
picture is sampled. A practical printing process also performs low-pass ltering because
it does not produce halftone dots with ideally sharp boundaries. Since low-pass ltering
is cumulative, for brevity, we do not distinguish low-pass ltering by a printer from that
by a sampling device. A scanner may subsample the optoelectronically sampled image to
produce an output image of a lower resolution. The subsampling may also be preceded by
digital low-pass ltering.
Assuming that a low-pass lter precedes the ideal sampler, Eqs. 5.4 and 5.5 become,
respectively,
h
s
(x) = [h(x x) lp(x)]
_
_

k,l=
(x md ne)
_
_
, (5.6)
and
H
s
(u) =
_
H(u)e
j2xu
LP(u)
_

_
_

k,l=
(u ks lt)
_
_
=

k,l=
e
j2x(ukslt)
H(u ks lt)LP(u ks lt)
=

k,l=
_
_

m,n=
e
j2x(ukslt)
(u ks lt mp nq)
_
_
C(u ks lt)LP(u ks lt), (5.7)
where lp(x) is the low-pass lter function and LP(u) its Fourier transform. We assume
that lp(x) is even so that LP(u) is real and does not aect the phases of any frequency
components. Through the convolution, lp(x) performs a local weighted averaging of h(x).
Correspondingly, the (k, l)th replica in H
s
(u) is now subject to an additional attenuation
by LP(u ks lt). LP(u) usually is a Gaussian-like, bell-shaped surface. Theoretically,
LP(u) may be negative, thus aecting the phase of a frequency component.
In reality, a sampled halftone is almost always displayed on a screen in which pixels
are equally spaced horizontally and vertically. We can then regard the sampling lattice as
38
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
s
t
q
p
Figure 5.7: The layout of the Fourier spectrum of a typical, practically sampled uniform
halftone
being square and attribute any possible non-orthogonality or unequal spacing to the picture
being sampled. For example, in case of unequal sampling frequencies in the two directions
of sampling, we still regard the sampling frequencies as being equal but consider an input
picture to be stretched. So, from now on, unless otherwise stated, d and e are assumed
to be perpendicular, of equal length, and horizontally/vertically oriented. Correspondingly,
s and t are perpendicular, have equal length, and are horizontally/vertically oriented.
As mentioned earlier, most halftones use a screen that is a 45 degree square lattice. We
now also assume [a[ = [b[ and a b. Thereby, [p[ = [q[ and p q.
Figure 5.7 shows the layout of the spectrum of a typical sampled uniform halftone. The
gure shows only the spectral band in which frequencies are lower than the sampling fre-
quency. Here the halftone lattice is oriented at about 48 degrees from horizontal, indicating
a slight rotation of the halftone under the sampling lattice. The sampling frequency [s[ is
about 1.63 times the halftone frequency [p[, and the halftone frequency vectors p and q fall
39
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
in the Nyquist band of the sampled halftone which is shown by the dotted square in the
gure. Halftone frequency harmonics of orders higher than 3 are assumed to be insignicant
because of low-pass ltering; they are not drawn in the gure. The dashed squares show
the same partition of the frequency plane as in Figure 5.6.
The frequency components in Figure 5.7 have a clear structure: clusters of nine close
components in each of the partitioning squares. There are three types of clusters. The
rst type consists of a single cluster; the cluster is centered at the origin and is formed by
one zeroth order and eight second order halftone frequency harmonics. The second type
consists of four clusters; each cluster is centered at a rst order halftone frequency harmonic
of the zeroth replica and is formed by four rst order and ve third order halftone frequency
harmonics. The third type of clusters are those near the centers of the non-zeroth order
replicas.
By Theorem 5.4.1, for any two pairs of components such that one pair has the same
frequency dierence (in the vector sense) as the other pair, the phase dierence between the
rst pair is equal to that between the second pair modulo . It is also easy to verify that
if the low-pass lter has a linear phase shift, that is, the complex phase of LP(u) is linear
in u, the frequency components in H
s
(u) still have the same phase dierence property. In
Section 5.7 we show that each of the rst two types of clusters produces a dierent kind of
periodic pattern in the spatial domain, but the two kinds of patterns have the same beat
frequency. The third type of clusters have frequencies near the sampling frequency. They
have very little inuence on the sampled halftone when the latter is displayed, because they
are heavily low-pass ltered by the nite pixel (aperture) size of the display.
From the sampling theory point of view, the eight second order halftone frequency har-
monics in the rst type cluster are aliasing frequencies. However, the four rst order halftone
frequency harmonics in each of the second type clusters are not aliasing frequencies. The
other ve third order halftone frequency harmonics in a second type cluster may be aliasing
frequencies, but the corresponding components are much weaker than the components at
the rst order harmonics. The third type of clusters consist only of non-aliasing frequencies.
40
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
When the halftone is rotated under the sampling lattice, each replica will also rotate
accordingly. Then the frequency components from dierent replicas will recombine to form
new clusters.
When a moire pattern has aliasing components, we call it an aliasing moire. When
a moire consists only of non-aliasing components, we call it a non-aliasing moire. Most
often, a moire pattern is an aliasing moire. A non-aliasing moire occurs when adjacent
replicas do not overlap but they have strong components near the border of the Nyquist
band to beat with one another. In Figure 5.7, if components at both the second and third
order halftone frequency harmonics are insignicant compared to the components at the
rst order harmonics, the resultant moire can be considered to be a non-aliasing moire.
Another example of a non-aliasing moire is in Chapter 9. Usually, part of an aliasing moire
is a non-aliasing moire.
5.6 The Fourier spectrum of a practically sampled general halftone
Kermisch and Roetling [KR75] derive, in the continuous space, the Fourier spectrum of
a general halftone picture. Such a spectrum is the sum of the spectrum of the graphical
content and the nonlinear distortions of the latter spectrum. The distorted spectra are
centered at harmonics of the halftone frequency and are subject to a decreasing envelope
similar to C(u) in the preceding two sections.
Let 0 f(x) 1 be a source picture, and let h(x) be its halftone with a square screen
lattice. Assume that the origin of the x coordinate system is at the center of a halftone
cell. Then ([KR75], adapted to our notation)
h(x) =

m,n=
c[x ma nb; f(x)],
where c[x; f(x)] maps the content of f(x) in the halftone cell area [x[ < [a[/2 into a dot in
the cell:
c[x; f(x)] =
_
1, if [x[ < [a[/2 and x in the dot area;
0, otherwise.
41
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
c[x; f(x)] denes the shape, size, and position of the halftone dot.
The Fourier transform H(u) of h(x) is ([KR75], adapted to our notation and coordinate
system)
H(u) = [p[
2

m,n=
_

=
C[mp + nq; f()]e
j2(umpnq)
d, (5.8)
where C[mp + nq; f()] is the Fourier transform C[u; f()] of c[x; f()] evaluated at u =
mp+nq, with some input picture value f(). The transform C[u; f()] is calculated taking
f() as a constant. In other words, letting f
0
= f(), C[u; f
0
] is the Fourier transform of
c(x; f
0
) which denes a dot for a uniform tone of gray level f
0
. c(x; f
0
) is equivalent to c(x)
in Section 5.4 except that the gray level f
0
is not explicitly expressed in c(x).
The (m, n)th term in the summation in Eq. 5.8 is the Fourier transform of C[mp +
nq; f()] translated to center at u = mp + nq. Dene
H
m,n
(u) =
_

=
C[mp +nq; f()]e
j2u
d. (5.9)
Then Eq. 5.8 becomes
H(u) = [p[
2

m,n=
H
m,n
(u mp nq).
Note that H
0,0
(u) is the transform of f(x).
In Kermisch and Roetling [KR75], Eq. (5.8) is derived for square screens. It can be shown
that the equation also applies to non-square and non-orthogonal screens. The equation also
applies to dispersed-dot ordered dither halftone which is outside the scope of this research.
An analysis of Eq. 5.9 reveals some of its characteristics. First, it is easy to see that
H
m,n
(0) is real. Second, H
0,0
(u) achieves maximum magnitude at u = 0 because f(x) is
real and non-negative. Third, the magnitude of H
m,n
(x) becomes smaller as m or n assumes
larger value, because C[mp + nq; f
0
] becomes smaller in magnitude (for any value of f
0
).
In addition, we have the following theorem regarding the [m[ +[n[ = 1 cases of Eq. 5.9.
Theorem 5.6.1 If c(x; f
0
) denes a clustered dot th at is symmetric with respect to th e
center of its parallelogram h alftone cell for any value of f
0
, (th at is, c(x; f
0
) is even),
42
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
H
1,0
(u) in Eq. 5.9 ach ieves maximum magnitude at u = 0, and H
1,0
(0) is real and positive.
Th e oth er th ree rst order terms in Eq. 5.9 h ave th e same property.
To prove the H
10
(0) case in Theorem 5.6.1, note that the function C[mp + nq; f()] in
Eq. 5.9 is now
C[p; f()] =
_

x=
c[x; f()]e
i2px
d x,
which can be shown to be real and non-negative for any . Theorem 5.6.1 states that, under
fairly general conditions, [H(u)[ has peaks at the four rst order harmonics of the halftone
frequency.
It can be veried that C[u; f
0
], as a function of f
0
, varies very slowly with f
0
. Therefore
the magnitude of H
m,n
(u) in Eq. 5.9 is likely to achieve maximum value near u = 0
and decreases quickly with increasing [u[. Furthermore, small changes in f() do not
signicantly aect the magnitude. In other words, the value of [H
m,n
(u)[ near u = 0 is
largely determined by the larger features of the source picture. In Chapter 8 we make use
of this property of the spectrum of a halftone to devise a moire reduction algorithm.
We may now visualize the magnitude of the Fourier spectrum H(u) of a general halftone
as an array of sharp hills located at the harmonics of the halftone frequency, similar to
the spectrum of a uniform halftone which is an array of impulses at the harmonics of the
halftone frequency. However, as shown in Section 5.8, a hill at a higher than 1 order
harmonic of the halftone frequency may have its tip missing. In other words, [H(u)[ may
not have a maximum value at a higher order harmonic of the halftone frequency.
Figure 5.8 shows a test halftone picture House that is used in this dissertation. The
picture is produced using a digital halftoning algorithm with an 85 dpi, 45 degree square
screen. We scan a portion of the test halftone picture at 400 dpi using an HP ScanJet
IIc scanner and obtain a scanned halftone image of 1024 1024 pixels. Figure 5.9 shows
a 512 512 pixel portion of the scanned halftone image. Figure 5.10(a) shows in 3D the
magnitude of the Fourier spectrum of the 1024 1024 pixel scanned image. The scanning
is preceded with sucient low-pass ltering to prevent aliasing. The Fourier spectrum of
43
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.8: House, a test halftone picture
the original halftone should be similar except that the higher order hills are taller. To
speed up the reproduction of the graphics and save storage, the spectrum is smoothed
before being printed. The hills in the actual spectrum are sharper. The angle between
the halftone lattice and the sampling lattice is nearly 45 degrees. It can be seen that the
halftone structure is more pronounced in one direction than in the other.
Figure 5.10(b) shows the distribution of frequency components in Figure 5.10(a) that
are stronger than a certain threshold. This spectrum looks similar to the Fourier spec-
trum layout plot in the previous two sections because here a strong frequency component is
surrounded by frequency components that are also relatively strong, which can be seen in
Figure 5.10(a). Three factors contribute to this coincidence: 1) the nature of the halftone,
2) the leakage phenomenon of the spectrum of the sampled image , and 3) the smoothing
performed on the spectrum. Leakage ([Bel84, ch. 2], [Bri74, ch. 6 and 9], and Section 7.2)
occurs when a frequency component in an input signal is not a multiple of the sampling
44
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.9: A 512512 pixel portion of an image scanned from the House halftone picture
at 400 dpi
45
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
u
v
|H(u,v)|
u
(a)
(b)
Figure 5.10: The Fourier spectrum of the scanned halftone House
46
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
frequency, in which the energy of that component is distributed or leaked to neighboring fre-
quencies that are multiples of the sampling frequency in the sampled signal. The coincidence
helps identify the positions of strong frequencies and their strengths in Figure 5.10(b).
The spectrum of a sampled general halftone is that of the halftone replicated at har-
monics of the sampling frequency. Assuming a low-pass lter preceding the sampler and a
shift of the halftone under the sampling lattice, we have
H
s
(u) = [H(u)e
j2xu
LP(u)]
_
_

k,l=
(u ks lt)
_
_
=

k,l=
e
j2x(ukslt)
H(u ks lt)LP(u ks lt)
= [p[
2

k,l=
_
_

m,n=
e
j2x(ukslt)
H
m,n
(u ks lt mp nq)
_
_
LP(u ks lt). (5.10)
H
s
(u) can be visualized as the spectrum of a sampled uniform halftone but with impulses
replaced by hills. Thus, the analysis of the distribution of frequency components for
a sampled uniform halftone applies to a sampled general halftone. Again, for a scanned
halftone only components near the several low order harmonics of the halftone frequency
may survive the low-pass ltering of the scanner. Only these components need to be con-
sidered to analyze a moire pattern.
The initial phases of the frequency components near a higher than 1 order halftone
frequency harmonic or an alias of the harmonics are more complicated than the case of a
sampled uniform halftone. In Section 5.8 we discuss the phase issue.
Figure 5.11(a) shows the above 400 dpi scanned test halftone resampled at 133
1
3
dpi
with no additional low-pass ltering. Figure 5.11(b) shows the thresholded spectrum of
the resampled image. Also marked in the Figure 5.11(b) are the orders of several halftone
frequency harmonics or their aliases. Figure 5.11(b) shows only the Nyquist band of the
resampling rate, and the halftone frequencies fall in the band. Experiments show that the
visible moire is caused by two types of frequency components. The rst type are those
47
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
(a)
(-1,0,2,-1)
(1,1,-1,-1) (0,1,-1,1)
(0,1,-2,1)
(0,0,0,1)
(b)
Figure 5.11: Resampled scanned halftone House and its Fourier spectrum
48
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
at the aliasing halftone frequency harmonics near the zero frequency (six are signicant;
two are marked). The second type are those at the four rst order halftone harmonics and
their counterparts from the neighboring replicas which are not shown in the gure. These
components belong to the rst two types of clusters discussed at the end of Section 5.5.
Figure 5.7 indicates the positions of the neighboring replicas.
The resampled image in Figure 5.11(a) itself is halftoned when reproduced on paper.
The halftone screen used in the reproduction is carefully chosen so that the moire pattern
resulting from the interference between this screen and the halftone screen in the resampled
image is minimized. The same care is taken to reproduce all the other sampled halftone
images in this dissertation.
As can be seen from Figures 5.10, the hills at harmonics of the halftone frequency are
very sharp, so they can be modeled or approximated by impulses. In a sampled general
halftone the beat of these components with their aliased counterparts cause the visible
moire. It is also exactly these frequency components about which we have some knowledge;
their frequencies are determined by the geometry of the halftone lattice and that of the
sampling lattice. However, the strength of these components, or the height of the hills,
depends upon the image content as shown in Eq. (5.8), as well as upon the amount of
low-pass ltering the halftone picture receives before being sampled. The strength of these
frequency components is therefore dicult to compute in practice. Computation of the
strengths of these components may be considered in future research.
5.7 The geometric shape of a moire pattern
In Sections 5.45.6 we have explained in the frequency domain that the spectrum of a
sampled halftone image may have strong frequency components that are close to one another
and form clusters that have a regular pattern. In this section we study in the spatial
domain the consequence of having these clusters of frequency components. Specically,
since each strong frequency component and its complex conjugate correspond to a sinusoid
49
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.12: The beat pattern of two sinusoids
in the spatial domain, we study the behavior of a sum of sinusoids to gain insight into the
appearance of a moire pattern.
5.7.1 Sum of sinusoids
In this section we derive the equations for the sum of sinusoids. These equations facilitate
the analysis of the interference between sinusoids. We begin with the sum of two sinusoids
in one variable of the form a cos(
1
x +
1
) + b cos(
2
x +
2
). Next we extend the result
to the sum of an arbitrary number of sinusoids in one variable. Finally we extend the 1D
result to 2D.
When two sinusoidal waves, whose frequency dierence is much smaller than their re-
spective frequencies, are added, the result is a wave that oscillates at a high and periodically
changing frequency. The wave is also modulated by a periodically oscillating envelope at a
much lower frequency. Figure 5.12 shows such a sum of two sinusoids. If the sum represents
the density prole in the horizontal direction of an image that is vertically uniform, then
the image is a bar pattern whose contrast varies at a frequency that is much lower than the
the frequency of the bar pattern itself. Figure 5.13 shows such a bar pattern. Because the
human eye is more sensitive to bright areas, and because the bright bars in high contrast
areas are brighter than the bright bars in low contrast areas, high contrast areas of the bar
pattern appear to be brighter than low contrast areas even though the average brightness in
both areas is the same. Related to this phenomenon of the human eye, Adar Pelah [Pel94]
50
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.13: The bar pattern formed by the sum of two sinusoids
shows how to produce a (constant contrast) bar pattern that has white bars and dark bars
of perceptually equal width by printing the bright bars narrower than the dark bars.
A periodic pattern is a contrast pattern if it is a high frequency oscillation whose ampli-
tude is modulated by a lower frequency. A periodic pattern is a density pattern if it is an
oscillation with a constant amplitude.
To derive the equation that shows the beat between two sinusoids, we consider each
sinusoid to be an oscillation. We represent an oscillation using a complex function of the
form A(x)e
j(x)
([Cuc52]). The complex function corresponds to a vector rotating in the
complex plane. The real part Re[A(x)e
j(x)
] = A(x) cos (x) represents the instantaneous
position of the oscillation, the magnitude A(x) represents the instantaneous amplitude of
the oscillation, and the argument angle (x) represents the instantaneous phase of the
oscillation.
In such a scheme a cos(
1
x +
1
) is represented as v
a
(x) = ae
j(
1
x+
1
)
and b cos(
2
x +

2
) as v
b
(x) = be
j(
2
x+
2
)
. The sum of the two sinusoids is then represented as v(x) =
A(x)e
j(x)
= v
a
(x) + v
b
(x), as is shown in Figure 5.14. A(x) can be calculated as
51
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
+
v (x)
1
v(x)
x
2
x
a
1
2
(x)
b
v (x)
+

Figure 5.14: Vector sum of amplitudes of sinusoids


A(x) =
_
(a cos(
1
x +
1
) + b cos(
2
x +
2
))
2
+(a sin(
1
x +
1
) + b sin(
2
x +
2
))
2
=
_
a
2
+ b
2
+ 2ab cos[(
1

2
)x +(
1

2
)]. (5.11)
We also have
(x) = arctan[a sin(
1
x +
1
) +b sin(
2
x +
2
), a cos(
1
x +
1
) + b cos(
2
x +
2
)]. (5.12)
The function arctan(x, y ) in Eq. 5.12 is dened as arctan(x, y ) = tan
1
_
y
x
_
, but has range
[, ).
Now we can write the sum of the two sinusoids as
a cos(
1
x +
1
) + b cos(
2
x +
2
) = A(x) cos (x). (5.13)
The sum is an oscillation whose amplitude is given by Eq. 5.11 and whose instantaneous
phase is given by Eq. 5.12. A(x) is periodic and has a frequency
1

2
. The instantaneous
frequency of the oscillation, that of cos (x), is given by
(x) =
d
dx
(x)
= +
(
1
)a
2
+ (
2
)b
2
a
2
+b
2
+2ab cos[(
1

2
)x + (
1

2
)]
= +
(
1
)a
2
+(
2
)b
2
A
2
(x)
, (5.14)
52
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
where
=

1
+
2
2
. (5.15)
It is clear that (x) is a periodic function in x and has the same period as A(x). Also, (x)
has either the same initial phase or the opposite initial phase as A(x). In Figure 5.12, A(x)
and (x) are in phase. The oscillation has a higher frequency when its amplitude is higher.
As a validity check, let a = b. We have
A(x) = 2a

cos
(
1

2
)x +
1

2
2

and
cos(x) =
_

_
cos
(
1
+
2
)x + (
1
+
2
)
2
, if (4n 1) (
1

2
)x +
1

2
< (4n +1),
cos
(
1
+
2
)x + (
1
+
2
)
2
, otherwise,
where n is any arbitrary integer. A(x) cos (x) then reduces to the familiar formula:
2a cos
(
1

2
)x +
1

2
2
cos
(
1
+
2
)x + (
1
+
2
)
2
.
Eqs. 5.115.15 can be readily extended to include any number of sinusoids. Let
i
(x) =

i
x +
i
, where i 1. It can be shown that
n

i=1
a
i
cos
i
(x) = A(x) cos (x), (5.16)
where
A(x) =
_
_

a
i
cos
i
(x)
_
2
+
_

a
i
sin
i
(x)
_
2
=
_

a
2
i
+ 2

i<j
a
i
a
j
cos[
i
(x)
j
(x)], (5.17)
(x) = arctan
_

a
i
sin
i
(x),

a
i
cos
i
(x)
_
, (5.18)
and
(x) =
d
dx
(x)
= +

(
i
)a
2
i
+

i<j
a
i
a
j
(
i
+
j
2 ) cos[
i
(x)
j
(x)]
A
2
(x)
, (5.19)
53
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
where
=
1
n

i
. (5.20)
Both A(x) and (x) become very complicated for an arbitrary number of sinusoids. However,
we know that sinusoids whose frequencies are far dierent from one another do not beat
to create appreciable pattern of low frequencies. In the spectrum of a practically sampled
halftone, there are probably only three signicant frequency components that are lined up
and close to one another. In the next subsection we will examine the sum of three sinusoids
using Eqs. 5.165.20.
Next we extend Eqs. 5.165.20 to two dimensions. We consider a sum of sinusoids of
the form
n

i=1
a
i
cos
i
(x), where
i
(x) =
i
x +
i
, i = 1, 2, . . . and
i
= (
i,x
,
i,y
).
Notice that the functions cos and sin take only one argument, the instantaneous phase,
and in our sum of sinusoids in 1D the instantaneous phase of a sinusoid is
i
(x), a function
of one variable. Here the instantaneous phase is
i
(x). We need only to revise Eqs. 5.16
5.20 by replacing x with x,
i
(x) with
i
(x), and one derivative to two partial derivatives
to obtain the equations in 2D. So we have
n

i=1
a
i
cos
i
(x) = A(x) cos (x),
where
A(x) =
_
_

a
i
cos
i
(x)
_
2
+
_

a
i
sin
i
(x)
_
2
=
_

a
2
i
+ 2

i<j
a
i
a
j
cos[
i
(x)
j
(x)],
(x) = arctan
_

a
i
sin
i
(x),

a
i
cos
i
(x)
_
,

x
(x) =

x
(x)
=
x
+

(
i,x

x
)a
2
i
+

i<j
a
i
a
j
(
i,x
+
j,x
2
x
) cos[
i
(x)
j
(x)]
A
2
(x)
,
54
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
and

y
(x) =

y
(x)
=
y
+

(
i,y

y
)a
2
i
+

i<j
a
i
a
j
(
i,y
+
j,y
2
y
) cos[
i
(x)
j
(x)]
A
2
(x)
,
where

x
=
1
n

i,x
and

y
=
1
n

i,y
5.7.2 The geometric shape of a moire pattern in 1D
Now we examine the 1D analogy of a moire pattern that is a curve consisting of three
clusters of close frequency components. The rst cluster is centered at the origin of the
frequency space and has three components. The component at the origin corresponds to a
constant in the spatial space and the other two components are a complex conjugate pair
corresponding to a low frequency sinusoid. The other two clusters are complex conjugates
and represent one set of sinusoids with frequencies close to one another but each being
signicantly higher than the frequency of the sinusoid in the rst cluster. We consider
two cases for the two high frequency clusters: having two components or having three
components, corresponding to having two sinusoids or three sinusoids. More specically, we
examine the curves for
g(x) = a
0
+ a
1
cos(
1
x +
1
) +a
3
cos(
3
x +
3
) + a
4
cos(
4
x +
4
) (5.21)
and
g(x) = a
0
+a
1
cos(
1
x+
1
) +a
2
cos(
2
x+
2
) +a
3
cos(
3
x+
3
) +a
4
cos(
4
x+
4
) (5.22)
where a
0
, a
1
, a
2
, a
3
> 0,
1
=
3

2
=
4

3
,
1

2
,
3
,
4
, and
1
= (
3

2
) mod =
(
4

3
) mod . Figure 5.15 illustrates the Fourier spectrum of g(x) in magnitude. The
55
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES

3
|G(u)|

4
u
Figure 5.15: The Fourier spectrum of the sum of four sinusoids
spectrum shows the distribution of frequencies
1
,
2
,
3
, and
4
.
The second term a
1
cos(
1
x+
1
) in Eq. 5.21 corresponds to the component of an aliasing
halftone frequency harmonic that comes close to the zero frequency. It is a periodic pattern
in density with frequency
1
. The third and fourth terms in Eq. 5.21 interfere to create a
contrast pattern. More importantly, a
3
cos(
3
x +
3
) + a
4
cos(
4
x +
4
) can be written as
A(x) cos (x) where A(x) =
_
a
2
3
+a
2
4
+ 2a
3
a
4
cos(
1
x +

), and

=
4

3
=
1
mod .
The envelope A(x) has the same frequency
1
as the low frequency density pattern, and
the envelope has either the same or the opposite initial phase as that of density pattern.

1
is the frequency of the moire pattern. Figure 5.16(a) shows the sum of the rst two
terms of Eq. 5.21, Figure 5.16(b) shows the sum of the last two terms and their envelope
for

=
1
, and Figure 5.16(c) shows the sum of all four terms. In this case,
1
is the
initial phase of the moire. Recall from Section 5.4 that
1
is determined by 1) the halftone
lattice, the sampling lattice, and their relative orientation and oset and 2) the density of
the graphical content that the halftone represents. To the human eye, the contrast pattern
reinforces the density pattern. If

=
1
+ , the contrast pattern will compensate the
density pattern, and the initial phase of the moire is not well dened.
The sum of the last three terms a
2
cos(
2
x +
2
) + a
3
cos(
3
x +
3
) +a
4
cos(
4
x +
4
)
in Eq. 5.22 also produces a contrast pattern. The sum can be written as A(x) cos (x), and
it can be shown that A(x) is of the form
_

2
+cos(
1
x +

)[ + cos(
1
x +

)], where

=
1
mod , and , , and are constants dependent upon a
2
, a
3
, and a
4
. A(x) is an
56
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
(a)
a density pattern
(b)
a contrast pattern of two sinusoids
(c)
sum of the density pattern in (a) and the contrast pattern in (b)
(d)
a contrast pattern of three sinusoids
(e)
sum of the density pattern in (a) and the contrast pattern in (d)
Figure 5.16: 1D moire: density pattern, contrast pattern, combined pattern
57
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
oscillation with a base frequency of
1
, and it may also contain a secondary oscillation with
frequency 2
1
. Figure 5.16(d) shows a typical curve that A(x) may exhibit, together with
the corresponding contrast pattern. Figure 5.16(e) shows the combined moire, the sum of
all ve terms in Eq. 5.22.
5.7.3 The geometric shape of a moire pattern in 2D
Here we consider surfaces of sums of sinusoids whose frequencies form square clusters in
the frequency space, similar to Figure 5.7. Not only does each cluster have a square shape,
but the frequencies within the cluster also form smaller squares because of the assumption
of square halftone lattice and square sampling lattice. The components (sinusoids) at these
frequencies have the same phase dierence property of Theorem 5.4.1. We rst plot two
surfaces corresponding to the aliasing cluster centered at zero frequency, each under a
dierent assumption. We next plot a surface corresponding to a high frequency, non-
aliasing cluster. Lastly, we plot a surface corresponding to two non-aliasing clusters. We
do not plot the combined surface for the sum of the aliasing cluster and the non-aliasing
clusters, because the combined surface is too complicated to provide much insight. The
analysis in this section applies to non-orthogonal halftone and/or sampling lattices that
have equal spacing in both directions. The general halftone lattice and sampling lattice
case is analogous but more complicated.
The nine frequency components in the aliasing cluster come from the nine replicas
closest to zero frequency. They represent a constant and four sinusoids. Sometimes the four
components from the four second order replicas are insignicant compared to the ones from
the four rst order replicas. Thus, we consider two cases, assuming that the components
from the second order replicas are insignicant in the rst case and signicant in the second
case. Specically, we plot
g
d
(x) = a
0
+ a
1
cos(
1
x +
1
) + a
2
cos(
2
x +
2
)
and
58
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
u

3
4
v

6
7
8
5
v
(a) (b)
Figure 5.17: Frequency component distribution of (a) an aliasing and (b) a non-aliasing
frequency clusters
g
d
(x) = a
0
+ a
1
cos(
1
x +
1
) + a
2
cos(
2
x +
2
) +a
3
cos(
3
x +
3
)
+a
4
cos(
4
x +
4
),
where
1

2
, [
1
[ = [
2
[,
3
=
1
+
2
,
4
=
1

2
,
3
=
1
+
2
, and
4
=
1

2
.
Note that in more general cases,
3
= (
1
+
2
) mod , and
4
= (
1

2
) mod . Fig-
ure 5.17(a) shows the aliasing cluster in the frequency space. Figures 5.18(a) and (b) show
respectively the two g
d
(x) surfaces in the same scale. They are density patterns. The two
surfaces have similar hills, but the valleys in Figure 5.18(a) are lled in Figure 5.18(b)
by a
3
cos(
3
x +
3
) + a
4
cos(
4
x +
4
), and the hills are reinforced by the same two
terms.
A non-aliasing cluster has a complex conjugate cluster. Each component in a non-
aliasing cluster (together with its complex conjugate component in the conjugate cluster)
represents a sinusoid. Usually, only four components in the cluster, which form a small
square in the frequency space, are signicant. Therefore, we look at the surface formed by
the four strong components. Specically, we plot
g
c
(x) = a
5
cos(
5
x +
5
) +a
6
cos(
6
x +
6
) +a
7
cos(
7
x +
7
)
+a
8
cos(
8
x +
8
),
59
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
x
y
g_d(x,y)
x x
y
g_d(x,y)
x
(a) (b)
x
y
g_c(x,y)
x
(c)
x
y
g_c(x,y)
x
(d)
Figure 5.18: 2D moire: density patterns and contrast patterns (part (d) enlarged to show
detail)
60
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
where
6

5
=
7

8
=
1
,
7

6
=
8

5
=
2
,
6

5
=
7

8
=
1
, and

7

6
=
8

5
=
2
. Figure 5.17(b) shows a non-aliasing frequency cluster. Using the
results in Section 5.7.1 it can be shown that g
c
(x) has an envelope A(x) of the form
A(x) = [
2
0
+
1
cos(
1
x +
1
) +
2
cos(
2
x +
2
) +
3
cos(
3
x +
3
)
+
4
cos(
4
x +
4
)]
1
2
,
where
0

4
are constants dependent on a
5
a
8
. The envelope has the same frequencies
and initial phases as the density pattern produced by the aliasing cluster. Figure 5.18(c)
shows the surface of a typical g
c
(x), which is a contrast pattern. The moire frequency
consists of two vectors:
1
and
2
. The initial phase of the moire also consists of two
numbers:
1
and
2
.
There are four strong high frequency clusters in Figure 5.7. They represent two clusters
of sinusoids. We considered one cluster above. It is easy to see that the other cluster of
sinusoids have the same low frequency envelope as the cluster we considered, but they have
a dierent high frequency oscillation pattern that is in the direction perpendicular to that of
the high frequency oscillation of the cluster we considered. The eect of the two clusters of
sinusoids together is a crisscross high frequency pattern modulated by the low frequency
envelope A(x) above. Figure 5.18(d) shows the eect. The high frequency crisscross pattern
is simply the reection of the halftone dots.
In 2D, it is also possible for the density pattern and the contrast pattern to be out of
phase, which may occur in three ways.
5.8 The half period phase reversal of a moire pattern
Experiments show that a moire pattern in a sampled general halftone may exhibit a half
period phase reversal in areas where the image content has dierent densities. Figure 5.19
shows a simulated sampled halftone. The halftone represents a picture of two gray levels.
The moire pattern exhibits a phase reversal near the center in the vertical direction. In
Figure 5.11(a), which shows an image resampled from the scanned halftone House, the
61
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
Figure 5.19: The phase reversal of a moire pattern
moire pattern in the cloud area also exhibits a phase reversal to the moire pattern in
other darker areas. In Figure 8(a) of Shu, Springer, and Yeh [SSY89], the phase reversal
phenomenon is also visible, but not discussed. In the following, we explain this phenomenon
in 1D in detail for simplicity.
In a halftone picture the density of image content is represented by the size of halftone
dots. Recall that in frequency space the size of the halftone dots determines the shape and
polarity of the envelope that modulates the Fourier spectrum of the halftone.
Consider a uniform halftone rst. Let h(x) be a uniform halftone that represents a
density level f
0
, 0 f
0
1, and let the halftone frequency be p and the halftone period be
a, p =
1
a
. Let the corresponding dot size (width) be ( = f
0
a a). We have
h(x) =
_

m=
(x ma)
_
c(x) =

m=
c(x ma),
where
c(x) =
_
1, [x[ <

2
0, otherwise
62
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
h(x)
a

x
Figure 5.20: 1D halftone representing a uniform gray
denes one halftone cell and has a Fourier transform C(u) of the form
C(u) = sinc(u).
Figure 5.20 illustrates h(x). The Fourier transform H(u) of h(x) is
H(u) =
_

m=
(u mp)
_
C(u) =
_

m=
(u mp)
_
sinc(u).
Figure 5.21 shows H(u) for two values of f
0
. The short dashed vertical bars show the
spectrum for f
0
= 0.75, and the long dashed vertical bars show the spectrum for f
0
= 0.35.
The short dashed curve shows the envelope C(u) for f
0
= 0.75, and the long dashed curve
u
|H(u)|
p
0.35
p 2p
0.75
p
Figure 5.21: The Fourier spectra of two uniform halftones of density 0.35 and 0.75
63
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
u
s
H (u)
p s
2p
(a)
u
s
H (u)
p 2p s
(b)
Figure 5.22: The Fourier spectra of two sampled uniform halftones: (a) density 0.75 and
(b) density 0.35
shows the envelope for f
0
= 0.35. Note that the vertical bars are at the harmonics of the
halftone frequency p, and that the zeros of C(u) occur at multiples of
1

.
The frequency components at the second order halftone frequency harmonics have op-
posite signs for the two spectra. Hence, even though both halftones are the synthesis of
sinusoids of the same frequencies, the synthesis requires opposite sign coecients for some
of the sinusoids, including the one at the second harmonic of the halftone frequency.
When the two halftones are sampled at a frequency s that is barely more than twice
the halftone frequency, aliases of the second order halftone frequency harmonics come close
to the zero frequency. Since they have opposite signs, they produce density patterns of
the opposite polarity. Figures 5.22(a) and (b) show respectively the spectra of the sampled
halftones.
If the sampling frequency is barely more than twice the halftone frequency, phase reversal
occurs once to the density pattern: when the frequency component at a second halftone
64
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
frequency harmonic changes sign at f
0
=
1
2
. The contrast pattern has a constant phase
because it is formed by components at the rst order halftone frequency harmonics, which
are always positive.
If the sampling frequency is barely more than three times the halftone frequency, aliases
of the third order halftone frequency harmonics come close to the zero frequency, and
aliases of the second harmonics come close to the rst order harmonics. Phase reversal
occurs twice to the density pattern: when the frequency component at a third halftone
frequency harmonic changes sign at f
0
=
1
3
and at f
0
=
2
3
. The contrast pattern formed by
the components at the rst and second harmonics exhibits phase reversal once when the
component at a second order harmonic changes sign at f
0
=
1
2
. Note that in this sampling
frequency case, the aliased second order harmonics contributing to the contrast pattern are
aliasing frequencies, and the contrast pattern is an aliasing pattern.
For a general halftone representing a picture that has dierent densities in dierent areas,
all halftone dots share the same halftone lattice, but their sizes are signicantly dierent in
areas where densities are distinctively dierent. The moire pattern in the sampled halftone
may then exhibit phase reversal in those areas.
The cause of the phase reversal phenomenon also becomes clear if we examine closely
the scanned House halftone in Figure 5.9 again. In light areas, there are individual black
dots on a white background. In dark areas, however, black dots are large and connected to
one another, and the halftone looks as if it consists of white dots on a black background.
The black dots in light areas and the white dots in dark areas are in opposite phase.
It is interesting to know what spectrum a halftone representing dierent density features
may have. It can be shown from Eq. 5.9 that if we compute the Fourier transform of a region
of the halftone that has similar densities, the resultant spectrum has a peak component at
a halftone frequency harmonic, and the component has a certain polarity. The region is
similar to a uniform halftone. For example, for a region where densities are below 0.5, the
spectrum of the region has a peak at a second order harmonic of the frequency harmonic.
However, if we compute the Fourier transform of the entire halftone, the resultant spectrum
65
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
may not have a peak at the same harmonic frequency.
Consider a halftone that represents two levels of gray, as shown in Figure 5.23(a). The
halftone has a frequency p = 1 and a period a = 1. It has a smaller dot size
1
= 0.35a
for x < 0 and a larger dot size
2
= 0.75a for x > 0. Since the Fourier transform of this
halftone has innite magnitude at many frequencies, and it is dicult to plot and compare
two innite magnitudes, we compute the Fourier transform H(u) of a w < x < w segment
of the halftone. It can be shown that
H(u) =

m=
_

1
sinc(
1
mp) e
jw(ump)
+
2
sinc(
2
mp) e
jw(ump)
_
wsinc[w(u mp)]
Figure 5.23(b) shows the magnitude of H(u) for w = 10. Figures 5.23(c) and (d) show,
respectively, the close-up of [H(u)[ near the +1 and the +2 halftone frequency harmonics.
The Fourier spectrum of a sampled general halftone may not have a peak at a halftone
frequency harmonic of order higher than 1. This is because sinusoids of dierent polarities
are needed at that harmonic to synthesize dierent regions of the halftone that have dierent
densities. The Fourier transform of the entire halftone provides a set of sinusoids near the
harmonic. The set of sinusoids collaborate to behave as one sinusoid with a certain initial
phase in each density region. The frequency of the sinusoid that has the largest magnitude
in the set may not equal the harmonic of the halftone frequency.
In 2D, suppose that the halftone lattice and the sampling lattice are square, and that the
halftone dots are at the centers of their cells and are symmetric horizontally and vertically.
Then the moire phase reversal occurs in the two directions of the moire pattern grid at
the same time. Otherwise, the phase reversal may occur in one direction. The example in
Figure 5.19 has square halftone lattice and sampling lattice, but the halftone dots are not
symmetric because of the digital halftoning algorithm used. The phase reversal occurs in
one direction for the two particular density levels.
66
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
-4 -2 2 4
x
1
h(x)
(a)
-3 -2 -1 1 2 3
u
|H(u)|
(b)
0.8 0.9 1.1 1.2
u
|H(u)|
1.8 1.9 2.1 2.2
u
|H(u)|
(c) (d)
Figure 5.23: A halftone of two levels of gray and its Fourier spectrum
67
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
5.9 Summary
The spectrum of a sampled uniform halftone consists of impulses at the halftone fre-
quency harmonics and their aliases. These impulses may form clusters in the frequency
space. All clusters have the same geometry, that is, impulses in one cluster have the same
set of frequency osets with respect to one another as any other clusters. Further, under
modulo arithmetic, impulses in one cluster have the same set of initial phase osets with
respect to one another as any other clusters. The phase oset property holds if two assump-
tions are true: 1) the halftone dots are symmetric with respect to the center of their cells,
and 2) the low-pass function has a linear phase shift. Both assumptions are reasonable in
practice.
The low-pass ltering by a real scanner reduces the number of signicant impulses in
the clusters in the spectrum of a scanned halftone, so the clusters are more distinct.
The impulse cluster at the origin of the frequency space corresponds to a periodic density
pattern. The other higher frequency clusters correspond to a periodic contrast pattern. The
envelope of the contrast pattern has the same frequency as the density pattern. The density
pattern and the contrast pattern may be in phase or out of phase, depending upon the gray
level that the uniform halftone represents. In the literature, the term moire pattern usually
refers to the density pattern.
The impulse cluster at the origin consists of aliasing impulses. The most prominent high
frequency clusters usually include non-aliasing impulses. There may exist less prominent
high frequency clusters that include aliasing components.
An aliasing moire usually consists of both a density pattern and a contrast pattern. A
non-aliasing moire is always a contrast pattern.
The spectrum of a sampled general halftone consists of peaks at the rst order halftone
frequency harmonics and their aliases. The spectrum also consists of hills around higher
order halftone frequency harmonics and their aliases. The peaks and hills may form clusters
similar to those in the spectrum of a sampled uniform halftone.
68
CHAPTER 5. CHARACTERIZATION OF MOIRE IN SAMPLED HALFTONES
In a sampled general halftone, the moire pattern in each area of similar density has a
behavior similar to a moire pattern in a sampled uniform halftone. However, across areas
of dierent densities, the moire pattern may exhibit a half period phase reversal.
We use continuous Fourier transform to analyze a halftone and a sampled halftone. We
also assume that a halftone or a sampled halftone has innitely large size. These choices
make the results of analysis relatively simple while still revealing the essence of a moire
pattern. Given the continuous Fourier transform of a sampled halftone, it is procedural to
derive the discrete Fourier transform.
69
Chapter 6
Case Study: Moire Patterns in Scanned Halftones
by a Commercial Desktop Scanner
6.1 Introduction
In Chapter 5 we discuss a moire pattern in a sampled halftone with an important assump-
tion: samples are taken uniformly in the two directions in which sampling is performed. In
the literature, sampling theory is also mostly developed under the same assumption. How-
ever, for various reasons, a practical scanner may not deliver samples that are uniformly
spaced. As a result, the moire pattern in a scanned halftone image may also have charac-
teristics specic to the scanner. In this chapter we examine one popular desktop scanner,
the HP ScanJet IIc, and extend the results of Chapter 5 to scanned halftones from this
scanner. The approach used should be applicable to other scanners.
6.2 The image data processing procedure by the ScanJet IIc
In this section we describe part of the image data processing algorithm of the ScanJet
IIc that aects its sampling model. This information is non-proprietary and was obtained
mainly through personal communication between the author and Mr. Greg Degi at Hewlett-
Packard Company ([Deg95]). Part of the information can also be found in Degi and Buck
[DB93]. General information about a CCD sensor array, a key component of most scanners,
can be found in sources such as Optoelectronics Data Book ([Tex83]).
The ScanJet IIc system, including both the scanner and the accompanying support soft-
ware installed on the host computer, is capable of producing scanned images of resolutions
from 12 to 800 dpi. We call the resolution of an output image the target resolution. The dis-
70
CHAPTER 6. CASE STUDY
cussion in this chapter is restricted to target resolutions in the range of 60400 dpi because
it is the most common range and because the ScanJet IIc has a native optical resolution
of 400dpi. Any target resolution higher than 400 dpi is realized through interpolation and
thus leads only to a larger image size but not truly higher resolution in image content. All
the experiments conducted in this dissertation use resolutions in the 60400 dpi range. In
this range, the ScanJet IIcx, the enhanced model of the ScanJet IIc, has exactly the same
image data processing procedure.
6.2.1 The optical sampling of the ScanJet IIc
The ScanJet IIc samples an input picture dierently in the horizontal direction and
the vertical direction. The sampling in the horizontal direction is realized by an optical
sampling followed by a digital subsampling while the sampling in the vertical direction is
achieved by optics only.
The ScanJet IIc uses a three-line CCD array as the image acquisition device. Each
line senses one primary color through a color lter. Each line has about 3400 photosites
to sample a width of about 8.5 inches in the horizontal direction and achieve an optical
sampling rate of 400 dpi. A lens-mirror optical system focuses the 8.5 inch eld of view onto
the CCD array. The optical system is so designed that it also provides low-pass ltering that
eectively lters out frequencies higher than 200 dpi, half the horizontal optical sampling
rate.
The CCD array and the optical system is mounted on a carriage that moves across the
vertical length of the input picture. When scanning a picture, the CCD array is read and
cleared periodically while the carriage is moving, thus achieving the sampling in the vertical
direction. The CCD cells have a required exposure time to collect enough charges for each
read. For a particular target resolution, the CCDoptics carriage is set to move at a speed
that provides the CCD array with the necessary exposure time over each scan line. The
carriage does not stop when the CCD cells are open for exposure but continues moving, thus
defocusing the image vertically and providing low-pass ltering. A lower target resolution
71
CHAPTER 6. CASE STUDY
requires a faster CCD motion which in turn provides a stronger low-pass ltering needed
by the lower target resolution.
In summary, the optical sampling produces a digital image sampled uniformly at 400
dpi horizontally and at the desired resolution vertically. The input picture is suciently
low-pass ltered by the optical system in the horizontal direction and is low-pass ltered by
both the optics and motion blur in the vertical direction. Experiments show that aliasing
from optical sampling is usually weak in the vertical direction.
6.2.2 Resolution reduction of an optically sampled row of pixels
To produce a row of pixels at the target resolution, the ScanJet IIc subsamples an opti-
cally sampled row of pixels using an improved binary rate multiplication (IBRM) algorithm.
IBRM has three steps [DB93]:
1. Linear interpolation of the 400 dpi raw data to form an intermediate row of pixels at
800 dpi (herewith referred to as 2 interpolation).
2. Selection of pixels from the 800 dpi interpolated row using binary rate multiplication
(BRM) to form a second intermediate row of pixels whose average resolution is twice
the target resolution.
3. Selection of every other pixel from the second intermediate row to obtain the output
row.
Prior to the IBRM algorithm, digital low-pass ltering by a two-pixel moving average is
performed on the 400 dpi raw data if the target resolution is below 300 dpi, and a four-
pixel moving average is computed if the target resolution is below 150 dpi. Figure 6.1(a)
summarizes the processing steps that image data goes through in the horizontal direction.
For comparison, the processing steps that image data goes through in the vertical direction
is illustrated in Figure 6.1(b).
72
CHAPTER 6. CASE STUDY
optics
low-pass
low-pass
moving
sampling
graphic
input
column
moving output
column of
pixels
(a)
optical
low-pass
CCD
sampling
digital
low-pass
2x inter-
BRM
selection
alternating graphic
row
input
output
row of
pixels
(b)
polation
IBRM
Figure 6.1: The image data processing procedure of ScanJet IIc in horizontal direction (a)
and in vertical direction (b)
The pixels in the 800 dpi interpolated row are uniformly spaced with respect to the
input picture. The pixels selected by BRM may not have uniform spacing with respect to
the input picture, nor may the pixels in the output row.
The binary rate multiplication, as used in the ScanJet IIc, works as follows. Let B =
0.b
1
b
2
b
3
b
4
b
5
b
6
b
7
b
8
be the 8 bit binary representation of the ratio
R =
2 target resolution
800dpi
.
Let the pixels in the interpolated row be indexed sequentially by binary integers starting
from 0, and let the lowest eight bit of the index be i = i
7
i
6
i
5
i
4
i
3
i
2
i
1
i
0
. A pixel is selected
by BRM if
a)b
1
= 1 and i
0
= 0 or
b)b
2
= 1 and i
1
i
0
= 01 or

h)b
8
= 1 and i
7
i
6
i
5
i
4
i
3
i
2
i
1
i
0
= 01111111
Condition a) selects half of the pixels in the interpolated rowpixels at indices 0,2,4,6,8,. . . ,
condition b) selects
1
4
of the pixelsthose at indices 1,5,9,13, . . . , and so forth. Conditions
a)h) are disjoint and their union selects all pixels but the ones with i = 255, which are
selected only when R = 1. If only one bit in B is 1, the selected pixels are uniformly spaced.
73
CHAPTER 6. CASE STUDY
pixels
1 3 5 7 9 11 13 15 17 19 21 23 25 27
800dpi row
row by BRM
selection
row by alternating
selected by b =1
3
pixels
2
selected by b =1
Figure 6.2: IBRM steps to produce an output row at 150 dpi
Otherwise, the selected pixels are non-uniformly but recurrently spaced. If only b
1
and b
2
are 1, the spacing between the selected pixels corresponds to nearest neigh bor sampling. In
all other cases, spacing is less uniform than for nearest neighbor sampling.
As an example of the IBRM in the ScanJet IIc, assume that a target resolution of 150
dpi is requested. IBRM performs BRM on the 800 dpi interpolated row of pixels for an
intermediate row whose average resolution is 300 dpi. The ratio of 300 dpi to 800 dpi is
2150
800
=
1
4
+
1
8
= 0.01100000
(2)
. Therefore, BRM produces an intermediate row of pix-
els whose indices in the interpolated row are 1, 3, 5, 9, 11, 13, 17, 19, 21, . . .. The spacing
between the selected pixels (in the 800 dpi row) has a pattern of 2, 2, 4, 2, 2, 4, . . .. The
alternating selection step then produces an output row 1, 5, 11, 17, 21, . . .. The pixels in
the output row have a spacing pattern of 4, 6, 6, 4, 6, 6, . . . and are more evenly spaced than
the intermediate row of pixels selected by the BRM. This example is illustrated in Fig-
ure 6.2, where a circle represents a pixel, and pixels with odd indices in the interpolated
row are marked. As another example, assume that a target resolution of 100 dpi is re-
quested. The BRM delivers an intermediate row 1, 5, 9, 13, 17, . . ., which is uniform. The
alternating selection produces an output row 1, 9, 17, 25, . . ., which is also uniform. The
IBRM produces uniform spaced pixels if the target resolution is a power of two fraction of
800 dpi.
74
CHAPTER 6. CASE STUDY
Resolution conversion is the main subject of the research area multirate digital signal
processing (MDSP) ([CR83, Vai93]). The main idea is that given input data of some reso-
lution and a desired output resolution, the input data is interpolated to a resolution that
is a common multiple of both the input and the output resolutions (interpolation). The
interpolated data is then low-pass ltered to eliminate frequencies that are higher than
half the output resolution. Finally, output data samples are selected uniformly from the
interpolated and low-pass ltered data at the output resolution (decimation).
The IBRM algorithm used by the ScanJet IIc is a binary approximation to the stan-
dard MDSP approach in that interpolation and decimation are done only at rates that are
combinations of powers of two. The digital low-pass ltering by the ScanJet IIc is also a
coarse quantization of that of the MDSP approach. The IBRM is a compromise between
eciency/cost and quality.
6.3 The Fourier spectrum of scanned halftones from the ScanJet IIc
When the target resolution is 400 dpi, the optical sampling is accompanied by sucient
low-pass ltering; no noticeable aliasing occurs. Figures 5.9 and 5.10 show such an example
in which the 85 dpi House halftone in Figure 5.8 is scanned at 400 dpi. At 400 dpi no
digital low-pass ltering is performed, and digital subsampling has no eect.
When the target resolution is lower than 400 dpi, digital subsampling goes into eect
in the horizontal direction. Aliasing may become appreciable in the horizontal direction
because the amount of the digital low-pass ltering may not be sucient for the particular
subsampling resolution. Figure 6.3 shows the same House halftone scanned at 150 dpi.
We will use the term nominal sampling frequency to refer to the target resolution. Fig-
ures 6.4(a) and (b) show in 3D and 2D, respectively, the portion of the Fourier spectrum
of the scanned halftone within the Nyquist band of the nominal sampling frequency. The
halftone frequency falls in the band, which is reected by the strong frequency components
at the rst order halftone harmonics of the zeroth order replica in the spectral band shown.
75
CHAPTER 6. CASE STUDY
Figure 6.3: House halftone scanned by ScanJet IIc at 150 dpi
76
CHAPTER 6. CASE STUDY
u
v
|H(u,v)|
u
(a)
(b)
Figure 6.4: The Fourier power spectrum of the scanned halftone in Figure 6.3
77
CHAPTER 6. CASE STUDY
Very weak aliasing frequency components appear in the vertical direction, but aliasing fre-
quency components are clearly present in the horizontal direction. The distribution of these
frequency components appears to be more complex than that discussed in Chapter 5.
In the remainder of this section we will explain that the complexity just mentioned
results from the non-uniform subsampling. We rst analyze the case of 150 dpi target
resolution as in the example in Figures 6.3 and 6.4. Then we extend the result to other target
resolutions. For 150 dpi target resolution we begin with the 1D Fourier spectrum of one row
of the scanned image and then extend the result to the 2D spectrum of the entire scanned
image. We also show the interesting phenomenon that, due to non-uniform sampling, the
frequency component at the center of a replica may always have zero magnitude even though
components at other frequencies of the same replicas do not.
6.3.1 The Fourier spectrum of one row of an image scanned at 150 dpi
The 800 dpi interpolated row can be considered to be the input picture sampled at 800
dpi preceded by low-pass ltering; the argument is as follows:
1. Because the 400 dpi CCD sampling is preceded by sucient (optical) low-pass ltering
to prevent aliasing, the digital low-pass ltering and the CCD sampling is functionally
commutable. In other words, the 400 dpi CCD sampling box and the digital low-
pass box in Figure 6.1(a) can be conceptually switched. Of course, the digital low-
pass lter must be replaced with an equivalent analog lter. Figure 6.5 illustrates the
equivalent steps to obtain an 800 dpi interpolated row. Parts (a) and (b) of the gure
show the above commutation.
2. 400 dpi sampling followed by 2x interpolation is equivalent to sampling at 800 dpi
preceded by some special low-pass ltering ([OS89]). Figure 6.5 parts (b) and (c)
illustrate this equivalence.
3. Multiple cascaded linear lters are functionally equivalent to a single lter. Figure 6.5
parts (c) and (d) show that the optical low-pass lter, the digital low-pass lter, and
78
CHAPTER 6. CASE STUDY
optical
low-pass
digital
low-pass
2x inter- graphic
row
input
polation
CCD
400dpi
sampling
row
c)
800dpi
optical
low-pass
2x inter- graphic
row
input
polation
low-pass
analog
to digital
equivalent row
800dpi
400dpi
sampling
optical
low-pass
800dpi
row
input
sampling
low-pass
analog
to digital
equivalent row
800dpi
low-pass
special
800dpi graphic
row
input
sampling row
800dpi equivalent to the
above three
low-pass filter
cascaded filters
a)
b)
d)
graphic
Figure 6.5: Equivalent low-pass ltering and sampling models for obtaining the 800 dpi
interpolated row of pixels
the 2x interpolation lter, each of which is linear, are cascaded to form one lter
preceding the 800 dpi sampler.
Let f(x) be the prole of the input picture at the scan row, let lp(x) be the low-pass
lter equivalent to the three cascaded low-pass lters in Figure 6.5, let f
2s
(x) be the 800
dpi interpolated row of pixels, and let x be measured in
1
800
inches. We can write
f
2s
(x) = [f(x) lp(x)]

m
(x m) = g(x)

m
(x m),
where the summation of functions represents the 800 dpi sampler. f
2s
(x) can be thought
of as sampled from g(x) = f(x) lp(x), the low-pass ltered version of f(x).
The output row of pixels is a subset of pixels of the 800 dpi interpolated row. When
the target resolution is 150 dpi, the output row of pixels have a periodic spacing pattern of
4, 6, 6, 4, 6, 6, . . ., with a period of 16 pixels in the interpolated row. Let g
s
(x) represent the
output row of pixels whose positions are measured in the interpolated row. g
s
(x) can be
79
CHAPTER 6. CASE STUDY
considered as directly sampled from g(x) with non-uniform spacing:
g
s
(x) = g(x)

m
(x x
m
),
where
x
m
=
_

_
16k + 0, m = 3k,
16k + 4, m = 3k + 1,
16k + 10, m = 3k + 2,
(6.1)
and k is an integer. We dene
q
i
=
_

_
0, i = 0,
4, i = 1,
10, i = 2,
and call q
0
, q
1
, q
2
the sampling pattern because it gives the relative positions of one period
of the non-uniformly spaced samples.
It is more convenient to adopt the nominal sampling period corresponding to the 150
dpi nominal sampling frequency as the unit of measure for x. In this unit, the nominal
sampling frequency is one sample per unit, and the nominal sampling period is one unit
length. With x measured in this new unit, Eq. 6.1 becomes
x
m
= 3k +q
i
where i = m mod 3, and
q
i
=
_

_
0, i = 0,
0.75, i = 1,
1.875, i = 2.
The output rowof pixels nowhas a spacing pattern of 0.75, 1.125, 1.125, 0.75, 1.125, 1.125, . . .,
with a period of 3. If the samples were uniformly spaced, we would have q
i
= i.
A standard technique to study non-uniform but recurrent samples is to partition them
into sequences of uniform samples. The row of pixels g
s
(x) can be considered to be the sum
of three sequences of uniformly sampled pixels, g
s
i
(x), i = 0, 1, 2; each sequence is sampled
with the same frequency of one third samples per unit but with a dierent shift:
g
s
i
(x) = g(x)

k
(x 3k q
i
).
80
CHAPTER 6. CASE STUDY
-1
sampled
pixels
displayed
pixels
0 1 2 3 4 5
g(x)
g (x)
g (x)
vs
q -3+q q q
3+q
0
3+q 3+q
2
s
0 1
2
1
g (x)
s
0
2
Figure 6.6: Non-uniform sampling and uniform display
g
s
(x) can then be written as
g
s
(x) =
2

i=0
g
s
i
(x) = g(x)
2

i=0

k
(x 3k q
i
).
The upper half of Figure 6.6 illustrates g(x) being sampled with the sampling pattern
q
0
, q
1
, q
2
to produce g
s
(x). Vertical bars represent the samples. Samples that belong to
g
s
0
(x) are marked out.
Once the row of pixels g
s
(x) is formed, it is just a dimensionless sequence. The row of
pixels is always viewed and processed as though the pixels were uniformly sampled, which
is rarely the case. The lower half of Figure 6.6 illustrates that samples are assumed to have
equal spacing when being displayed, thus forming a new function g
vs
(x) dierent from g
s
(x).
The dotted lines in the middle of the gure show the transition of role of the sampled
pixels. This transition can be modeled by snapping each sequence of uniformly sampled
pixels to the nearest nominal sampling points. We have
g
vs
(x) =
2

i=0
g
s
i
(x +q
i
i) (6.2)
=
2

i=0
g(x +q
i
i)

k
(x 3k q
i
) =

m
g(x x

m
)(x m),
81
CHAPTER 6. CASE STUDY
where
x

m
=
_

_
0, m mod 3 = 0,
0.25, m mod 3 = 1,
0.125, m mod 3 = 2.
From now on, without specic qualication, the term the Fourier spectrum of an output
image always means the spectrum of the image as it is displayed. It is amazing how tolerant
or insensitive the human eye is to such sample manipulationthe human eye does not seem
to notice any distortion of the image content if there is no signicant moire pattern.
The Fourier transform G
s
i
(u) of g
s
i
(x) is
G
s
i
(u) = G(u)
_
1
3

_
u
k
3
_
e
j2q
i
u
_
=
1
3

k
G
_
u
k
3
_
e
j2q
i
k
3
,
where G(u) is the transform of g(x). The Fourier transform G
vs
(u) of g
vs
(x) is
G
vs
(u) =
2

i=0
G
s
i
(u)e
j2(q
i
i)u
=
1
3

k
_
2

i=0
e
j2[q
i(u
k
3
)iu]
_
G
_
u
k
3
_
. (6.3)
For comparison, the transform G
s
(u) of g
s
(x) is
G
s
(u) =
2

i=0
G
s
i
(u)
=
1
3

k
_
2

i=0
e
j2q
i
k
3
_
G
_
u
k
3
_
.
Due to non-uniform sampling, in both G
vs
(u) and G
s
(u) the transform G(u) is replicated
at one third the nominal sampling frequency (50 dpi) apart, rather than at the nominal
sampling frequency (150 dpi) apart as in uniform sampling. The kth order replica in G
vs
(u)
is enveloped by
env
vs
k
(u) =
2

i=0
e
j2[q
i(u
k
3
)iu]
,
which generally has non-zero magnitude except at its center, u =
k
3
, where it becomes
env
vs
k
(
k
3
) =
2

i=0
e
j2(i
k
3
)
=
1 e
j2k
1 e
j2
k
3
=
_
3, k is a multiple of 3,
0, otherwise.
82
CHAPTER 6. CASE STUDY
-1
2
-(-)
3
1
-(-)
3
1
-
3
2
-
3
1
u
1
2
3
envelope magnitude
Figure 6.7: Magnitude of envelopes for sampling replicas in G
vs
(u)
Figure 6.7 shows the magnitude of env
k
(u) for four values of k. The two solid curves
are for k = 0 and 3. The short dashed curve is for k = 1, and the long dashed curve
for k = 2. A replica whose order is not a multiple of 3 is reduced to zero in mag-
nitude at its center. As a comparison, the kth order replica in G
s
(u) is enveloped by
env
s
k
(u) =
2

i=0
e
j2q
i
k
3
= env
vs
k
(0), which is a constant in u.
6.3.2 The Fourier spectrum of an image scanned at 150 dpi
Similar to one rowof a scanned image, the entire image can be considered as one sampled
from a low-pass ltered version of the input picture. Let g(x) be this low-pass ltered input
picture, and let G(u) be the Fourier transform of g(x).
The Fourier transform G
vs
(u) of the scanned image is G(u) replicated in both vertical
and horizontal directions. In the vertical direction, G(u) is replicated at the nominal sam-
pling frequency, 150 dpi, apart. In the horizontal direction, G(u) is replicated at a spacing
of 50 dpi. Horizontal replicas are modulated by non-constant envelopes.
We are now ready to re-examine Figure 6.4. First note that the spectrum in Figure 6.4
is computed using the DFT which regards input data points as uniformly spaced. The low-
pass ltering by the optical system and the moving cartridge essentially eliminates halftone
harmonics of orders higher than 1 in the vertical direction. The low-pass ltering by the
83
CHAPTER 6. CASE STUDY
optical system and by the circuitry eliminates halftone harmonics of orders higher than
2 and substantially attenuates the second order in the horizontal direction. Figure 6.8(a)
illustrates the distribution of signicant frequency components of G(u). Figure 6.8(b) shows
the distribution of signicant frequency components of G
vs
(u), where the non-lled circles
represent frequency components whose magnitude is reduced to zero by the modulating
envelopes. The dashed square represents the 75 dpi Nyquist band for the nominal sampling
frequency, and the dotted lines show to which replica some of the frequency components
belong. Signicant frequency components are those at the rst and second order halftone
harmonics from the zeroth, rst, second, and third replicas. The distribution of signicant
frequency components in Figure 6.4 is similar to that in the dashed square in Figure 6.8(b).
Experiments are performed to low-pass lter the scanned image in Figure 6.3 with dif-
ferent cut-o frequencies. These experiments show that there are two types of frequency
components in Figure 6.4 or Figure 6.8(b) that contribute most strongly to the moire. The
rst type includes components at the four second order halftone frequency harmonics of
the second and third replicas in the horizontal direction. They are aliasing components.
Two such components are encircled by the dashed oval in Figure 6.8(b). The second type
includes components at the rst order halftone frequency harmonics of the zeroth order
replica, the rst order replica in the vertical direction, and the third order replica in the
horizontal direction. These are non-aliasing components. Four such components are encir-
cled in the dashed circle in Figure 6.8(b). Other frequency components in Figure 6.4 or
Figure 6.8(b) contribute only to minor moire patterns because their magnitudes are low
and their frequencies are high.
The rst type of moire components have frequencies near 20 dpi and 30 dpi respectively,
and they form a pattern similar to the curve in Figure 6.9 which is the sum of two sinusoids
at the respective frequencies. This pattern should be considered a contrast pattern with
a 10 dpi envelope. However, since both contributing frequencies are relatively low, the
pattern can also be regarded to be an intensity pattern. The moire pattern in the cloud
area of Figure 6.3 indeed resembles the curve in Figure 6.9. The moire pattern in darker
84
CHAPTER 6. CASE STUDY
(b)
(a)
Figure 6.8: Distribution of signicant frequency components of (a) the low-pass ltered
input halftone and (b) the scanned halftone
85
CHAPTER 6. CASE STUDY
Figure 6.9: The low frequency moire pattern in the scanned halftone in Figure 6.3
areas demonstrates a half period phase reversal from that in the cloud area.
The second type of moire components form a contrast pattern. The contrast pattern
is the reection of the halftone dots. Figure 6.10 shows a h igh -pass ltered version of
the image in Figure 6.3, with a cut-o frequency that is just below the halftone frequency.
Figure 6.11 shows a portion of Figure 6.10 enlarged eight times. Clearly the moire pattern
is a contrast pattern. It is interesting that this high-pass ltered image still carries some
image information of the original picture. The contrast of the image content is very low, and
the image was contrast enhanced when reproduced. Recall that the frequency components
around each halftone frequency harmonic come from a non-linearly distorted copy of the
spectrum of the original picture. The interference between the dierent distorted copies
forms an image that resembles the original picture. Note that the high-pass ltered image
does not physically contain low-frequency components. The apparent low-frequency features
seen in the high-pass ltered image are formed by varying contrast.
Other aliasing components in Figure 6.4 or Figure 6.8(b) contribute only to secondary
moire patterns. Specically, the components at the rst order halftone frequency harmonics
of the rst order horizontal replica form a contrast pattern that has a 20 dpi envelope in the
horizontal direction. The components at the rst order halftone frequency harmonics of the
second order horizontal replica interfere with the two neighboring components at the rst
86
CHAPTER 6. CASE STUDY
Figure 6.10: The image in Figure 6.3 high-pass ltered to retain only frequency components
at the halftone frequency or higher
87
CHAPTER 6. CASE STUDY
Figure 6.11: A portion of the high-pass ltered image in Figure 6.10 enlarged eight times
88
CHAPTER 6. CASE STUDY
order halftone harmonic of the zeroth and third order replicas to form a 10 dpi envelope on
the 30 dpi envelope formed by the latter two components.
It can be shown that env
vs
k
(u) has a complex phase linear in u, and for dierent values
of k, the phases of env
vs
k
(u)s only dier by a constant. Therefore, the same phase dierence
property of Theorem 5.4.1 still holds for pairs of frequency components that have the same
frequency dierence.
6.3.3 The Fourier spectrum of an image scanned at arbitrary resolutions
For an arbitrary target resolution, the output image from the ScanJet IIc consists of
pixels that are uniformly spaced in the vertical direction. In each row of the output image,
however, the spacing between adjacent pixels (in the physical dimension) is generally non-
uniform but is always recurrent. Therefore each row can still be considered to be the union
of several sequences of uniformly spaced pixels. Thus, the Fourier spectrum G
vs
(u) of the
output image is still the spectrum G(u) of a low-pass ltered version of the input picture
replicated in both the vertical and the horizontal directions in the frequency plane. In
the vertical direction the distance between two adjacent replicas is the nominal sampling
frequency. In the horizontal direction the distance between two adjacent replicas is generally
shorter than the nominal sampling frequency, and each replica is modulated by a non-
constant envelope.
Recall that each output row of pixels is selected from an 800 dpi interpolated row
of pixels according to a binary fraction 0.b
1
b
2
b
3
b
4
b
5
b
6
b
7
b
8
that is the ratio of the target
resolution to 800 dpi. If b
i
is the least signicant non-zero bit in the fraction, 1 i 8,
the ratio can also be written as
p
r
where r = 2
i
. It can be shown that p is odd and equals
the number of sequences of uniformly spaced pixels, the union of which is the output row.
For example, if the target resolution is 150 dpi, we have the ratio
150
800
= 0.00110000
(2)
=
3
16
.
We already see that an output row of pixels in this case is the union of three sequences of
uniformly spaced pixels. In the spectrum of the output image, replicas of the spectrum of
89
CHAPTER 6. CASE STUDY
the low-pass ltered input picture are spaced at
1
p
the nominal sampling frequency apart
in the horizontal direction.
Let q
0
, q
1
, q
2
, . . . , q
p1
be the sampling pattern of the output row of pixels. It can be
shown that the envelope env
vs
k
(u) that modulates the kth order replica in the horizontal
direction is, similar to the 150 dpi case in Section 6.3.3,
env
vs
k
(u) =
p1

i=0
e
j2
_
q
i
_
u
k
p
_
iu

. (6.4)
It is easy to verify that env
vs
k
(
k
p
) = 0 when k is not a multiple of p, and thus the frequency
component at the center of the corresponding replica is reduced to zero in the spectrum of
the output image.
The strength and location of the replicas indicate the severity of aliasing. The strength
of the kth order replica is determined in part by its modulating envelope env
vs
k
(u). For
a target resolution that results in a sampling pattern of period p, we can plot [env
vs
k
(u)[
for k = 0, 1, 2, . . . , p 1 as we did in Figure 6.7. However, when p is large, it is not easy
to plot and compare the p curves. As an extreme example, if the user requests a scan
resolution of 147 dpi, the ratio of this resolution to 800 dpi is approximated in eight bits by
100+25+12.5+6.25+3.15
800
=
146.875
800
= 0.00101111
(2)
=
47
256
, resulting in p = 47.
Since we are only interested in the magnitude of each [env
vs
k
(u)[ near u = 0, and
[env
vs
k
(0)[ gives an indication of this magnitude, we now make one plot for a target res-
olution that consists of [env
vs
k
(0)[ for k = 0, 1, 2, . . .. For each k, we draw a vertical bar
of height [env
vs
k
(0)[ at u =
k
p
, the center of the kth order replica. Thus, the location of
a bar indicates the center of a replica, and the height of the bar indicates the potential
inuence of the replica on the neighborhood of u = 0. Figure 6.12 shows six plots for target
resolutions of 146.875 dpi, 150 dpi, 193.75 dpi, 200 dpi, 250 dpi, and 300 dpi, respectively.
The values of [env
vs
k
(0)[s are normalized so that [env
vs
0
(0)[ = 1. The resolution 146.875 is
one that can be represented by a binary fraction as shown earlier. If a user requests in a
scan resolution of 146 dpi or 147 dpi, the value 146.875 is used by the scanner.
There is another reason that we choose to plot [env
vs
k
(0)[. A sampling process on a
90
CHAPTER 6. CASE STUDY
0 50 100 150 200
0
0.2
0.4
0.6
0.8
1
146.875dpi
0 50 100 150 200
0
0.2
0.4
0.6
0.8
1
150dpi
(a) (b)
0 50 100 150 200
0
0.2
0.4
0.6
0.8
1
193.75dpi
0 50 100 150 200
0
0.2
0.4
0.6
0.8
1
200dpi
(c) (d)
0 50 100 150 200 250 300
0
0.2
0.4
0.6
0.8
1
250dpi
0 50 100 150 200 250 300
0
0.2
0.4
0.6
0.8
1
300dpi
(e) (f)
Figure 6.12: The distribution and strength of replicas created by IBRM sampling
91
CHAPTER 6. CASE STUDY
sequence of discrete data points can be modeled by a multiplication of the discrete data
with a discrete sampling impulse train consisting of 0s and 1s. For example, for the 150 dpi
target resolution, an output row of pixels is obtained from an 800 dpi interpolated row by
multiplying the latter with the impulse train
1, 0, 0, 0, 1, 0, 0, 0, 0, 0, 1, 0, 0, 0, 0, 0, . . .,
in which the positions of 1s have a repeated pattern of 0, 4, 10, with a period of 16. Notice
that
env
vs
k
(0) =
p1

i=0
e
j2q
i
k
p
(6.5)
looks similar in form to a discrete Fourier transform, and it is easy to show that the equation
actually is the DFT of one period of the periodic sampling impulse train corresponding to
the sampling pattern q
0
, q
1
, . . . , q
p1
. Consider the 150 dpi target resolution case as an
example again. The DFT of one period of the sampling impulse train is
e
j2k
0
16
+e
j2k
4
16
+e
j2k
10
16
= env
vs
k
(0).
Each plot in Figure 6.12 is then also the DFT spectrum (in magnitude) of the corresponding
sampling impulse train. In fact, all six plots in the gure are produced by computing the
DFT of the corresponding sampling impulse train.
Several observations can be made from the plots in Figure 6.12.
1) The height of the kth vertical bar counted from the left in a plot indicates the
magnication by the envelope env
vs
k
(u) to the kth order replica in G
vs
(u) near u = 0.
The kth order replica itself is a shifted transform G(u) of g(x), the low-pass ltered version
of the input picture. If G(u) is high in magnitude near u =
k
p
and the kth bar is high too,
the kth replica contributes strong aliasing frequency components to the region near u = 0 in
G
vs
(u). Since g(x) is a low-pass ltered version of the input picture, G(u) tends to become
small in magnitude when [u[ is large. Thus a high order replica does not introduce strong
aliasing even if the envelope so permits.
92
CHAPTER 6. CASE STUDY
2)The 150 dpi plot clearly shows the existence of two replicas between the zero frequency
and the nominal sampling frequency.
3)The 250 dpi plot shows the existence of strong replicas centered at 100 dpi and 150
dpi, while the 200 dpi plot shows that the lowest frequency replica is centered at 200 dpi.
An image scanned at 250 dpi may have more severe aliasing than one scanned at 200 dpi.
Figures 6.13 and 6.14 respectively show a portion of the 85 dpi house halftone scanned at
200 dpi and at 250 dpi. The 200 dpi scan does not exhibit appreciable moire while the 250
dpi scan does exhibit some. Note that at 200 dpi the pixels in an output row are uniformly
spaced.
4)The binary fraction 0.b
1
b
2
b
3
b
4
b
5
b
6
b
7
b
8
that represents the target resolution determines
what pixels from the 800 dpi intermediate row are selected to form the output row. If one of
the two or three least signicant bits is 1, the sampling pattern for the pixels in the output
row will have a long period, resulting in many closely spaced replicas, though most of the
replicas are weak in energy. The 146.875 dpi plot and the 193.75 dpi plot are two examples.
5)The plot of the 146.875 dpi resembles that of the 150 dpi in energy distribution among
the replicas. So does the 193.75 dpi plot to the 200 dpi plot. In general, energy distribution
is continuous in target resolution.
Finally, we mention the concept of a symmetricizable spacing pattern. We have used
the term spacing pattern before. We now formally dene it. Corresponding to a sam-
pling pattern q
0
, q
1
, . . . , q
p1
that is measured in the unit of the nominal sampling pe-
riod, we dene a spacing pattern t
0
, t
1
, . . . , t
p1
to be a list of numbers such that t
i
=
(q
(i+1)modp
q
i
) mod p, for i = 0, 1, . . . , p 1. If t
i
= t
pi1
for i = 0, 1, . . . , p 1, the
spacing pattern is said to be symmetric. For 0 l p 1, the l-wrap shifted pattern
of the above spacing pattern is t
l
, t
l+1
, . . . , t
p1
, t
0
, t
1
, . . . , t
l1
. If for some l the l-wrap
shifted pattern of a spacing pattern is symmetric, the spacing pattern is said to be sym-
metricizable. For example, the sampling pattern 0, 0.75, 1.875 corresponds to a spacing
pattern of 0.75, 1.125, 1.125, which is not symmetric but is symmetricizable because its
1-wrap shifted pattern is 1.125, 0.75, 1.125. It can be shown that if a sampling pattern
93
CHAPTER 6. CASE STUDY
Figure 6.13: A portion of the 85 dpi House test halftone scanned at 200 dpi by a ScanJet
IIc
94
CHAPTER 6. CASE STUDY
Figure 6.14: A portion of the 85 dpi house test halftone scanned at 250 dpi by a ScanJet
IIc
95
CHAPTER 6. CASE STUDY
q
0
, q
1
, . . . , q
p1
corresponds to a spacing pattern that is symmetricizable, then
env
vs
k
(u) =
2

i=0
e
j2[q
i(u
k
3
)iu]
has a phase linear in u for any k and the phase dierence between env
vs
k
(u) and env
vs
k+1
(u)
is a constant. If the sampling pattern is symmetric, env
vs
k
(u) is real. It can also be shown
that the IBRM algorithm always produces pixels whose spacing pattern is symmetricizable.
Thus, the same phase dierence property of Theorem 5.4.1 still holds for a halftone image
scanned at any resolution by the ScanJet IIc. Unfortunately, this property does not appear
to be very useful in devising a moire pattern reduction method. The next chapter will
make use of the symmetricizable spacing pattern concept to estimate the oset between the
halftone lattice and the sampling lattice.
6.4 Discussion
The spectrum of a scanned image consists of replicas of the spectrum of the input
picture. Dierent orders of replicas may overlap if the input picture has a wide spectrum
and the low-pass ltering of the input picture before sampling is not sucient. A halftone
picture has a wide spectrum, and its energy is concentrated at harmonics of the halftone
frequency. The eect of overlap of dierent replicas in the spectrum of a scanned halftone
may be especially noticeable because strong frequency components from dierent replicas
may beat with one another to form a periodic pattern, the moire pattern.
The moire pattern problem is then immediately related to the classical topic of how
to properly low-pass lter an input signal before sampling it. However, it is expensive
to perform the correct low-pass ltering in hardware for each of the supported sampling
resolutions over a wide range. If the ltering is insucient, aliasing occurs; if the ltering is
overdone, too much signal information is lost. The IBRM subsampling scheme used by the
ScanJet IIc is easy to implement and runs eciently, but it may produce non-uniform pixels
in the horizontal direction, creating replicas that are more densely spaced than uniform
sampling and making proper low-pass ltering more dicult.
96
CHAPTER 6. CASE STUDY
For a given nominal sampling frequency, uniform sampling produces the least aliased
output because it minimizes the overlap among replicas. Even for non-uniform sampling,
the more uniform the sampled pixels, the lower energy of the replicas that lie in between
the zero frequency and the nominal sampling frequency, as indicated in Figure 6.12.
Several measures may lead to more uniformly spaced pixels. Instead of 2x interpolation
of the optically scanned data, a higher interpolated resolution makes it possible to produce
uniformly spaced output pixels at more resolutions. If the interpolated resolution has small
prime factors such as 2, 3 and 5, uniform output pixels can be obtained at many common
resolutions. For example, a 1200 dpi interpolated resolution allows for uniform output at
resolutions of 60, 75, 80, 100, 120, 150, 200, 240, 300, and 400 dpi in the resolution range of
our interest (though the scheme of BRM followed by alternating selection does not produce
uniform pixels at all of the above resolutions).
Instead of low-pass ltering the optically scanned pixels, ltering the interpolated pixels
makes it possible to choose the amount of ltering at a greater precision. Averaging two pixel
and four optically scanned pixels is approximately equivalent to averaging four and eight
interpolated pixels. One may also consider averaging two interpolated pixels. Performing
low-pass ltering directly on the 2x interpolated pixels increases the amount of computation.
However, measures can be taken to reduce the computation ([CR83, Chapters 3 and 4]).
Instead of using twice the target resolution in the BRM and then retaining every other
pixel produced by the BRM, sending a larger multiple, say k > 2 times, of the target
resolution to the BRM and then retaining every kth pixel produced by the BRM ([DB93])
gives more uniform output pixels. For example, given 400 dpi optical resolution and 2x
interpolation, if the target resolution is less than 200 dpi, one can send four times the
target resolution to the BRM and retain every fourth pixel produced by BRM.
The nearest neighbor selection method may be used to subsample the interpolated rowof
pixels; it guarantees the most uniform output pixels possible. For a given target resolution,
uniform sampling of the interpolated row may not be possible because a sampling point
may fall between two pixels. The nearest neighbor selection always nds the pixel that
97
CHAPTER 6. CASE STUDY
is nearest to a nominal uniform sampling point. There exist ecient implementations of
nearest neighbor selection that only require integer additions and simple condition control.
The need to perform subsampling comes from the fact that the optically sampled image
has a xed resolution because the optical system of the ScanJet IIc uses a xed conguration
for the object (the input picture), the lens, and the image (the CCD array). The distance
between the object plane and the lens and the distance between the lens and the image
plane is so calibrated that the input plane is in focus on the CCD array. It is also possible
to design a scanner that mechanically adjusts the above two distances to focus the input
picture onto a portion of the CCD array, thus achieving dierent resolution. However, this
second approach is more dicult to implement. Even though there exist scanners designed
that way ([Flo92]), most commercial scanners adopt the xed optical system design and
achieve variable resolution digitally.
There are other algorithms for subsampling an optically scanned image. The Ri-
coh/Improvison IS 510/520 scanners achieve lower resolution by dropping every kth pixel
from the optically scanned row of pixels. The pixels in the output rowmay be non-uniformly
spaced too. This scheme can be modeled and analyzed similarly as the IBRM scheme.
Another scanner provides only a few output resolutions albeit the pixels in the output
image are always uniform ([BBM
+
92]).
98
Chapter 7
Estimating the halftone lattice from a sampled
halftone image
7.1 Introduction
To restore systematically a sampled halftone that has a moire pattern, it is necessary to
know what frequency components contribute to the moire, which in turn requires knowledge
of the sampling lattice and the halftone lattice. Since it is reasonable to assume that we
have perfect knowledge of the sampling lattice, we need to estimate the halftone lattice. The
more accurate the estimate of the halftone lattice, the higher the potential quality of the
restoration. Restoration of an image with an aliasing moire pattern requires more accurate
knowledge of the halftone lattice than does restoration of an image with a non-aliasing
moire pattern.
The halftone lattice consists of lines connecting the centers of halftone cells. Four param-
eters completely dene the lattice: the period and phase in the two directions perpendicular
to the lattice lines. We use the image coordinate system that coincides with the sampling
lattice to measure the halftone lattice. The phases of the halftone lattice are the oset of
the lattice with respect to the sampling lattice. Figure 7.1(a) illustrates a halftone lattice
relative to a sampling lattice, as well as the coordinate system we use. The solid lattice is
the halftone lattice and the intersection of two solid lines is the center of a halftone cell.
The dotted lattice is the sampling lattice. P and Q are the halftone period vectors, and d
p
and d
q
are the phases of the halftone lattice in the P and Q directions, respectively. In this
chapter we estimate P, Q, d
p
, and d
q
. We do so in the frequency domain.
In the frequency domain, the halftone period vectors P and Qcorrespond to the halftone
99
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
P
x
y
d
d
q
p
Q
-p
v
u
p
q
-q
(a) (b)
Figure 7.1: The halftone lattice relative to the sampling lattice (a) and the halftone fre-
quencies in the frequency domain (b)
frequency vectors p and q. The two pairs of vectors have a reciprocal relationship: p =
P
[P[
2
,
q =
Q
[Q[
2
. Figure 7.1(b) illustrates the halftone frequencies p and q, as well as their conju-
gate frequencies p and q.
In this chapter, h(x) represents a halftone picture with Fourier transform H(u), and
h
s
(x) represents an image sampled from h(x) with transform H
s
(u). Also, recall from
Chapter 5 that we use the four-tuple (0, 0, m, n) to refer to the order of mp + nq, the
(m, n)th harmonic of the halftone frequencies in the (0, 0)th replica in H
s
(u).
From Eq. 5.8, which is from Kermisch and Roetling [KR75], it can be shown that the
heights of the peaks in H(u) at u = p or u = q are second only to the height of the peak at
u = 0 under very general conditions. (Exceptions include contrived halftones that represent
only black and white image content, for which a halftone has no lattice structure at all,
100
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
and halftones that represent only a sinusoid whose amplitude is nearly 1). Therefore, the
heights of the peaks in H
s
(u) at u = p or u = q are second only to the height of the peak
at u = 0. We may perform a simple search on H
s
(u) to locate the peaks at u = p or u = q.
However, since we can compute only the discrete Fourier transform (DFT), which delivers
values only at discrete frequencies, a simple search of the DFT values can lead only to the
closest frequency representable by the DFT to p or q. The accuracy of the simple search is
usually not sucient for restoring images that have aliasing moire. Fortunately, we can take
advantage of the leakage phenomenon of the DFT to interpolate a more accurate estimate
of the halftone frequency.
H
s
(p) is a complex number. The ratio
Im[H
s
(p)]
Re[H
s
(p)]
reects the phase d
p
. We estimate d
p
by computing the ratio after obtaining an estimate of p.
p, q, d
p
, and d
q
can only be estimated from H
s
(u), if the sampling rate is suciently
high that the corresponding Nyquist band encloses p and q.
In addition to the frequency domain approach, there is also an image domain approach
for estimating the halftone lattice. The halftone frequency is estimated by computing
the autocorrelation of the sampled halftone and measuring the distance between adjacent
autocorrelation peaks. The halftone lattice phases are estimated by evaluating the cross-
correlation of the sampled halftone with a periodic function (e.g. a sinusoid) whose initial
phase is known and whose frequencies equal those of the halftone lattice. The location of
a cross-correlation peak indicates the phases of the halftone lattice. However, this image
domain approach is not as ecient as the frequency domain approach.
7.2 The leakage phenomenon of the Fourier transform
A real image has a nite size and can be modeled as the truncation of an innite-
size image. So far we have considered only the Fourier spectra of images of innite size
for simplicity. The Fourier spectrum of the truncation of a function exhibits the leakage
phenomenon ([Bri74] and [Bel84]). The energy associated with one frequency is distributed,
101
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
or leaked, to a spectrum of frequencies. In this section we discuss leakage in 1D and mention
its counterpart in 2D.
A function can be synthesized as the sum of (possibly uncountably) many sinusoids.
Each sinusoid corresponds to a pair of complex conjugate frequency components in the
spectrum of the function. The truncation of the function means the truncation of all the
contributing sinusoids. We nowconsider the spectrum of a truncated sinusoid. In particular,
we consider a length L truncation of cos 2px,
f(x) = rect
_
x
L
_
cos 2px,
where rect(x) is the rectangle function:
rect(x) =
_

_
0 [x[ >
1
2
1 [x[
1
2
,
and L and p are constants. cos 2px is said to be windowed by a rectangular window
function. The Fourier transform F(u) of f(x) is
F(u) = L
_
sin L(u + p)
L(u + p)
+
sinL(u p)
L(u p)
_
. (7.1)
F(u) is, except for a scaling factor, the sum of two translated sinc functions instead of the
sum of two translated functions as in the spectrum of cos(2px) itself. The dashed curve in
Figure 7.2 shows the magnitude of F(u) for p = 3.6 and L = 1. The two dot-dashed vertical
bars in the gure are [F(p)[ and [F(p)[ respectively. The energy originally concentrated
at frequencies p and p is leaked to other frequencies due to truncation.
We nowsample f(x) with a frequency s > 2p to obtain f
s
(n), n =
N
2
,
N
2
+1, . . . ,
N
2
1,
where N = sL and f
s
(n) = f(
n
s
). The DFT F
s
(m), m =
N
2
,
N
2
+ 1, . . . ,
N
2
1, of f
s
(n)
is
F
s
(m) = L
_
_
sinL(
m
L
+p)
N sin
L(
m
L
+p)
N
+
sinL(
m
L
p)
N sin
L(
m
L
p)
N
_
_
,
which can be considered to be sampled from
F
d
(u) = L
_
sin L(u + p)
N sin
L(u+p)
N
+
sinL(u p)
N sin
L(up)
N
_
.
102
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
[F(u)[, [F
s
(m)[
-5 -4 -3 -2 -1 1 2 3 4 5
u,m
-p p
Figure 7.2: Leakage of DFT
The function
1
N
sin u
sin
u
N
is sometimes called the digital sinc function, and it indeed approxi-
mates the conventional sinc very well when N is large. For our purpose, N is on the order
of at least 256 because a 256 256 pixel image is already small. For N=256, the relative
error between the digital sinc and the conventional sinc is less than 3 10
5
for [u[ 1.
Except for a scaling factor, F
d
(u) is the sum of two translated digital sinc functions. Thus,
F
s
(m) is essentially a sampled version of F(u).
F
s
(m) assumes values at frequencies that are multiples of
1
L
: F
s
(m) = F
d
(
m
L
)
F(
m
L
), m =
N
2
,
N
2
+ 1, . . . ,
N
2
1. Note that F(u) = F
d
(u) = 0 for u =
n
L
p, n =
1, 2, . . .. If
k
L
= p, for some integer k, that is, p is a multiple of
1
L
, F
s
(m) is zero except
at m = k = p L and m = k = p L. The leakage in the continuous transform is, by
coincidence, not reected in the DFT. If p is not a multiple of
1
L
, F
s
(m) assumes values
that approximately satisfy Eq. 7.1 at u =
m
L
, m =
N
2
, . . . ,
N
2
, as illustrated in Figure 7.2 for
N = 256. In the gure the tick marks represent frequency sampling points of the DFT, and
the solid vertical bars represent the magnitude of [F
s
(m)[. The DFT is scaled horizontally
in such a way that a value m coincides with the frequency it represents,
m
L
. The two largest
magnitude DFT values in [s/2, 0) occur at pL| and pL| respectively, the two closest
103
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
frequencies of the DFT to p. Similarly, the two largest magnitude DFT values in [0, s/2)
occur at pL| and pL| respectively. Leakage actually helps preserve information about a
frequency component that is not at a frequency sampling point.
In practice, samples of a function can only be collected in a nite interval. In the DFT of
the samples of a general function, all components of the function whose frequencies are not
at a frequency sampling point of the DFT are leaked according to the same model above.
Leakage is generally considered to be an adverse property of the Fourier transform and
the discrete Fourier transform because it mixes energy from dierent frequencies. At each
frequency, the transform of a truncated signal is inuenced by many frequency components
of the original signal, thus making the measurement of energy distribution (in frequency) of
the original signal inaccurate. Many tapering window functions are proposed in the litera-
ture to suppress leakage so that a strong frequency component does not overwhelm weak
components whose frequencies are distinctly dierent. A nite-length signal is multiplied
with a tapering window (instead of the rectangular window) before its DFT is computed.
We do not use a tapering window in estimating parameters of a halftone lattice for
three reasons: 1) A tapering window smoothes out sharp peaks in the Fourier spectrum,
making accurate frequency measurement dicult. 2) We use only a small neighborhood of
the spectrum of a sampled halftone at each rst order harmonics of the halftone frequencies.
Leakage does not signicantly aect the energy distribution in such a neighborhood because
the energy there is much stronger than the energy in neighboring frequencies. 3) Without
a tapering window, the energy distribution near a halftone frequency harmonic is easy to
model.
In 2D, let f(x, y ) be the truncation of a sinusoid cos 2(p
x
x + p
y
y ) in a square area
_

L
2
,
L
2
_

L
2
,
L
2
_
,
f(x, y ) = rect
_
x
L
_
rect
_
y
L
_
cos 2(p
x
x +p
y
y ).
The Fourier transform F(u, v) of f(x, y ) is
F(u, v) =
sin L(u p
x
)
(u p
x
)
sinL(v p
y
)
(v p
y
)
+
sin L(u +p
x
)
(u +p
x
)
sinL(u + p
y
)
(u + p
y
)
. (7.2)
104
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
-6
-4
-2
0
2
4
6
u
-6
-4
-2
0
2
4
6
v
|F(u,v)|
-6
-4
-2
0
2
4
6
u
Figure 7.3: Leakage of Fourier transform in 2D
Figure 7.3 shows the magnitude of F(u, v). The two peaks are located at (p
x
, p
y
) and
(p
x
, p
y
), respectively. Leakage is most notable in vertical and horizontal directions.
The 2D DFT of a truncated and sampled function is still essentially a sampled version
of the continuous transform of the truncated function. A component of the function whose
frequency is not at a frequency sampling point is reected in the DFT through leakage. In
this dissertation the computation of the Fourier transforms of all sampled halftone images
is performed via the DFT. Leakage of strong frequency components is evident in all plots
of such transforms, as in Figures 5.10, 5.11(b), and 6.4.
The Fourier transform of a general halftone picture is usually continuous. Both the
zero frequency component and its neighbors have much greater energy than high frequency
components. In the DFT of a sampled general halftone, the zero frequency component of the
halftone exhibits no leakage because zero frequency is always a frequency sampling point.
Leakage of the near-zero frequency components, however, is present in the DFT. Since
leakage is most signicant in vertical and horizontal directions, the leakage of near-zero
frequency components aects only frequencies near the two frequency axes appreciably.
105
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
When a halftone picture is sampled, if the sampling lattice and the halftone lattice are
at such an angle that the (0, 0, 1, 0)th and the (0, 0, 0, 1)st halftone frequency harmonics
do not appear near the u or v axis, leakage from the near-zero frequency components does
not aect the energy distribution at those harmonics; this is conrmed by experiments as
shown in Section 7.3. Since most halftone lattices are oriented at near 45 degrees with
respect to the sampling lattice, leakage from the near-zero frequency components to the
rst order halftone frequency harmonics is already minimized. Furthermore, if the halftone
lattice is square, orienting the halftone lattice at an angle slightly away from 45 degrees to
the sampling lattice reduces the leakage among the components at the rst order halftone
frequency harmonics.
7.3 Halftone lattice estimation from a uniformly sampled halftone
Recall that H(u) is the Fourier transform of a halftone picture, and p and q are the
halftone frequency vectors. From Theorem 5.6.1 we know that H(p), H(p), H(q), and
H(q) are positive if the center of a halftone cell coincides with the origin of the coordinate
system. Therefore, the pair of conjugate components H(p) and H(p) correspond to a
cosinusoid, cos 2(p x). If the halftone is shifted with respect the coordinate system
as shown in Figure 7.1, H(u) is multiplied by a complex phase factor reecting the oset.
H(p) and H(p) then correspond to cos 2(px[p[d
p
). This cosinusoid achieves maximum
value on each solid line perpendicular to P in Figure 7.1(a). The pair of components H(q)
and H(q) represent another cosinusoid, cos 2(q x [q[d
q
).
In the spectrum of a sampled real (nite-size) halftone picture, the components in a small
neighborhood of the (0, 0, 1, 0)th or (0, 0, 0, 1)st harmonic of the halftone frequencies can
be modeled by the product of two translated sinc functions.
There are ve types of modeling errors. Let u
0
be a rst order harmonic of the halftone
frequencies: u
0
= p or u
0
= q.
1. The zeroth order replica is enveloped by a low-pass lter function that is not a constant
106
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
but decreases with increasing frequency. The lter tends to suppress more the portion
of a peak that is on the opposite side to the zero frequency.
2. Since it is eventually the DFT of the sampled halftone that is computed, the DFT
spectrum near u
0
is not exactly the product of two sinc functions and is better ap-
proximated by the product of two translated digital sinc functions.
3. In the spectrum of a halftone, the frequency components in the immediate neigh-
borhood of u
0
are usually not zero. These components also leak in the DFT of the
samples of the halftone and thus aect the energy distribution near u
0
.
4. Leakage from strong components at other frequencies is not zero near u
0
. For example,
if another rst order halftone frequency harmonic has the same u or v coordinate as
u
0
, the components at the two harmonics leak to each other. Figures 5.10, 5.11(b),
and 6.4 all show such leakage. Figure 7.2 also shows the same type of leakage in 1D.
The two highest peaks of [F(u)[ in the gure are not located exactly at p and p.
5. The halftone lattice of a real halftone picture may not be consistent. The halftone
frequency may vary across the halftone picture ([FJ94]).
The rst three types of errors are small in our application. The fourth types of errors can
be made small if the angle between the halftone lattice and the sampling lattice is properly
chosen. The fth type of error is small if the input halftone is of good quality and is free
of geometric distortion.
7.3.1 Estimating the halftone frequencies
In this subsection we rst consider estimating the frequency of a cosinusoid in 1D and
then extend the method to 2D. Finally, we apply the method to a sampled halftone.
Let a cosinusoid cos 2p(x d
0
) be truncated in [0, L) and then sampled at
n
L
, n =
0, 1, . . . , N 1, such that
N
L
> 2p. The reason we place the truncation in [0, L) is that it is
customary to express a nite-size image in an x 0 interval. (Figure 7.1(a) also indicates
107
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
that the halftone is in an x 0 and y 0 region.) We are to estimate p from the N samples
of the cosinusoid. It can be shown that the Fourier transform of the truncated cosinusoid is
F(u) =
L
2
_
e
j2pd
0
e
jL(u+p)
sinc[L(u +p)] +e
j2pd
0
e
jL(up)
sinc[L(u p)]
_
. (7.3)
If Lp is suciently large, the magnitude of F(u) is dominated by
L
2
sinc[L(u + p)] near
u = p because the leakage from the other sinc is small. Likewise, the magnitude of F(u)
is dominated by
L
2
sinc[L(u p)] near u = p.
Let A
m
, m =
N
2
,
N
2
+ 1, . . . ,
N
2
1 be the DFT of the cosinusoid samples. Then
A
m
F(
m
L
). Assume that
m
0
L
p <
m
0
+1
L
, for some m
0
. Such an m
0
exists because of
our sampling rate assumption. It is easy to show that A
m
0
and A
m
0
+1
are the two largest
magnitude A
m
for m near m
0
. Figure 7.2 can be used as a reference, where m
0
= 3.
For m near m
0
, we have
[A
m
[

F
_
m
L
_

L
2

sinc
_
L
_
m
L
p
__

sin(mLp)
(mLp)/L

.
Let m = Lp m
0
, that is, p =
m
0
L
+
m
L
. It is easy to show that
[ sin(mLp)[ sinm
for all values of m. We then have
[A
m
0
[m [A
m
0
+1
[(1 m),
which leads to
m
[A
m
0
+1
[
[A
m
0
[ +[A
m
0
+1
[
. (7.4)
The approximation happens to have the same form as a linear interpolation. The estimate
is correct even if p is a multiple of
1
L
, in which case p =
m
0
L
and A
m
0
+1
= 0.
Now we have an algorithm to estimate p:
1. Compute the DFT of the samples of the cosinusoid;
2. search the DFT array in the positive frequency range for the element A
m
0
that has
the largest magnitude;
108
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
3. select the element A
m
1
adjacent to A
m
0
that has the larger magnitude (m
1
= m
0
1);
4. compute m =
|Am
1
|
|Am
0
|+|Am
1
|
;
5. p =
m
0
+(m
1
m
0
)m
L
is an estimate of p.
We can search the negative frequency segment of the DFT array and estimate p instead.
However, it does not improve the accuracy of the estimate of p if we estimate both p and
p and average the two estimates because DFT is hermitian. In fact, the DFT is most likely
to be computed through FFT algorithms which simply reverse the sign of the DFT values
in the positive frequency range to obtain the negative frequency components.
In 2D, the Fourier transform of the [0, L) [0, L) truncation of a cosinusoid cos 2(p
x [p[d
p
) = cos 2(p
x
x + p
y
y [p[d
0
) is
F(u, v) =
L
2
2
_
e
j2|p|d
0
e
jL(u+px+v+py)
sinc[L(u +p
x
)]sinc[L(v + p
y
)]+
e
j2|p|d
0
e
jL(upx+vpy)
sinc[L(u p
x
)]sinc[L(v p
y
)]
_
. (7.5)
Assuming that L[p[ is suciently large, the magnitude of F(u, v) is dominated by sinc[L(u+
p
x
)]sinc[L(v+p
y
)] near (u, v) = (p
x
, p
y
) and is dominated by sinc[L(up
x
)]sinc[L(vp
y
)]
near (u, v) = (p
x
, p
y
). Note that u and v are completely separable in each product of two
sinc functions.
Let the truncated cosinusoid be sampled at (
k
L
,
l
L
), k, l = 0, 1, . . . , N1, and let A
nm
, m, n =

N
2
, . . . ,
N
2
1 be the 2D DFT of the samples. Then A
nm
F(
m
L
,
n
L
). Note that we use m
as the column subscript which corresponds to the u variable. We perform the 1D algorithm
on the A
mn
s in the horizontal and the vertical directions separately to estimate p
x
and p
y
.
Without loss of generality, we dene the right half of the frequency plane, u 0 and v 0
or u > 0 and v < 0, as the positive frequency region. We have the following algorithm:
Algorithm 7.1
1. Compute the 2D DFT of the samples of the cosinusoid;
2. search the 2D DFT array in the positive frequency region for the element A
n
0
m
0
that
has the largest magnitude;
109
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
3. select the element A
n
0
m
1
horizontally adjacent to A
n
0
m
0
that has the larger magnitude;
select the element A
n
1
m
0
vertically adjacent to A
n
0
m
0
that has the larger magnitude;
4. compute m =
|An
0
m
1
|
|Am
0
n
0
|+|An
0
m
1
|
and n =
|An
1
m
0
|
|Am
0
n
0
|+|An
1
m
0
|
;
5. p
x
=
m
0
+(m
1
m
0
)m
L
is an estimate of p
x
and p
y
=
n
0
+(n
1
n
0
)n
L
is an estimate of p
y
.
Since the spectrum of a nite-size halftone can be approximated by the product of two
sinc functions in the neighborhood of u
0
= p or u
0
= q, we can apply Algorithm 7.1
to a sampled halftone to estimate p and q. We restrict the search step to frequencies away
from the zero frequency so that strong components near the zero frequency are not included.
We also nd the two largest local maximum components and perform interpolation at both
frequencies.
Forchhammar and Jensen [FJ94] briey discuss the above interpolation method. The
authors also suggest that interpolation be applied to components near higher order har-
monics of the halftone frequencies if such harmonics are within the DFT spectral band of
the sampled halftone. As discussed in Section 5.8, the continuous Fourier transform of a
sampled halftone may not have a peak right at a higher than 1 order halftone frequency
harmonic. It can be easily veried that the DFT of a sampled halftone image may not
have a local maximum magnitude component at that halftone frequency harmonic either.
Therefore, the interpolation method does not always work at a higher order harmonic of
the halftone frequencies.
7.3.2 Estimating the phases of the halftone lattice
In this section we rst consider estimating the phase of a sampled cosinusoid in 1D and
then extend the method to a sampled cosinusoid in 2D. Lastly, we apply the method to a
sampled halftone.
Rewrite Eq.7.3 as
F(u) = Neg(u) + Pos(u),
110
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
where
Neg(u) =
L
2
e
j[2pd
0
L(u+p)]
sinc[L(u + p)]
and
Pos(u) =
L
2
e
j[2pd
0
+L(up)]
sinc[L(u p)].
Pos(u) dominates in F(u) near u = p. We can write Pos(u) as
Pos(u) =
L
2
sinc[L(u p)]cos[2pd
0
+ L(u p)] j sin[2pd
0
+ L(u p)]
= Re[Pos(u)] + jIm[Pos(u)].
Then we have
d
0
=
arctanIm[Pos(u)]signsinc[L(u p)], Re[Pos(u)]signsinc[L(u p)] L(u p)
2p
.
The range of arctan is (, ). If [u p[ <
1
L
, sinc[L(u p)] is positive, and the above
equation is simpler in form. In particular, let m
0
be such that [Pos(
m
L
)[ achieves maximum
among all integer values of m. Then

m
0
L
p

<
1
L
. We have
d
0
=
arctan
_
Im
_
Pos
_
m
0
L
_
, Re
_
Pos
_
m
0
L
__
L
_
m
0
L
p
_
2p
.
The DFT A
m
of the cosinusoid samples also achieves maximum magnitude at m
0
. Since
Pos
_
m
0
L
_
F
_
m
0
L
_
A
m
0
, we have

d
0
=
arctan[Im(A
m
0
), Re(A
m
0
)] L
_
m
0
L
p
_
2 p
d
0
. (7.6)
In 2D, we rewrite Eq.7.5 as
F(u, v) = Neg(u, v) + Pos(u, v),
where
Neg(u, v) =
L
2
2
e
j[2|p|dpL(u+px+v+py)]
sinc[L(u +p
x
)]sinc[L(v + p
y
)],
and
Pos(u, v) =
L
2
2
e
j[2|p|dp+L(upx+vpy)]
sinc[L(u p
x
)]sinc[L(v p
y
)].
111
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
Pos(u, v) dominates in F(u, v) near (u, v) = (p
x
, p
y
). We can write Pos(u, v) as
Pos(u, v) =
L
2
2
sinc[L(u p
x
)]sinc[L(v p
y
)]
cos[2[p[d
p
+L(u p
x
+v p
y
)] j sin[2[p[d
p
+ L(u p
x
+ v p
y
)]
= Re[Pos(u, v)] +jIm[Pos(u, v)].
For [(u, v) (p
x
, p
y
)[ <
1
L
,
d
p
=
arctanIm[Pos(u, v)], Re[Pos(u, v)] L(u p
x
+ v p
y
)
2[p[
.
Let (m
0
, n
0
) be such that the DFT A
mn
of the cosinusoid samples achieves maximum
magnitude. We have

d
p
=
arctan[Im(A
n
0
m
0
), Re(A
n
0
m
0
)] L
_
m
0
L
p
x
+
n
0
L
p
y
_
2[ p[
d
p
. (7.7)
For the same reason that the cosinusoid frequency estimation algorithm can be used
to estimate halftone frequencies from a sampled halftone, Eq. 7.7 can be used to estimate
the phases d
p
and d
q
of a halftone lattice from a sampled halftone. We estimate d
p
and d
q
separately by using Eq. 7.7 twice.
7.3.3 Measuring the accuracy of halftone lattice estimation
The best wayto test the performance of Algorithm7.1 is to apply it to a sampled halftone
whose frequency is accurately known. However, the author of this dissertation does not have
the means to obtain a halftone with an accurately known frequency, nor does the author
have the means to measure the frequency of an existing halftone to an accuracy that is
comparable to the level of accuracy that can be achieved by the algorithm. Furthermore,
since it is dicult to physically measure the oset of a halftone lattice relative to a sampling
lattice, it is also dicult to test the performance of Eq. 7.7 using a real halftone and a real
sampler (e.g. a scanner).
Synthesized test images are then used as substitutes to truly sampled halftones. The
test images are uniform samples of the following two functions:
h(x) = rand(x) + 0.25(1 )[2 + cos 2(p x +[p[d
p
) + cos 2(q x +[q[d
q
)] (7.8)
112
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
h(x) = g(x) +0.25(1 )[2 + cos 2(p x +[p[d
p
) +cos 2(q x +[q[d
q
)], (7.9)
where rand(x) is a unit magnitude, uniformly distributed random noise, and g(x) is a digital
image. 0 < < 1 is a coecient that determines the relative strength of the sinusoidal
screen structure to that of the image content in the test image. The two numeric
constants 0.25 and 2 maintain h(x) in the range [0, 1]. The two sinusoids are chosen to be
perpendicular to each other and have the same absolute frequency: p q and [p[ = [q[.
We use the following metrics to measure, respectively, the error of the frequency estimate
p and the error of the phase estimate

d
p
for the sinusoid cos 2(p x+[p[d
p
) in a test image:
Err( p) = N[ p p[,
Err(

d
p
) = [p[[

d
p
d
p
[,
where N is the image width in pixels. The quantity N[p[ is the number of period [P[ =
1
[p[
of the sinusoid that can be embedded in the width of the image. N[ p[ is the number of
estimated period [

P[ =
1
[ p[
that can be embedded in the width of the image. Err( p) is
the number of periods in excess that is accumulated across the width of the image due
to the error in p. An Err( p) of 0.5 means if we were to construct a sinusoid across the
width of the image using the true phase d
p
and the estimated frequency p, the constructed
sinusoid would have exactly the opposite phase at the other end of the image to one that
is constructed using the true period. The phase d
p
can be expressed as a percentage of the
length of the period [P[. [p[d
p
is this percentage. Err(

d
p
) is the error of

d
p
in terms of
the percentage of the true period [P[. The worst value of Err(

d
p
) is 0.5, which means the
estimated phase

d
p
is exactly opposite to the true phase d
p
. An Err(

d
p
) of 0.9 is equivalent
to an Err(

d
p
) of 0.1.
Six sets of test images were created. The rst three sets use Eq. 7.8 with = 0.3, 0.6,
and 0.8, respectively, and the other three sets use Eq. 7.9 with the same three values,
respectively. There are four images in each set; each image has a dierent value of p but
the same values of d
p
(=
0.29
|p|
) and d
q
(=
0.69
|q|
). Figure 7.4 shows the four values of p (solid
line vectors) used in the test images. Also shown are the frequencies (dotted line vectors)
113
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
u (0.3,0.0)
(0.3,0.3)
(0.3,0.38)
(0.45,0.45)
v
Figure 7.4: Frequencies used in the test images for frequency and phase estimation
of q. All images have size 512 512 pixels (N=512). g(x) is the digital image that is used
to produce the test halftone House in Figure 5.8.
We use the following two metrics to measure the overall frequency estimation perfor-
mance and phase estimation performance for a test image:
Err( p, q) = max[Err( p), Err( q)]
Err(

d
p
,

d
q
) = max[Err(

d
p
), Err(

d
q
)].
Table 7.1 lists Err( p, p) and Err(

d
p
,

d
q
) for the rst three sets of images. The image
Table 7.1: Errors of the frequency and phase estimates on test images that are sums of
random noise and two sinusoids
image = 0.3 = 0.6 = 0.8
p = (p
x
, p
y
) Err( p, q) Err(

d
p
,

d
q
) Err( p, q) Err(

d
p
,

d
q
) Err( p, q) Err(

d
p
,

d
q
)
(0.30,0.00) 5 10
3
2 10
3
2 10
2
6 10
3
5 10
2
2 10
2
(0.30,0.30) 3 10
3
1 10
3
5 10
3
3 10
3
1 10
2
8 10
3
(0.30,0.38) 1 10
3
5 10
4
5 10
3
2 10
3
1 10
2
5 10
3
(0.45,0.45) 5 10
3
5 10
3
8 10
3
5 10
3
1 10
2
7 10
3
114
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
with p = (0.30, 0.00) gives most of the largest errors and the image with p = (0.30, 0.38)
gives the smallest errors. In the spectrum of the former image, leakage from the near-zero
frequencies aect the peaks at p and q; in the spectrum of the latter image, leakage has the
little eect on the same peaks. The (0.45,0.45) image gives larger errors than the (0.3,0.3)
image because in the spectrum of the former image, peaks at p and q and their aliases in
neighboring replicas are closer to one another, so leakage between them is stronger.
Overall, the errors in Table 7.1 are very small. For the simulation method in Chapter 8,
which requires an estimate of the halftone lattice, errors of less than 0.01 in both the
frequency and the phase estimates appear to be sucient while errors of larger than 0.05
result in unsuccessful restoration.
Table 7.2 lists Err( p, p) and Err(

d
p
,

d
q
) for the other three sets of test images. The
Table 7.2: Errors of the frequency and phase estimates on test images that are sums of the
house image and two sinusoids
image = 0.3 = 0.6 = 0.8
p = (p
x
, p
y
) Err( p, q) Err(

d
p
,

d
q
) Err( p, q) Err(

d
p
,

d
q
) Err( p, q) Err(

d
p
,

d
q
)
(0.30,0.00) 3 10
3
1 10
3
8 10
3
3 10
3
2 10
2
7 10
3
(0.30,0.30) 2 10
3
5 10
4
2 10
3
7 10
4
2 10
3
7 10
4
(0.30,0.38) 1 10
4
2 10
5
3 10
4
7 10
5
8 10
4
2 10
4
(0.45,0.45) 5 10
3
5 10
3
5 10
3
5 10
3
5 10
3
5 10
3
errors here have a similar distribution as in Table 7.1. However, the errors here are smaller,
and the (0.45,0.45) image gives consistent errors for all three values of . The reason for
these dierences appears to be that the image g(x, y ) does not have as strong components
at high frequencies as rand(x, y ) does.
A question of interest is what values of give rise to a test image whose sinusoidal
screen is comparable in strength to the halftone lattice in a sampled halftone image. We
compare the ratio of the energy at frequency p to the energy at zero frequency of a test
image with the ratio of the energy at a halftone frequency to the energy at zero frequency
of a sampled halftone. By this comparison, an value of 0.5 to 0.6 appears to approximate
a sampled halftone well.
115
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
7.4 Halftone lattice estimation from a non-uniformly sampled halftone
Pixels in an image scanned by the ScanJet IIc are sampled from the input picture uni-
formly in the vertical direction and non-uniformly yet recurrently in the horizontal direction.
As discussed in Chapter 6, each replica in the spectrum of such an image is modulated by an
additional envelope that varies in the horizontal direction but is a constant in the vertical
direction. In the following, we take into account the eect of the envelope that modulates
the zeroth order replica. We still begin with the 1D case. For simplicity, we use the term
non-uniform to mean non-uniform and recurrent in this section.
Let a sinusoid cos 2p(xd
0
) be truncated in [0, L) and then sampled non-uniformly. In
the neighborhood of u = p, the continuous Fourier transform of the truncated and sampled
sinusoid can be modeled by the product of a sinc function and an envelope function due
to non-uniform sampling. The DFT of the sinusoid samples, A
m
, m =
N
2
, . . . ,
N
2
1,
approximates the continuous transform. For values of m such that
m
L
is near p, we have
A
m

L
2
e
j[2pd
0
+L(up)]
sinc[L(u p)]env
vs
0
(u)

u=
m
L
, (7.10)
where env
vs
0
(u) is dened in Eq. 6.4. Dene
B
m
=
L
2
e
j[2pd
0
+L(up)]
sinc[L(u p)]

u=
m
L
.
A
m
can be written as
A
m
B
m
env
vs
0
_
m
L
_
.
The magnitude of env
vs
0
(u) does not vary dramatically over a short interval in u, so
env
vs
0
(u) does not change appreciably the relative magnitude of A
m
for values of m such
that
m
L
is very close to p. Figure 6.7 shows the magnitude of one env
vs
0
(u) (for a sampling
pattern of period 3). It turns out that if we simply apply the frequency estimation algo-
rithm in Section 7.3.1 on a non-uniformly sampled sinusoid, the estimated frequency is as
accurate as one obtained by applying the algorithm on a uniformly sampled sinusoid. In
other words, the error introduced by the varying magnitude of env
vs
0
(u) is smaller than
116
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
errors of other causes (such as leakage or noise in the input data). Nevertheless, we still
have the following algorithm that compensates the varying magnitude of env
vs
0
(u).
1. Compute the DFT of the non-uniformly sampled sinusoid;
2. search the DFT array in the positive frequency range for the element A
m
0
that has
the largest magnitude;
3. select the element A
m
1
adjacent to A
m
0
that has the larger magnitude (m
1
= m
0
1);
4. compute [B
m
0
[ =
|Am
0
|
|envvs
0
(
m
0
L
)|
and [B
m
1
[ =
|Am
1
|
|envvs
0
(
m
1
L
)|
;
5. compute m =
|Bm
1
|
|Bm
0
|+|Bm
1
|
;
6. p =
m
0
+(m
1
m
0
)m
L
is an estimate of p.
Experiments show that the above algorithm produces estimates that are virtually indistin-
guishable from those by the uniform algorithm. An important note is that the envelope
function env
vs
0
(u) is dependent of the sampling pattern. Thus, the above algorithmrequires
knowledge of the sampling model.
env
vs
0
(u) changes the complex phase of the spectrum of the sampled data. However,
we can compute the complex phase of env
vs
0
(u) and compensate it in a phase estimate.
Specically, write env
vs
0
(u) as
env
vs
0
(u) = [env
vs
0
(u)[e
j2(u)
,
and let Pos(u) be as dened in Section 7.3.2. We have
d
0
=
arctanIm[Pos(u)]signsinc[L(u p)], Re[Pos(u)]signsinc[L(u p)] L(u p) + (u)
2p
.
Let m
0
be such that A
m0
has the largest magnitude among all A
m
s for m > 0, we have

d
0
=
arctan[Im(A
m
0
), Re(A
m
0
)] L(m
0
/L p) +(
m
0
L
)
2 p
d
0
. (7.11)
Recall from Chapter 6 that if the sampling pattern is symmetric, env
vs
0
(u) is real, so
(u) is zero. In this case, Eq. 7.11 reduces to Eq. 7.6. In other words, if the sampling
117
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
pattern is symmetric, the data samples can be treated as if they were uniformly spaced
both in frequency estimation and in phase estimation. In addition, if the sampling pattern
of a sequence of data samples is symmetricizable, dropping a few samples at the beginning
of the sequence makes the sampling pattern symmetric.
In 2D the development is analogous to 1D. Here we just list the frequency estimation
algorithm and the phase estimation formula without further explanation. Only the phase
estimation formula is necessary.
Algorithm 7.2
1. Compute the 2D DFT of the non-uniform samples of the input halftone;
2. search the DFT array in the positive frequency region for the element A
n
0
m
0
that has
the largest magnitude;
3. select the element A
n
0
m
1
horizontally adjacent to A
n
0
m
0
that has the larger magnitude;
select the element A
n
1
m
0
vertically adjacent to A
n
0
m
0
that has the larger magnitude;
4. compute [B
n
0
m
0
[ =
|An
0
m
0
|
|envvs
0
(
m
0
L
)|
and [B
n
0
m
1
[ =
|An
0
m
1
|
|envvs
0
(
m
1
L
)|
;
5. compute m =
|Bn
0
m
1
|
|Bn
0
m
0
|+|Bn
0
m
1
|
and n =
|An
1
m
0
|
|An
0
m
0
|+|An
1
m
0
|
;
6. p
x
=
m
0
+(m
1
m
0
)m
L
is an estimate of p
x
and p
y
=
n
0
+(n
1
n
0
)n
L
is an estimate of p
y
.
The phase estimate is

d
p
=
arctan[Im(B
n
0
m
0
), Re(B
n
0
m
0
)] L
_
m
0
L
p
x
+
n
0
L
p
y
_
+
_
m
0
L
_
2[ p[
d
p
. (7.12)
Again, both Algorithm7.2 and Eq. 7.12 require knowledge of the sampling model to evaluate
(
m
0
L
).
Test images from Eqs.7.8 and 7.9 were created to measure the performance of Algo-
rithm 7.2 and Eq. 7.12. The images are non-uniformly sampled using the sampling pat-
tern 0, 0.75, 1.185 and its two wrap-shifted variants 0, 1.125, 1.875 and 0, 1.125, 2.25.
118
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
The spacing patterns corresponding to the three sampling patterns are 0.75, 1.125, 1.125,
1.125, 0.75, 1.125, and 1.125, 1.125, 0.75, respectively. Let a test image be T
lk
, k, l =
0, 1, . . . , N 1. For a sampling pattern q
0
, q
1
, q
2
,
T
lk
= h(x
k
, y
l
),
where h(x, y ) is one of Eqs.7.8 and 7.9, and
x
k
=
1
L
(k i +q
i
) , i = k mod 3,
y
l
=
1
L
l .
The test results show that, for 0.3, Algorithm 7.2 and Eq. 7.12 achieve comparable
accuracy as Algorithm 7.1 and Eq. 7.7. For brevity, we do not list the test results.
7.5 Summary and discussion
The halftone lattice of a sampled halftone can be estimated from the DFT components
of the sampled halftone in the 2 2 pixel neighborhoods of the (0, 0, 1, 0)th and the
(0, 0, 0, 1)st harmonics of the halftone frequencies. In this regard the halftone lattice can
be modeled as the sum of two cosinusoids. The components at the relevant frequencies are
distinctly stronger than surrounding components and thus are easy to identify.
In the neighborhood of a (0, 0, 1, 0)th or a (0, 0, 0, 1)st harmonic of the halftone
frequencies, the spectrum of a uniformly sampled halftone can be modeled by the product
of two sinc functions. Our algorithms and formulas for estimating the frequencies and
phases of a halftone lattice are based upon the sinc function model.
For a non-uniformly but recurrently sampled halftone, the sinc function model is modu-
lated by an extra complex envelope function. This envelope function does not introduce sig-
nicant variation in magnitude but may introduce signicant phase shift. For non-uniform
samples, the frequency estimation algorithms for uniform samples and for non-uniform
samples perform equally well. The complex phase of the modulating envelope, on the other
hand, must be compensated to obtain a correct phase estimate. Knowledge about the
sampling model is necessary to evaluate the envelope function.
119
CHAPTER 7. ESTIMATING THE HALFTONE LATTICE
The estimation algorithms and formulas developed in this chapter make use of the fact
that the DFT approximates the continuous Fourier transform at the frequency sampling
points. The continuous model is used even though the DFT is computed. The approxima-
tion makes the mathematical derivation simpler.
We can also follow the DFT model to develop estimation algorithms and formulas.
Specically, in 1D, the DFT of a truncated and uniformly sampled sinusoid is, except for a
scaling factor, the sum of two digital sinc functions. However, the digital sinc model does
not lead to a simple frequency estimation formula. Rather, a transcendental equation of
the form

A
m
1
A
m
0

sinm
sin
1m
N

results. Solving this equation by approximation leads to the same interpolation formula
in Eq. 7.4. On the other hand, the digital sinc model does lead to a more accurate phase
estimation formula:

d
0
=
arctan[Im(A
m
0
), Re(A
m
0
)]
N1
N
L(
m
0
L
p)
2 p
d
0
.
However, because other sources of error are usually larger than the error of approximating
the digital sinc by the conventional sinc, the improved accuracy in phase estimation is
usually insignicant. For example, in the tests in Section 7.3 the improved accuracy can
barely be seen when = 0.3, which is already unrealistic compared to an of 0.50.6 for
a typical scanned halftone.
120
Chapter 8
Restoration of Sampled Halftones with Moire
Patterns
8.1 Introduction
A moire pattern in a sampled halftone is usually an aliasing moire. In this chapter we
explore three approaches to reduce an aliasing moire pattern in a sampled halftone. These
include notch ltering, simulation, and relaxation. We will use three sampled halftones as
examples and attempt to suppress the moire patterns in them. These include the images in
Figure 1.1, Figure 5.11(a), and Figure 6.3. Without further qualication, the term moire
means aliasing moire in this chapter.
Since a moire pattern is a manifestation of aliasing, it is theoretically impossible to
remove it from a sampled halftone while keeping the image content intact. However, since
a moire pattern is periodic, that is, its energy largely concentrates at some individual
frequencies in the frequency domain, it is still possible to reduce the appearance of a moire
without severely damaging the image content.
An important issue in moire pattern reduction is how to recognize whether a periodic
pattern is a moire pattern or a graphic feature. If we know how the pixels of a halftone
image are sampled, it is relatively easy to identify the components of a moire pattern
deterministically. Otherwise, some heuristic method or human intervention is necessary. In
this chapter we assume that the sampling model is known.
In Section 8.2 we discuss the assessment of the quality of moire pattern reduction. We
use two subjective measures and one quantitative measure. A good measure should take
into account of two issues: 1) to what degree is the moire pattern suppressed? and 2) how
121
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
well is the image information preserved? We also implement the dual-sampling method by
Ohyama et al. [OYT
+
86] (also see Section 4.3) to provide comparisons.
In the spectrum of a sampled halftone, a moire pattern consists of sharp hills located at
harmonics of the halftone frequencies and their aliases. It is quite intuitive that, if we remove
these hills from the spectrum, the moire pattern will be suppressed. Individual frequency
component removal is sometimes called notch ltering. In Section 8.3, we demonstrate the
performance of this approach and discuss its applicability.
While the frequency and orientation of a moire pattern is determined by the halftone
lattice and the sampling lattice, its amplitude is largely determined by the low frequency
content of the source picture. This is because in the spectrum of a halftone picture the
frequency components at harmonics of the halftone frequencies are determined by the larger
features of the source picture, as discussed in Chapter 5. We may then premise that a source
picture and a low-pass ltered copy of itafter being halftoned and sampled the same way
give rise to similar moire. We may low-pass lter a sampled halftone to obtain an image
that has no halftone structure in it. Then we algorithmically halftone and sample the low-
pass ltered image to obtain a simulated sampled halftone. The dierence between the
simulated sampled halftone and the low-pass ltered halftone is a simulated moire pattern.
This simulated moire pattern can be used to compensate for the actual moire in the real
sampled halftone. In Section 8.4 we elaborate this approach.
Two pieces of information completely determine a template-dot halftone picture: the
halftone dot lattice and the sizes of the halftone dots. If we can accurately estimate both
pieces of information, we obtain the graphical content of the halftone. Suppose that we
are able to accurately estimate the halftone lattice from a sampled template-dot halftone,
and assume that we are able to accurately model the low-pass ltering eect applied to
the halftone prior to sampling. We can then compute the values of pixels in the sampled
halftone if we also know the sizes of the halftone dots. Conversely, we can hypothesize
the sizes of the halftone dots and compare the computed pixel values against the actual
pixels in the sampled halftone. We then use the dierences between the computed pixel
122
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
values and the actual pixel values to improve our hypothesis. We repeat the process until it
converges. This approach attempts to extract the graphical content of a sampled halftone
and thus implicitly overcomes the moire pattern problem. If the input halftone is a general
clustered-dot halftone, we still treat it as if it were a template-dot halftone to approximate
the graphical content it carries. We describe this approach in more detail in Section 8.5
though it has not been implemented.
8.2 Measurement of the quality of moire pattern suppression
8.2.1 Subjective measures
The principal method used in the literature to assess the quality of moire suppression,
including suppressing moire patterns when halftoning source pictures with periodic content,
is through human evaluation of the processed images. In this dissertation we also use this
method and always show the processed images. A second and more revealing subjective
measure we use is to examine the dierence between the processed image and its unprocessed
counterpart. The dierence should be only a pure moire pattern.
8.2.2 A quantitative measure
Subtle dierences between images may well be measured quantitatively. The quanti-
tative measure we use consists of two metrics. The rst metric describes how much of a
moire pattern is suppressed from the sampled halftone or how much residue moire still re-
mains in a processed image. The second metric describes how close a processed image is to
its unprocessed counterpart in image content, or how well the image content is preserved.
Let h
s
(x) be a sampled halftone image with discrete Fourier transform H
s
(u) and h
p
(x)
be the processed image with DFT H
p
(u). Let pres(h
s
, h
p
) be the preservation metric and
123
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
supp(h
s
, h
p
) the suppression metric. Then
supp(h
s
, h
p
) =
1
A

U
i
Ua

uU
i
H
2
p
(u)
[U
i
[

uU
i
H
2
s
(u)
,
pres(h
s
, h
p
) =
1
[UU
a
[

uUUa
[H
s
(u) H
p
(u)]
2
,
where U is the set of frequencies on which the DFTs are calculated, A is the number of
aliases of the halftone frequency harmonics in U, U
i
is the set of frequencies that form
the neighborhood of such an alias, [U
i
[ is the number of discrete frequencies in U
i
, and
U
a
is the union of all the U
i
s. Even though the components at the halftone frequencies
themselves may form a contrast pattern, we do not include the halftone frequencies because
they are relatively easy to lter out without loss of much image information. In general,
the smaller each metric value, the better quality the processed image. The preservation
metric is simply a mean squared error. The metrics are dependent upon the choice of U
i
s;
to compare the metric values of two processed images, the same sets of U
i
s should be used.
If we compare a sampled halftone h
s
(x) with itself using the two metrics, we will have
supp(h
s
, h
s
) equal 1 and pres(h
s
, h
s
) equal 0. The reason we use frequency domain metrics
is that a moire pattern appears only in small neighborhoods of some individual frequencies,
making the distinction of image information and the moire pattern relatively easy.
To identify the frequencies in a U
i
, we recall from Chapters 5 and 6 that an alias of a
harmonic of the halftone frequencies is of the form ks +lt+mp+nq, where s and t are the
distances between replicas in the horizontal and the vertical directions, respectively, p and q
are the halftone frequencies, and k,l,m, and n are integers with k ,= 0 and l ,= 0. The values
of s and t depend upon the sampling model. We determine the size of a neighborhood U
i
empirically. It is on the order of a few pixels in the DFT of an image. The stronger the
frequency component at the center of a neighborhood, the larger the neighborhood, because
the center component inuences a larger area due in part to leakage. Heuristic methods to
determine the size of a neighborhood can be devised if the necessity is established.
124
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
There are three major problems with the two metrics. First, they are computationally
expensive. Second, they do not apply to certain reduction methods, such the ideal notch
ltering method in Section 8.3. Third, neither metric is very accurate. To see the third
problem, note that the image content may have energy at frequencies in a U
i
. Thus,
a perfect removal of moire does not yield a zero energy ratio

uU
i
H
2
p
(u)

uU
i
H
2
s
(u)
in a U
i
for the
suppression metric. This is especially true if a U
i
is close to the zero frequency. Likewise, the
preservation metric does not include frequencies that are in U
a
; loss of image information
at those frequencies due to over-correction is not reected. Such examples are shown in
Section 8.3. Because of this inaccuracy, the metric values are not conclusively comparable
if they are calculated from images that are sampled dierently, have dierent sizes, or even
have dierent content.
The two metrics work relatively well when all the strong aliasing frequency components
are located away from the zero frequency.
The two quality metrics may be improved in several ways. For the preservation metric,
instead of treating errors at all frequencies equally, we may use the human eye sensitivity
model to weigh the errors. We may also normalize the preservation metric in some way so
that it is easy to judge the preservation quality by the metric value itself, without comparing
to the metric values of other processed images. Currently, the preservation metric tends to
have very small values.
We may improve the suppression metric by distinguishing aliases of the halftone fre-
quency harmonics that contributes most signicantly to the moire from other aliasing fre-
quencies using some heuristic method. Then we can weigh the energy ratio at dierent
aliasing frequencies to make the metric reect more accurately the state of moire appear-
ance.
The relationship of the two metrics to human perception also remains to be investigated.
125
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
8.2.3 Comparison with the dual-sampling method
Since there is no known method that attempts to restore a sampled halftone that has a
moire pattern, we implement the dual-sampling method by Ohyama et. al. [OYT
+
86] and
compare its results with those produced by our methods. The conditions of the comparison
favor the dual-sampling method, because the dual-sampling method requires two sampled
images of the same input halftone at dierent resolutions to begin with, while our methods
operate on only one sampled image of the input halftone. We intentionally set the compar-
ison conditions to favor the dual-sampling method more by applying our methods to the
lower resolution image of the two used by the dual-sampling method.
We apply the dual-sampling method to the two House examples, Figure 5.11(a) and
Figure 6.3. Since the author does not have the means to scan exactly the same (portion of
an) input picture at two truly dierent resolutions, the dual-sampling process was simulated
by rst scanning a portion of the 85 dpi House halftone using a ScanJet IIc at 400 dpi to
produce a uniformly sampled master image that has a size of 1024 by 1024 pixels, and then
digitally subsampling the master image at two dierent resolutions by bilinear interpolation
[Sch89]. Figure 5.9 shows a 512 512 pixel portion of the master image, and Figure 5.10
shows the Fourier spectrum of the master image.
In the rst example we attempt to reduce the moire pattern in the sampled halftone in
Figure 5.11(a) which is subsampled uniformly at 133
1
3
dpi (one out of every three pixels)
from the master image and has a size of 342 by 342 pixels. For the dual-sampling method
to work, we subsample the master image uniformly at 150 dpi to obtain a second sampled
image of the same halftone. Figure 8.1 shows the second sampled halftone which has a size
of 384 by 384 pixels. Figure 8.2 shows the output image from the dual-sampling method,
that is additionally low-pass ltered to eliminate the halftone structure. The moire pattern
is largely suppressed. Figure 8.3 shows the Fourier spectrum of the output image, and it is
clear that the aliasing frequency components such as those marked as (0,1,-1,1), (1,1,-1,-1),
(-1,0,2,-1), and (0,1,-2,1) in Figure 5.11(b) are indeed absent here. Figure 8.4 shows the
126
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.1: The 400 dpi scanned House halftone by a ScanJet IIc subsampled uniformly
at 150 dpi
dierence between the images in Figure 5.11(a) and Figure 8.2. Table 8.1 lists values of
the two quality metrics discussed in Section 8.2.2 applied to the image in Figure 5.11(a)
Table 8.1: Quality metric values for the moire-reduced image in Figure 8.2
suppression preservation neighborhood size
0.084 0.00093 3 3
and the image in Figure 8.2. Also listed is the size of neighborhood used to compute the
metrics.
Figure 8.4 reveals that there is some loss of information in the image in Figure 8.2 which
looks slightly grainy. The loss of image information is primarily due to the misalignment
between the two input sampled images to the dual-sampling method. The 133
1
3
dpi sampled
image covers the whole 1024 1024 pixel area of the master image. The 150 dpi sampled
127
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.2: The image produced by the dual-sampling method using the images in Fig-
ure 5.11(a) and Figure 8.1 as input
Figure 8.3: The Fourier spectrum of the image in Figure 8.2
128
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.4: The dierence between the image in Figure 5.11(a) and the image in Figure 8.2
image covers a 1022
1
3
1022
1
3
pixel area of the master image. The coverage of the two
sampled images diers by about 0.3%. Experiments show that if the coverage of the two
images diers by 0.5%, the output image exhibits clear distortion of image content.
The second example is related to the sampled halftone in Figure 6.3 which is the House
halftone image scanned at 150 dpi by a ScanJet IIc. We sample the master image uniformly
in the vertical direction and non-uniformly in the horizontal direction using the improved
binary rate multiplication algorithm. Two sampled images are created, one at 150 dpi
resolution with a size of 384 by 384 pixels and the other at 160 dpi with a size of 408 by 408
pixels. The resolutions of the two sampled images are carefully chosen so that the coverage
of the two images diers by only about 0.016%. Figure 8.5 and Figure 8.6 show the two
sampled images, respectively. Figure 8.5 is indeed similar to Figure 6.3. Figure 8.7 shows
the output image from the dual-sampling method, and Figure 8.8 shows the dierence
between the images in Figure 8.5 and Figure 8.7. Table 8.2 lists values of the two quality
129
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.5: The 400 dpi uniformly scanned House halftone subsampled at 150 dpi using
the ScanJet IIc subsampling procedure
130
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.6: The 400 dpi uniformly scanned House halftone subsampled at 160 dpi using
the ScanJet IIc subsampling procedure
131
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.7: The image produced by the dual-sampling method using the images in Figure 8.5
and Figure 8.6 as input
132
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.8: The dierence between the images in Figure 8.5 and Figure 8.7
Table 8.2: Quality metric values for the moire-reduced image in Figure 8.7
suppression preservation neighborhood size
0.41 0.00017 3 3
metrics applied to the image in Figure 8.5 and the image in Figure 8.7, as well as the size
of neighborhood used to compute the metrics. Moire reduction is not as successful as in the
rst example because some moire components in the two sampled images have the same
frequencies, as indicated in Section 6.3.3.
8.3 Notch ltering
The notch ltering method is closely related to the quantitative measure discussed in
the previous section. It includes six steps:
133
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
1. compute the DFT of the sampled halftone;
2. estimate the halftone frequencies;
3. compute the aliases of halftone frequency harmonics that fall in the DFT spectral
band;
4. smooth the DFT components in a small neighborhood of each aliasing halftone fre-
quency harmonic;
5. low-pass lter the DFT components to remove the halftone structure;
6. compute the inverse DFT of the smoothed DFT.
The frequencies in Step 3 are easily computed, since each has the form ks +lt +mp +nq.
For Step 4, the neighborhood in which to smooth the DFT is determined empirically. Two
types of smoothing are investigated: 1) zero-out the frequency components in a neighbor-
hood (ideal notch ltering), and 2) interpolate new components interior to a neighbor-
hood using components on the neighborhood perimeter. The interpolation uses the bilinear
blending functions proposed by Gordon ([Gor71]). It appears that the ideal notch ltering
produces better visual results. The low-pass ltering of step 5 often appreciably improves
the appearance of the output image because it eliminates the (potential) non-aliasing moire.
As a rst example, we apply the notch ltering algorithm to the uniformly sampled
halftone image in Figure 5.11(a) using dierent neighborhood sizes and smoothing meth-
ods. Figures 8.9(a) and (b) show, respectively, two ideal notch ltered images. A 33
neighborhood size is used for Figure 8.9(a), and a 77 neighborhood size is used for Fig-
ure 8.9(b). Figures 8.10(a) and (b) show, respectively, the dierences between the image
in Figure 5.11(a) and the images in Figures 8.9(a) and (b). Figure 8.11 shows the Fourier
spectrum of the image in Figure 8.9(a).
Neither image in Figure 8.9(a) and (b) shows successful moire reduction. Figure 8.9(a)
is under-corrected. Figure 8.9(b) is also slightly under-corrected, and a ringing distortion
appears (in the house area). The dierence image in Figure 8.10(b) shows the ringing more
134
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
(a) 3 3 neighborhood
(b) 7 7 neighborhood
Figure 8.9: Two ideal notch ltered images of the sampled halftone in Figure 5.11(a)
135
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
(a) 33 neighborhood
(b) 77 neighborhood
Figure 8.10: The dierence between (a) the image in Figure 5.11(a) and the image in
Figure 8.9(a) and (b) the image in Figure 5.11(a) and the image in Figure 8.9(b)
136
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.11: The Fourier spectrum of the image in Figure 8.9(a)
clearly. Use of larger neighborhoods also causes ringing as well as obvious loss of image
information. Ringing is a characteristic of ideal ltering. Experiments to use an inverted
Hamming window at each ltering notch show that ringing is slightly reduced, but the
residue moire pattern is also stronger.
Figure 8.12 shows a notch ltered image of the same sampled halftone in Figure 5.11(a)
using blending interpolation in a 33 pixel neighborhood of each aliased halftone frequency
harmonics. Figure 8.13 shows the dierence between the images in Figure 5.11(a) and
Figure 8.12. The moire pattern is somewhat suppressed in the center portion of the image,
but severe ringing distortion appears near the boundary of the image.
Table 8.3 lists values of the two quality metrics for the three notch ltered images.
The metrics fail on the notch ltered images because notch ltering smoothes most of the
frequency components that the suppression metric accounts for and does not aect most of
the frequency components that the preservation metric accounts for. If an image is ideal
notch ltered, and the metric computation uses exactly the same neighborhood size as the
137
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.12: The sampled halftone in Figure 5.11(a) notch ltered using blending interpo-
lation in 3 3 pixel neighborhoods
Table 8.3: Quality metric values for the moire-reduced images in Figure 8.9 and Figure 8.12
image suppression preservation neighborhood size
Figure 8.9(a) 0.000002 0.00043 3 3
Figure 8.9(b) 0.000002 0.00063 3 3
Figure 8.12 0.11 0.00056 3 3
ltering, both metrics theoretically have zero value. The non-zero metric values in Table 8.3
for ideal notch ltering are due in part to the fact that the inverse transformed image of
the ltered DFT spectrum has pixels whose values exceed the representable range, and in
implementation, such pixel values are converted to the nearest value in the range. We will
not list the quality metric values for the next two notch ltering examples.
Notch ltering is unsuccessful in this example because some of the aliased halftone
frequency harmonics are close to the zero frequency. The image content has a non-negligible
amount of energy at those frequencies. Ideal notch ltering removes that energy and also
138
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.13: The dierence between the image in Figure 5.11(a) and the image in Figure 8.12
causes ringing.
An examination of the Fourier spectra of several natural images, including the original
House image, indicates that such a spectrum is not smooth, especially near the zero
frequency. Therefore, articial smoothing using blending functions also distorts the image,
especially the phase information in a spectrum.
As a second example, we apply notch ltering to the scanned halftone image in Fig-
ure 6.3. The image has a size of 512 512 pixels. Figure 8.14 shows an ideal notch ltered
output image. Since the halftone is scanned by a ScanJet IIc at 150 dpi, an alias of a
harmonic of the halftone frequencies has the form of ks + lt + mp + nq, where [s[ = 50
dpi and [t[ = 150 dpi. Table 8.4 lists the sizes of neighborhoods of the aliased halftone
harmonics used in ltering. Here we use dierent sizes of neighborhoods to show how the
algorithm can be adapted to images that are scanned dierently. Experiments show that
better visual results are sometimes obtained by using non-square neighborhoods, as shown
139
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.14: An ideal notch ltered image of the scanned halftone in Figure 6.3
140
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Table 8.4: Neighborhood sizes used in notch ltering to obtain the image in Figure 8.14
harmonic (any , any , 0, 1) (1, 0, 1, 1) (2, 0, 1, 1) others
order (any , any , 1, 0)
size 7 7 3 9 5 17 5 5
in Table 8.4, perhaps because the image has non-uniform pixels horizontally. Figure 8.15
shows the Fourier spectrum of the image in Figure 8.14, and Figure 8.16 shows the dierence
between the images in Figure 6.3 and Figure 8.14. Notch ltering is largely successful in
this example because all the strong aliasing frequency components are relatively far away
from the zero frequency.
As a last example, we apply the notch ltering algorithm to a 512512 pixel portion of
the Mercedes image in Figure 1.1 which was scanned by an unknown scanner. Figure 8.17
shows the square image. By visually examining the Fourier spectrum of the image, which
is shown in Figure 8.18, we can infer that the image was scanned non-uniformly in both
horizontal and vertical directions. We can also guess that the spacing pattern of the pixels is
a variant of about 2,3,3 in both directions. We use this guessed information in the notch
ltering algorithm to determine (guess) the distances [s[ and [t[ between replicas to be one
third of the nominal sampling frequency. Figures 8.19, 8.20, and 8.21 showthree images thus
obtained. Ideal notch ltering is performed in Figure 8.19 and Figure 8.20, and blending
interpolation is performed in Figure 8.21. For Figure 8.19, a neighborhood size of 1 by 1
pixel is used for all aliasing halftone frequency harmonics. For Figure 8.20 and Figure 8.21,
a neighborhood size of 3 by 3 pixels is used for all aliasing halftone frequency harmonics.
Figures 8.22, 8.23, and 8.24 show the dierences between the image in Figure 8.17 and the
images in Figures 8.19, 8.20, and 8.21, respectively. Figure 8.25 shows the Fourier spectrum
of the ideal notch ltered image in Figure 8.20. Notch ltering is unsuccessful because some
of the aliasing frequency components are extremely close to the zero frequency. It is also
possible that our guess of the pixel spacing pattern is not accurate.
141
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.15: The Fourier spectrum of the image in Figure 8.14
142
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.16: The dierence between the images in Figure 6.3 and Figure 8.14
143
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.17: A 512512 pixel portion of the Mercedes image in Figure 1.1
144
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.18: The Fourier spectrum of the Mercedes image in Figure 8.17
145
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.19: The Mercedes image ideal notch ltered using 11 pixel neighborhoods
146
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.20: The Mercedes image ideal notch ltered using 33 pixel neighborhoods
147
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.21: The Mercedes image notch ltered using blending smoothing in 33 pixel
neighborhoods
148
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.22: The dierence between the images in Figure 8.17 and Figure 8.19
149
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.23: The dierence between the images in Figure 8.17 and Figure 8.20
150
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.24: The dierence between the images in Figure 8.17 and Figure 8.21
151
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.25: The Fourier spectrum of the ideal-notch ltered Mercedes image in Figure 8.19
152
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
To conclude this section, we note that notch ltering is an eective method when all the
aliasing frequency components in a sampled halftone are far away from the zero frequency
relative to the majority of frequency components of the image content.
8.4 Simulating the halftoning and the sampling processes
8.4.1 Overview of the method
In contrast to the notch ltering method which operates in the frequency domain, the
simulation method is primarily a spatial domain technique. It consists of ve steps:
1. from the input sampled halftone h
s
(x) estimate the halftone frequencies and the oset
between the halftone dot lattice and the sampling lattice;
2. low-pass lter h
s
(x) to obtain a defocused version h
d
(x) of the source picture that is
free of halftone dot structure.
3. using the estimated halftone frequencies and oset, algorithmically halftone h
d
(x) to
obtain a simulated halftone h
dh
(x); then sample h
dh
(x) to obtain a simulated sampled
halftone image h
sim
(x) of the source picture;
4. add h
d
(x) h
sim
(x) to h
s
(x) to obtain a moire reduced image;
5. optionally low-pass lter the moire reduced image to eliminate the halftone dot struc-
ture.
Steps 24 can be iterated multiple times. Dene T,H, and 5 to be the defocusing (low-pass
ltering), halftoning, and sampling operators respectively. We can describe an iteration of
steps 24 as
h
j+1
s
(x) = h
s
(x) +cT[h
j
s
(x)] 5HT[h
j
s
(x)], (8.1)
where h
j
s
(x) (j 0) is the restored image from the jth iteration, h
0
s
(x) = h
s
(x), and c is a
gain factor that controls the amount of correction applied in each iteration.
153
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
The low-pass ltering in step 2 prevents the interference between the halftone dot struc-
ture in the sampled halftone and the simulated halftoning process in step 3. The ltering
does not need to be accurate and can be very aggressive. Only components with very low
frequencies in the sampled halftone need to be retained, in contrast to the low-pass ltering
in step 5 which should eliminate only the halftone dot structure while minimally aecting
the image content. The low-pass ltered image h
d
(x) from step 2 may have a moire pattern.
This moire pattern has a low contrast compared to the dynamic range of pixel values that a
normal image may have, because it is the result of aliasing halftone frequency components
which have low energy compared to that of the image content. This moire pattern does
not signicantly aect the formation of the new moire pattern in the simulated sampled
halftone image in step 3.
Step 3 is the crucial step of the method. Section 8.4.2 provides more details about its
implementation. Here we list four general issues regarding the step:
1. Realize that there is no need to separate the simulated halftoning process from the
simulated sampling process. Rather, we can just halftone the low-pass ltered image
h
d
(x) from step 2 in the neighborhood of each sampling point and then compute the
pixel value for that sampling point.
2. The physical positions of the pixels in the simulated sampled halftone image h
sim
(x)
should be the same as those of the pixels in the real sampled halftone image h
s
(x).
3. The simulated halftoning process should resemble the one from which the actual input
halftone is produced. Though the input halftone is likely to be of clustered-dot ordered
dither type, it is dicult to estimate the corresponding halftone function (dened in
Chapter 3). There has been research on the estimation of halftone function given
the exact binary pixel representation of a digital halftone image (e.g. one that is
sent to a laser printer) [AA92]. But an input sampled halftone image under our
consideration has a much lower resolution and has been distorted by the low-pass
ltering by the scanner. For simplicity, we use template-dot halftone with circular
154
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
dot to approximate the input halftone. (Other dot shapes can also be implemented).
The simulated halftone is dened by a set of dots where each dot is dened by its
center position and a radius.
4. The simulated sampling process should take into account the low-pass ltering eect
by the scanner. The simulated halftoning process creates only halftone dots with
ideally sharp boundaries.
Step 4 compensates the moire pattern in the input sampled halftone by adding a neg-
ative moire pattern to the input sampled halftone. The use of a gain factor in step 4
provides a choice of the amount of compensation. It may be possible to vary the gain factor
adaptively according to the image content. In the experiments later in the section, however,
the gain factor is 1.
8.4.2 Generating the simulated sampled halftone image
The basic algorithm
Using the notation of the previous subsection, we have
h
dh
(x) = H[h
d
(x)],
h
sim
(x) = 5[h
dh
(x)].
The algorithm below generates h
sim
(x) from h
d
(x).
1. for each pixel of h
sim
(x) do steps 25;
2. compute the physical position r = (x
r
, y
r
) of the sampling point corresponding to the
pixel;
3. compute the positions of the halftone dots of h
dh
(x) that are in the vicinity of r;
4. compute the sizes of those halftone dots of h
dh
(x) using h
d
(x) as source;
155
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
5. compute the value of the pixel of h
sim
(x) by integrating over the dots of h
dh
(x)
obtained in steps 3 and 4.
Step 2 of the above algorithm is simple when pixels are uniformly sampled. Remember
that the image coordinate system we use coincides with the sampling lattice as shown in
Figure 7.1(a). We have
x
i
=
i
[s[
,
y
j
=
j
[t[
,
where s and t are the sampling frequencies in the two directions. When pixels are non-
uniformly sampled, the non-uniformity needs to be taken into account.
The relative position of a sampling point and the halftone dots
To compute the value of the pixel corresponding to a sampling point, it is necessary
to know the positions of the halftone dots in the vicinity of the sampling point. The
sampled halftones we consider are all scanned by scanners that use CCD arrays as the
image acquisition device. Figure 8.26 shows the sampling point r of a CCD sensor and a
few halftone dots in the vicinity. The solid lattice is the halftone lattice, and the circles
represent the halftone dots. The dotted lines show the boundaries of halftone cells. The
vectors a and b dene the shape of the halftone lattice and the halftone cells.
To compute the positions of halftone dots that are close to a sampling point r, we think
of r as a vector and express it in terms of the vectors a and b, as shown in Figure 8.27. In
the gure the solid lattice is the halftone lattice. We have
r = r
a
+r
b
= a +b,
where the coecients and are easy to determine using vector algebra. Note that by our
choice of the coordinate system, r is always in the range x 0, y 0, though r
a
or r
b
may
not be.
156
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
r
b
a
Figure 8.26: The relative position of a CCD sensor and the halftone dots in its vicinity
157
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
r
y
x
r
r
a
b
a
b
d
b
d
a
Figure 8.27: The position of a sampling point expressed in terms of the halftone lattice
vectors
158
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
If we also know, in the directions of a and b, the oset d
a
and d
b
of the halftone dot
lattice with respect to the origin of the coordinate system, as shown also in Figure 8.27, the
position of one halftone dot that is an immediate neighbor of the sampling point is given
by
_
d
a
[a[
+
_

d
a
[a[
__
a +
_
d
b
[b[
+
_

d
b
[b[
__
b.
The positions of the other halftone dots are given by
_
d
a
[a[
+
_

d
a
[a[
_
+k
_
a +
_
d
b
[b[
+
_

d
b
[b[
_
+l
_
b,
where k and l are integers. In particular, For k, l =0 or 1, we obtain the four halftone dots
that are the immediate neighbors of the sampling point.
In the above calculation of halftone dot positions, the halftone dot lattice is expressed
in terms of a, b, d
a
, and d
b
, which are parallel to the lattice. In Chapter 7, we estimate a
dierent set of parameters, P, Q, d
p
, and d
p
, which are perpendicular to the halftone dot
lattice they dene. Figure 8.28 shows the relationship between the two sets of parameters.
a, b, d
a
, and d
b
can be calculated from P, Q, d
p
, and d
p
using simple trigonometry.
Computing the size of a halftone dot
To compute the size of a halftone dot, we average h
d
(x) near the dot and then normalize
the average to the range [0, 1]. Let the normalized pixel value be w. We dene the local
density of h
d
(x) to be z = 1 w. This denition of density is dierent from that used
in photography. Given a and b, we know the area of a halftone cell. The area of the
halftone dot is then such that its ratio to the area of the halftone cell is equal to z. When
a dot is small, its shape is circular; when a dot is large, its shape is the intersection of the
parallelogram halftone cell and a circle, as shown in Figure 8.29. In implementation the size
of a dot is always expressed in terms of the radius of the corresponding circle. For eciency,
the relationship of h
d
(x) pixel value to halftone dot size is pre-calculated and stored in a
lookup table. Figure 8.30 shows a pixel value to dot size correspondence curve for a square
halftone that has a unit side length.
159
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
P
a
b
x
y
d
d
d
d
p
q
a
b
Q
Figure 8.28: Two sets of parameters to dene a halftone dot lattice
Figure 8.29: A large halftone dot in a parallelogram cell
160
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
0 50 100 150 200 250
pixel value
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
d
o
t

r
a
d
i
u
s
Figure 8.30: The pixel value to dot radius correspondence curve for a square halftone cell
that has a unit side length
Three notes can be made about the generation of halftone dots. 1) We have used
template-dot scheme with circular dots to model the halftoning process for simplicity. We
can also consider more complex schemes such as diamond-shaped template-dot, or even
clustered-dot ordered dither. However, the purpose of the process is not to create a high
quality halftone but one that creates a moire pattern similar to the moire pattern in the
input sampled halftone h
s
(x) when sampled the same way. 2) Since h
d
(x), the input image
to the simulated halftoning process, has only low frequency content, experiments show that
using the average value of pixels of h
d
(x) near a halftone dot to estimate the size of the dot
is almost equivalent to using the value of the one pixel of h
d
(x) that is closest to the dot.
3) For a scanned halftone image, the pixel values may not be exactly linear to the sizes of
the halftone dots in the input halftone picture. If a more accurate dot size to pixel value
relationship is known, it can be used in the generation of simulated halftone dots.
161
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Computing the pixel values in the simulated sampled halftone
Given an input picture f(x) to a scanner, the reading of a CCD sensor at a point r can
be modeled by the convolution
_

f(x)psf(r x)dx,
where psf(x) is the point spread function of the scanner optics, including the aperture
eect of the nite size CCD photosite. psf(x) can be modeled by a Gaussian function:
psf(x) = e
x
2
, where controls the amount of ltering. For modeling convenience, we
write psf(x) in such a form that its gain is 1, that is, the lter does not amplify or diminish
the energy of the input but only redistributes it spatially.
For the ScanJet IIc, the CCD sensor moves in the vertical direction of the input picture
while an exposure is taking place. In the following, we consider modeling the formation of
the value of an output image pixel from this particular scanner.
For the case of 300400 dpi scanning resolution, the value of a pixel in the output image
from the ScanJet IIc can be modeled by
pixel(r) =
1
2
_
yr+
yr
__

psf(x

x)f(x)dx
_
dy

=
_

f(x)
_
1
2
_
yr+
yr
psf(x

x)dy

_
dx
=
_

f(x)lp(r x)dx. (8.2)


Here r = (x
r
, y
r
) is the nominal sampling point corresponding to the pixel, x

= (x
r
, y

) is
a point on the vertical line with x-coordinate x
r
, and is half of the distance that the CCD
sensor travels during one exposure. These are shown in Figure 8.31(a), where the square
represents the nominal sampling point, and the thick vertical bar represents the path along
which the CCD sensor moves. lp(x) =
1
2
_

psf(x x

)dy

, where x

= (0, y

), is the
cumulative low-pass lter function of both the optics and the motion of the CCD. lp(x)
looks very much like a Gaussian function except it is usually fatter but approaches the
Gaussian function when the length of the integration interval 2 approaches zero.
162
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
r
r 2
2
2
(b) (a)
Figure 8.31: The CCD motion path(s) and the nominal sampling point for one output pixel
of the ScanJet IIc for (a) 300400 dpi resolution and (b) 150297 dpi resolution
For the case of 150297 dpi scanning resolution, the readings of two adjacent CCD
sensors in the horizontal direction are averaged to form the value of one output pixel. The
value of the pixel can be modeled by
pixel(r) =
1
2
_
1
2
_
yr+
yr
__

psf(x

1
x)f(x)dx
_
dy

+
1
2
_
yr+
yr
__

psf(x

2
x)f(x)dx
_
dy

_
=
_

f(x)
_
1
4
_
yr+
yr
_
psf(x

1
x) + psf(x

2
x)

dy

_
dx
=
_

f(x)lp(r x)dx. (8.3)


Here x

1
= (x
r
, y

) and x

2
= (x
r
+ , y

) are, respectively, points on the two lines
along which the two CCD sensors travel, and is half of the distance between the two
CCD sensors. These are shown in Figure 8.31(b), where the square represents the nominal
sampling point, and the two thick vertical bars represent the paths along which the two
CCD sensors move. lp(x) =
_
1
4
_

_
psf(x x

1
) +psf(x x

2
)

dy

_
, where x

1
= (, y

)
and x

2
= (, y

), is the cumulative low-pass lter function of the optics, the motion of the
CCDs, and the average of two CCD readings.
Note that, strictly speaking, 2 is the distance between the nominal sampling points
of two adjacent CCD cells, which is much larger than the physical spacing between two
163
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
)
sensor integration
region r (
r
b
a
Figure 8.32: The region that inuences the value of the pixel at a sampling point
adjacent CCD cells on a CCD array IC chip. For brevity, we call 2 the distance between
two adjacent CCD sensors.
In either the 300400 dpi or the 150297 dpi scan resolution case, lp(x) decreases quickly
as [x[ increases. In practice, the innite integral in x in Eq. 8.2 and Eq. 8.3 can be approx-
imated by a nite integral:
pixel(r)
_
x(r)
f(x)lp(r x)dx, (8.4)
where (r) is a near-rectangular region centered at r in which lp(r x) is signicant.
Figure 8.32 shows the same sampling point r to halftone dot conguration as Figure 8.26
but with the region (r) marked.
For the case of 60147 dpi scan resolution, each output pixel is the average of four CCD
readings. The analysis is analogous to that for the 150297 dpi case.
164
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
In the implementation of the simulated sampling process, the calculation of a pixel value
corresponding to a sampling point involves four or nine nearby halftone dots, depending
upon the sampling rate. We could substitute the general input picture f(x) in Eq. 8.4 with
h
dh
(x), the binary simulated halftone of h
d
(x). However, since it is dicult to compute the
integral in Eq. 8.4, we use the summation
pixel(r)
[(r)[
B

x
ij
(r)
h
dh
(x
ij
)lp(r x
ij
), (8.5)
to approximate the integral. Here x
ij
s are a set of B points that are uniformly spaced in
(r) whose area is [(r)[. Eq. 8.5 is a weighted average of h
dh
(x) in (r) with weights
lp(r x
ij
)s. The values of lp(r x
ij
)s are independent of r and are pre-calculated. In the
examples to be presented later in this section, the value of B is on the order of 100 to 400.
The advantage of the above summation method is its simplicity, though it may not be very
ecient.
To compute the function lp(x), it is necessary to know the coecient for the Gaussian
function psf(x) and the size of (r). For the ScanJet IIc, we know that the optics virtually
eliminate frequencies above 200 dpi and attenuate by about 50% the 100 dpi frequency.
This corresponds approximately to a Gaussian lter e
0.00002u
2
in the frequency domain,
as shown in Figure 8.33(a) in 1D prole, with [u[ measured in dpi. (More accurately, the unit
is cpi, cycles per inch.) The spatial domain representation of the lter is lp(x) = e


0.00002
x
2
( =
1
0.00002
), as shown in Figure 8.33(b) in 1D prole, with [x[ measured in inches. lp(x) is
signicant in an area that is about 0.01 inches across or about ve 400 dpi pixel spacing (
5
400
inches). When the scan resolution is 400 dpi, a CCD sensor moves at about
1
400
inches per
exposure time. in Eq. 8.2 is about
0.5
400
inches, and (r) in Eq. 8.4 is about
5
400

5
400
inches.
Computation shows that lp(x) in this case is similar to a Gaussian function. Figure 8.34(a)
and (b) show, respectively, the 3D surface plot and 2D contour plot of lp(x) for 400 dpi
scan resolution. The unit of measure in the gure is
1
400
inches. When the scan resolution is
150 dpi, a CCD sensor moves about
2.67
400
inches per exposure time, and the digital low-pass
ltering averages the output of two CCD sensors that are
1
400
inches apart. and in
165
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
0 50 100 150 200 250
frequency (dpi)
0
0.2
0.4
0.6
0.8
1
0 0.002 0.004 0.006 0.008
spatial coverage (inches)
0
0.2
0.4
0.6
0.8
1
Figure 8.33: The mathematical model for the low-pass ltering by the optics of the ScanJet
IIc: in frequency domain (a) and in spatial domain (b)
-2
0
2
x -2
0
2
y
lp(x,y)
-2
0
2
x
-3 -2 -1 0 1 2 3
-3
-2
-1
0
1
2
3
(a) (b)
Figure 8.34: The spatial domain low-pass ltering model by the optics and moving carriage
of the ScanJet IIc at 400 dpi scan resolution: (a) 3D plot, (b) contour plot
166
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
-2
0
2
x -2
0
2
y
lp(x,y)
-2
0
2
x
-3 -2 -1 0 1 2 3
-3
-2
-1
0
1
2
3
(a) (b)
Figure 8.35: The spatial domain low-pass ltering model by the optics and moving carriage
of the ScanJet IIc at 150 dpi scan resolution: (a) 3D plot, (b) contour plot
Eq. 8.3 are, respectively, about
1.35
400
and
0.5
400
inches, and (r) in Eq. 8.4 is about
6
400

7
400
inches. Figure 8.35(a) and (b) show, respectively, the 3D surface plot and 2D contour plot
of lp(x) for 150 dpi scan resolution.
Finally, recall that the boundaries of a halftone dot printed by a real printer are slightly
defocused, whereas in our simulation the dots have ideally sharp boundaries. In the im-
plementation, we use the same low-pass lter function psf(x) to account for the low-pass
ltering of the printing device. The value of is reduced to increase the amount of low-pass
ltering. Experiments show that a =
1
0.00003
appears to give the best results for halftone
images scanned from input halftone pictures that are printed using an HP LaserJet 4 printer.
8.4.3 Experimental results
We rst apply the simulation method to the image in Figure 5.11(a). Since the image
is obtained by selecting every third pixel in each direction from a master image scanned
at 400 dpi by a ScanJet IIc, it can be considered to be uniformly sampled from the input
halftone picture at 133
1
3
dpi preceded by the low-pass ltering by the ScanJet IIc for a 400
167
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.36: The output image obtained by applying the simulation method on the image
in Figure 5.11(a)
dpi scan. Figure 8.36 shows an image obtained after two iterations. Figure 8.37 shows the
dierence between the images in Figure 5.11(a) and Figure 8.36. Table 8.5 lists values of
the two quality metrics for the image in Figure 8.36.
Table 8.5: Quality metric values for the moire-reduced image in Figure 8.36
suppression preservation neighborhood size
0.21 0.00073 3 3
The output image still exhibits some moire pattern. More iterations of the simulation
method without variation in the parameter settings introduce noticeable distortion. More
research is needed to improve the method. One current weak point of the method is its sim-
ulated halftoning process. Examination of the input House halftone picture reveals that
in the sky and the shadow areas the halftone dots are not symmetric vertically and horizon-
tally at all; they are connected diagonally from upper left to lower right but are separated
168
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.37: The dierence between the images in Figure 5.11(a) and Figure 8.36
diagonally from upper right to lower left. The halftone almost looks like a line h alftone.
Since the input halftone picture is halftoned digitally and printed using a laser printer, the
halftone dots cannot be symmetric for all levels of gray. However, the simulated halftoning
process generates dots that are symmetric vertically and horizontally. The printed dots
may be more accurately modeled by 45 degree oriented ellipses instead of circles.
Overall, the simulation method does not perform as well as the dual-sampling method,
but outperforms the notch ltering method in this example. Note, though, that the preser-
vation metric has a smaller value here than for the dual-sampling method processed image
in Figure 8.2. From the dierence images in Figure 8.4 and Figure 8.37 it is also clear that
the simulation method preserves image content better than the dual-sampling method in
this example.
In the second example, we apply the simulation method to the image in Figure 6.3.
This image was scanned at 150 dpi using a ScanJet IIc. In the horizontal direction, the
169
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
image pixels are not uniform and have a spacing pattern that is one of 0.75, 1.125, 1.125,
1.125, 0.75, 1.125, or 1.125, 1.125, 0.75, where the unit of measure is
1
150
inches, the
nominal sampling period.
We need to know which of the three patterns the image pixels actually possess. For this
purpose, we run steps 13 of the simulation method on the scanned image three times, each
time using one of the three patterns in step 3, to obtain three simulated sampled halftone
images of the same defocused scanned image. We then compute the cross-correlation of the
scanned image with each of the three simulated images. The simulated image giving rise to
the largest cross-correlation is sampled using the correct spacing pattern. We are then able
to apply the whole simulation method to the scanned image to reduce the moire pattern in
it.
Figure 8.38 shows a processed image after two iterations of the simulation method.
Figure 8.39 shows the dierence between the images in Figure 6.3 and Figure 8.38. Some
moire pattern still remains, and it is clear the some image information is lost in the processed
image. The dierence image shows that the processed image has a higher contrast than the
input image. The loss of image information appears to be related to the inaccuracy of the
simulated halftoning process in modeling the actual process from which the input halftone
was generated. Table 8.6 lists values of the two quality metrics for the processed image in
Table 8.6: Quality metric values for the moire-reduced image in Figure 8.38
suppression preservation neighborhood size
0.42 0.00022 3 3
Figure 8.38.
Comparison of Figure 8.38 to Figures 8.7 and 8.14 shows that, in this example, the
notch ltering method and the simulation method produce visually comparable results, and
both methods arguably outperform the dual-sampling method. The notch ltering method
removes more moire, and the simulation method causes no ringing.
Comparison of Table 8.6 to Table 8.2 shows that, in this example, the simulation method
170
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.38: The output image obtained by applying the simulation method on the image
in Figure 6.3
171
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
Figure 8.39: The dierence between the images in Figure 6.3 and Figure 8.38
172
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
and the dual-sampling method render comparable moire reduction but the dual-sampling
method preserves image information better. However, since the images in Figures 8.38 and
8.7 do not have the same size, the metric comparison is not very conclusive.
We are not able to apply the simulation method to the Mercedes image in Figure 1.1
because we do not know how that image was scanned. Examination of the Fourier spectrum
of the image indicates that in both horizontal and vertical directions the image pixels have
a spacing pattern that is very close to one of 0.75, 1.125, 1.125, 1.125, 0.75, 1.125, or
1.125, 1.125, 0.75, but not exactly one of the three patterns. Attempts were made to run
the simulation method on the image assuming that the image had been sampled with each
of the nine possible combinations of the above three spacing patterns in the two directions,
but the results were not good. This again shows the heavy dependence of the simulation
method upon the knowledge of the sampling model.
8.5 The Relaxation method
In the relaxation method we attempt to answer this question: what input halftone
picture could have caused the sampled halftone image we have? Assuming that we know
the sampling model, we are able to determine the halftone dot lattice of the input halftone
picture accurately. If the input halftone is of the template-dot type, that is, a halftone dot
is always centered in its halftone cell and has a xed shape, the only other information
we need to completely specify the halftone picture is the size of each halftone dot. More
specically, since we know the positions of the halftone dots, which form the halftone dot
lattice, we can express the halftone picture as a discrete 2D function g(I), where I = (i
a
, i
b
)
is the index variable for halftone dots, and g(I) is the size of the halftone dot with index
I. The physical position x
I
of a dot I can be calculated using vector algebra, similar to the
vector calculation in Section 8.4.2. We can also express the sampled halftone as a discrete
2D function h
s
(J), where J = (j
x
, j
y
) is the index variable for pixels in the sampled image,
and h
s
(J) is the value of the pixel with index J. Also as in Section 8.4.2, we can calculate
173
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
the position x
J
of the sampling point corresponding to pixel J using the sampling model.
We know h
s
(J), and we want to know g(I).
From Section 8.4.2 we see that the value h
s
(J) of a pixel J depends upon the sizes of
the halftone dots in a neighborhood (x
J
) of x
J
, and the dependence relationship is an
integral that collectively involves all the dots in (x
J
). Conceptually and computationally,
we can also consider the contribution to h
s
(J) from each dot in (x
J
) separately and sum
the individual contribution up to obtain h
s
(J). Let dot(I) be the area occupied by dot I,
and let lp(x) be as dened in Eqs. 8.2 and 8.3. We have
h
s
(J) =

x
I
(x
J
)
_
xdot(I)
lp(x
J
x)dx
=

x
I
(x
J
)
[x
J
, x
I
, g(I)], (8.6)
where [x
J
, x
I
, g(I)] =
_
xdot(I)
lp(x
J
x)dx is the contribution from dot I. [x
J
, x
I
, g(I)]
is a function of g(I) because the latter determines the size of dot(I). Strictly speaking, the
integral on lp(x
J
x) should be performed only in (x
J
) dot(I). However, since lp(x)
diminishes quickly as [x[ increases, for simplicity, we can use a large enough (x
J
) and
count a dot as either completely inside or completely outside (x
J
). Figure 8.40 shows a
sampling point x
J
, its neighborhood (x
J
), and a few dots near x
J
among which is one
that has a position x
J
and a dot area dot(I).
If we have an estimate g(I) of g(I) for each halftone dot I and substitute g(I) for g(I)
in Eq. 8.6, we obtain an estimated value for pixel J. The dierence between the estimated
pixel value and the true pixel value h
s
(J) is the error of the estimated pixel value.
We try to nd better estimate g(I) to reduce the errors in the estimated pixel val-
ues. We relate each halftone dot I to pixels whose corresponding sampling points fall in a
neighborhood (x
I
) of the halftone dot. Dene
err(I) =

x
J
(x
I
)
_
_
_
h
s
(J)

I
(x
J
)
[x
J
, x

I
, g(I

)]
_
_
_
. (8.7)
err(I) is the cumulative error of the estimated pixels whose sampling points are close to x
I
.
174
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
( ) x
J
dot( )
I
x
I
J
x
Figure 8.40: A sampling point and the dots in or near its neighborhood
err(I) is a function of g(I

)s that appear on the right-hand-side of Eq. 8.7, where a


halftone dot I

is in the neighborhood (x
J
) of a sampling point x
J
that falls in the neigh-
borhood (x
I
) of halftone dot I. In fact, halftone dot I itself is in (x
J
) for all x
J
s in (x
I
)
if the size of (x
I
) is smaller than (x
J
). Figure 8.41 shows a few halftone dots represented
by circles and a few sampling points represented by small squares. The position of one
halftone dot is marked as x
I
, and the solid large square represents (x
I
). Four sampling
points fall in (x
I
), one being marked as x
J
. The two large dotted squares represent the
neighborhoods of two sampling points, respectively, one being (x
J
). To clarify the centers
of neighborhoods, a dashed line connects the center and the border of a neighborhood.
We take the partial derivative of err(I) with respect to g(I) in Eq. 8.7. All the terms
175
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
x
J
( )
x
I
( ) x
I
J
x
Figure 8.41: The overlap of the neighborhood (x
I
) of a halftone dot at x
I
and the neigh-
borhoods of the sampling points that fall in (x
I
)
that do not contain g(I) drop out so that
err(I)
g(I)
=

x
J
(x
I
)
_

[x
J
, x
I
, g(I)]
g(I)
_
. (8.8)
Let err(I) be the change in err(I) due only to the change g(I) in g(I). We can write
the Taylor expansion of err(I) as
err(I) =
err(I)
g(I)
g(I) +

2
err(I)
2 g
2
(I)
g
2
(I) + .
For a rst order approximation, we have
err(I)
g(I)

err(I)
g(I)
.
We now choose g(I) to be such that err(I) = err(I) so that the new err(I) becomes
176
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
zero. Therefore,
g(I) =
err(I)
err(I)
g(I)
. (8.9)
Now we have an algorithm that starts with an initial guess g
0
(I) of g(I) for all halftone
dots I and then iterates to generate new estimate g
k
(I) of g(I):
1. estimate the halftone lattice from h
s
(J);
2. low-pass lter h
s
(J) to obtain a defocused input image, and algorithmically halftone
the defocused image, using parameters from Step 1, to obtain g
0
(I); then repeat steps
2 and 3 for k = 0, 1, . . ., until g
k
(I)s converge;
3. for each halftone dot I compute err(I),
err(I)
g(I)
, and g(I) using Eqs. 8.7, 8.8, and 8.9;
4. for each halftone dot I update g
k
(I) with
g
k+1
(I) = g
k
(I) + g(I).
Note that to compute err(I) and
err(I)
g(I)
, the sampling of g
k
(I) is performed.
We have not implemented the relaxation method. Many details of the method still
remain to be worked out, and the eectiveness or even convergence of the method is yet
to be veried. We may foresee at least one issue that may complicate the method. err(I)
in Eq. 8.7 is the sum of errors of the estimated pixel values, and the errors may cancel
one another. A remedy is to change err(I) to the sum of squared errors. However, the
computational complexity for evaluating
err(I)
g(I)
will be dramatically increased.
8.6 Discussion
A moire pattern in a sampled halftone usually consists of an aliasing pattern and a non-
aliasing pattern. The aliasing pattern may consist of many components. In the frequency
domain, these components are from the replicas that are adjacent to the zeroth replica.
The non-aliasing pattern is usually caused by the rst order harmonics of the halftone
frequencies of the zeroth replica and its adjacent replicas.
177
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
The non-aliasing pattern is a contrast pattern and can be easily eliminated by low-pass
ltering without obvious loss of image information because a sampled halftone image does
not carry signicant image information at frequencies as high as the halftone frequency.
The low-pass ltering removes the halftone dot structure from the image, increasing the
adjacent pixel correlation and thus making the image easier to compress. Also, resizing
the image will not produce a moire pattern. Sometimes the aliasing pattern in a sampled
halftone is much weaker than the non-aliasing pattern, and in such cases, the removal of
the non-aliasing pattern signicantly improves the appearance of the image.
It is still relatively easy to remove those components of an aliasing pattern whose fre-
quencies are higher than the frequencies of the majority components of the image content.
Such moire components can be removed by notch ltering without signicant loss of image
information.
It is much more dicult to suppress the low frequency components of an aliasing moire.
Very often these components form an intensity pattern. A complete removal of the energy
at these low frequencies from the spectrum of the sampled image is not acceptable because
the image content has a signicant amount of energy at these frequencies. Both the simu-
lation method and the relaxation method have the potential to suppress the low frequency
components of an aliasing moire, but more work is needed to make the methods practical.
One diculty in suppressing a low frequency intensity pattern is that the pattern am-
plitude is dependent upon the density of the image features. Another possible approach to
overcome this problem is a computer assisted interactive procedure wherein the computer
estimates the frequency and phase of the intensity pattern and relies on the the human
operator to adjust the local amplitude of the pattern interactively.
The three moire reduction techniques proposed in this chapter have dierent charac-
teristics. The notch ltering method is a frequency domain technique. It is simple and
intuitive in logic and requires only the estimation of the halftone frequency. The method is
relatively eective if all the moire components have high frequencies. Both the simulation
method and the relaxation method, on the other hand, are primarily spatial domain tech-
178
CHAPTER 8. RESTORATION OF SAMPLED HALFTONES
niques. They attempt to provide consistent reduction of moire regardless of the distribution
of moire components in the frequency plane. A sampled halftone can be considered to be
the sum of a moire pattern and the image content. The simulation method attempts to
estimate the moire pattern while the relaxation method attempts to estimate the image
content. Both methods are much more complex than notch ltering. They are also more
dependent upon the knowledge of the sampling model and require a very accurate estimate
of the halftone lattice, since a small change in either the sampling lattice or the halftone dot
lattice causes a large change in the appearance of the moire. The simulation method and
the relaxation method share the same core techniques for modeling the halftoning process
and the sampling process that generate an input sampled halftone.
Dierent non-uniform sampling models produce vastly dierent moire patterns. Sys-
tematic moire reduction requires knowledge of the sampling model. Heuristic analysis of
the Fourier spectrum of a sampled halftone image to identify components of a moire pattern
is complicated and may not be conclusive.
A good quantitative moire reduction measure must be able handle the ambiguity caused
by the loss of image information due to an aliasing moire pattern. It must also correlate
with human visual perception.
179
Chapter 9
Prevention of Moire Patterns When Sampling a
Halftone
9.1 Introduction
An aliasing moire pattern in a sampled halftone image is due to insucient low-pass
ltering prior to sampling. If we can perform proper low-pass ltering, we can prevent
aliasing moire altogether. For a xed sampling rate, a xed anti-aliasing lter is easy to
implement. It is more dicult to implement a variable lter or a bank of lters for a variable
sampling rate.
For the ScanJet IIc, the CCD array samples an input picture at a xed rate. The scan-
ner optics are designed to act as an anti-aliasing lter for that rate. The most signicant
aliasing in an output image is due to insucient low-pass ltering and non-uniform subsam-
pling performed by the scanner circuits to reduce the resolution of the optoelectronically
sampled data. In this chapter we apply a well-known resolution reduction method to the
optoelectronically sampled data directly to produce an output image of an arbitrary lower
resolution. The output image is still free of aliasing.
9.2 Partial inverse Fourier transform
The partial inverse Fourier transform method converts an image of M N pixels to a
lower resolution image of mn pixels, m M, n N. It has two steps:
1. compute the M N point DFT of the input image; let the DFT be G[k, l], where
k =
N
2
|,
N
2
| + 1, . . . ,
N
2
| 1 and l =
M
2
|,
M
2
| +1, . . . ,
M
2
| 1;
180
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
2. compute the mn point inverse DFTof the center portion of G[k, l]: k =
n
2
|,
n
2
|+
1, . . . ,
n
2
| 1 and l =
m
2
|,
m
2
| + 1, . . . ,
m
2
| 1; the result is the desired mn
output image.
By inverse transforming only the center portion of the Fourier transform of the input image,
a uniform subsampling preceded by an ideal low-pass ltering is implicitly performed on the
input image. No aliasing occurs in the output image. The spectrum of the output image is
simply the selected portion of the input image spectrum replicated in both horizontal and
vertical directions.
Assume that the input image to the above algorithm is a sampled halftone that has
virtually no aliasing, as is the case with a halftone image scanned at 400 dpi by a ScanJet
IIc. Figures 9.1(a) and (b) depict, respectively, the distribution of signicant frequency
components of the input and output images. In Figure 9.1(a) the dashed square shows
the portion of the input image spectrum selected for the inverse transform, and the dashed
square in Figure 9.1(b) shows the Nyquist band of the output image, which is the same
band in which the selected portion of the input image spectrum resides. The dotted lines
in the gure show to which replica each component belongs. The reader may compare
Figure 9.1 with Figure 6.8. Also realize that the input image is a sampled image itself, and
its spectrum consists of multiple replicas. Figure 9.1(a) shows only a portion of the zeroth
order replica larger than the selected band.
It is clear in Figure 9.1(b) that there is no overlap of replicas due to the implicit subsam-
pling. However, a halftone frequency harmonic of one replica may still be close to halftone
frequency harmonics of adjacent replicas, and the interference among these harmonics forms
a contrast pattern in the output image. The dashed circle in Figure 9.1(b) shows that for
the particular output image size the rst order halftone frequency harmonics of adjacent
replicas are close to one another. To overcome this problem, we can lter the selected por-
tion of the DFT of the input image before the inverse transform to smooth the components
at halftone frequency harmonics near the band border. Equivalently, we can perform the
ltering on the inverse transformed image, though it is computationally more expensive. In
181
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
(b)
(a)
Figure 9.1: Distribution of signicant frequency components of (a) a halftone scanned
uniformly with sucient low-pass ltering and (b) an output image from the partial inverse
Fourier transform method taking the scanned halftone as input
182
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
Figure 9.2: The 342 342 pixel center portion of the DFT of the 400 dpi, 10241024 pixel
master image used in Chapter 8
the following examples simple low-pass ltering or notch ltering are used.
Now we apply the partial inverse Fourier transform method to the 400 dpi, 10241024
pixel master image used in Chapter 8, whose DFT spectrum is shown in Figure 5.10. We
make two sample output images, one at 133
1
3
dpi, the other at 150 dpi. Figure 9.2 shows the
342342 pixel center portion of the DFT of the master image, corresponding to a 133
1
3
dpi
spectral band. Figure 9.3 shows the inverse transformed image. The moire pattern in the
image in Figure 9.3 is a non-aliasing contrast pattern and is caused by the four rst order
harmonics of the halftone frequency at the corners of the (portion of) DFT in Figure 9.2.
Figure 9.4 shows the image in Figure 9.3 ideally low-pass ltered to eliminate the rst order
halftone frequency harmonics. Figure 9.5 shows the inverse transform of the notch ltered
384 384 pixel center portion DFT of the master image. The notch lter eliminates the
rst order halftone frequency harmonics.
183
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
Figure 9.3: The 133
1
3
dpi output image obtained by inverse transforming the DFT in
Figure 9.2
184
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
Figure 9.4: The 133
1
3
dpi output image obtained by low-pass ltering the image in Figure 9.3
The partial inverse Fourier transform method is conceptually simple and produces im-
ages of good quality. However, the general DFT, the building block of the algorithm, is
complex in logic and is expensive to implement in hardware. The method can be imple-
mented in software as a backup procedure to the digital resolution reduction procedure of
a scanner. The method can also be implemented to reduce the resolution in only one direc-
tion, e.g. the rows, of an image. It should be noted that it is always possible to compute
the DFT of an N N image in O(N
2
log N) time for an arbitrary N [Nus82].
Inverse transforming the center portion of the DFT of an image is equivalent to con-
volving the image with a lter of the form sinc(
x
x)sinc(
y
y ) and then subsampling the
ltered image. Here
x
and
y
are constants dependent upon the size of the output image.
The inverse Fourier transform method can therefore be implemented in the spatial domain.
185
CHAPTER 9. PREVENTION OF MOIRE PATTERNS
Figure 9.5: The 150 dpi output image obtained by low-pass ltering the 384 384 pixel
center portion of the master image DFT followed by inverse transform
186
Chapter 10
Concluding Remarks
10.1 The nature of a moire pattern
In Chapters 5 and 6 we showed that a moire pattern in a sampled halftone image is
usually complex, even though it may appear visually as a simple periodic pattern. It may
have many contributing components, and can be considered to be the superposition of many
periodic patterns. For a practical scanner, the blurring of the optics reduces the number of
signicant moire components, but a non-uniform sampling procedure may create new moire
components.
A typical moire pattern can be decomposed into an aliasing moire pattern and a non-
aliasing moire pattern. A typical moire pattern can also be considered to be composed of a
density pattern and a contrast pattern. The density pattern and the contrast pattern have
the same low visual frequency and have either the same phase or the opposite phase. A
density pattern usually has lower energy than a contrast pattern, but a density pattern is
also more easily observed. An aliasing moire pattern consists of a density pattern and a
contrast pattern. A non-aliasing moire pattern is always a contrast pattern.
It is relatively easy to remove a non-aliasing moire without loss of image information.
However, it is theoretically impossible to remove an aliasing moire from the image while
keeping the image information intact. We can only seek approximation methods to reduce
an aliasing moire.
Compared with restoration of a sampled halftone image that already has an aliasing
moire, prevention of an aliasing moire when a halftone picture is sampled is much easier.
The spatial frequency and orientation of a moire pattern is dependent upon, and is
187
CHAPTER 10. CONCLUDING REMARKS
very sensitive to, the halftone dot lattice of the input halftone picture, the sampling point
lattice of the sampler, and the relative position between the two lattices. A halftone picture
typically has a uniform dot lattice, but there are many variations in the sampling models
of dierent scanners. To identify the composition of a moire pattern, knowledge of the
sampling model is necessary.
10.2 Future work
Even though we now have a solid understanding of a moire pattern and have proposed
some techniques to suppress it, there is still much to be done to improve these techniques. It
is also desirable to develop new techniques, and the techniques proposed in this dissertation
provide references against which the new techniques can be judged.
The simulation method developed in Section 8.4 reduces the signicance of a moire
pattern in a sampled halftone. A key to the practical utility of the method appears to lie
in the accuracy of the model of the halftoning process that generates the input halftone
picture.
The relaxation method described in Section 8.5 opens up a new spatial domain approach
to restore an image that has a moire pattern. Its eectiveness is yet to be assessed.
The moire reduction quality metrics developed in Chapter 8 are a rst attempt to
quantitatively measure the signicance of residue moire pattern in a restored image in
conjunction with the preservation of image information. The metrics may be improved to
yield consistent values across images with dierent sizes and contents, or sampled dierently.
The relationship of the metrics with the human visual perception is also to be established.
The resolution reduction method described in Chapter 9 eectively prevents moire pat-
terns, but it requires much more complex computation than the improved binary rate mul-
tiplication method discussed in Chapter 6. An immediate question is whether new methods
can be developed that both prevent moire patterns and are computationally simple and
suitable for hardware implementation.
188
CHAPTER 10. CONCLUDING REMARKS
An interesting question is how to make the connection between the moire patterns in
moire metrology (Chapter 2) and the moire patterns in sampled halftones, so that knowl-
edge from moire metrology can be used to analyze or suppress moire patterns in sampled
halftones.
It is also of practical value to investigate moire patterns in sampled color halftone images.
A color halftone picture usually consists of four halftone dot lattices oriented at dierent
angles, each lattice in one primary color or in black. Experience shows that only those
lattices whose colors have high contrast to the paper on which the picture is printed can
cause noticeable moire patterns. The black dot lattice is most likely to cause a moire. While
a moire pattern can be analyzed in each color, concentration should be placed on the black
lattice.
189
REFERENCES
[AA92] M. Analoui and J. Allebach. New results on reconstruction of continuous-tone
from halftone. In IEEE ICASSP, pages III313316, San Francisco, March 1992.
[AL77] J.P. Allebach and B. Liu. Analysis of halftone dot prole and aliasing in the
discrete binary representation of images. Journal of Optical Society of America,
67:11471154, 1977.
[All78] J.P. Allebach. Random nucleated halftone screen. Ph otograph ic Science and
Engineering, 22:8991, 1978.
[BBM
+
92] S. Boyd, S. Bencuya, D. McGrath, S. Saylor, and W. Smyth. Photographic
scanner with integrated charge-coupled device. In Proceedings of SPIE, volume
1656, pages 270280, 1992.
[Bel84] M. Bellanger. Digital Processing of Signals, chapter 2. John Wiley & Sons,
1984.
[Bra86] R.N. Bracewell. Th e Fourier Transform and Its Applications. McGraw-Hill,
1986.
[Bri74] E.O. Brigham. Th e Fast Fourier Transform. Prentice-Hall, 1974.
[Bry74] O. Bryngdahl. Moire: Formation and interpretation. Journal of Optical Society
of America, 64:12871294, October 1974.
[CR83] R.E. Crochiere and L.R. Rabiner. Multirate Digital Signal Processing. Prentice-
Hall, 1983.
[Cuc52] C.L. Cuccia. Sidebands and Transients in Communication Engineering, pages
287289. McGraw-Hill, 1952.
[DB93] G. Degi and D.C. Buck. Image processor. U.S. Patent 5185817, 1993.
[Deg95] G. Degi. Personal communication. Author at Howlett-Packard Company, Gree-
ley, CO, 1995.
[FJ94] S. Forchhammer and K.S. Jensen. Data compression of scanned halftone images.
IEEE Transactions on Communications, 42(2/3/4):18811893, 1994.
[Flo92] O. Florant. Technical choices for a very high denition dia scanner. In Proceed-
ings of SPIE, volume 1656, pages 207211, 1992.
190
REFERENCES
[FS75] R.W. Floyd and L. Steinberg. Adaptive algorithm for spatial grey scale. SID
Int. Sym. Digest of Tech . Papers, pages 3637, 1975.
[Gor71] W.J. Gordon. Blending-function methods of bivariate and multivariate interpo-
lation and approximation. SIAM Journal of Numerical Analysis, pages 158177,
August 1971.
[Hew93] Hewlett-Packard Company. HP DeskScan II Users Guide, Apple Macintosh
Version. Hewlett-Packard Company, November 1993.
[Hua74] T.S. Huang. Digital transmission of halftone pictures. Computer Graph ics and
Image Processing, 3:195202, 1974.
[JJN76] J.F. Jarvis, C.N. Judice, and W.H. Ninke. A survey of techniques for the display
of continuous tone pictures on bilevel displays. Computer Graph ics and Image
Processing, 5:1340, 1976.
[KR75] D. Kermisch and P.G. Roetling. Fourier spectrum of halftone images. Journal
of Optical Society of America, 65:716723, 1975.
[Kra92] Kai Krause. Photoshop: Kais power tips & tricks. From ftp.netcom.com in
/pub/hsc/Kais Power Tips, also oating on the internet, author at kai@aol.com,
1992.
[Mal92] D. Malacara, editor. Optical Sh op Testing, chapter 1 and 16. John Wiley &
Sons, 1992.
[MB93] Mercedes-Benz. Personal communication, 1993.
[MP91] T. Mitsa and K.J. Parker. Digital halftoning using a blue noise mask. In
Proceedings of SPIE, volume 1452, pages 4756, 1991.
[MP92] C.M. Miceli and K.J. Parker. Inverse halftoning. Journal of Electronic Imaging,
1:143151, 1992.
[MS74] J.L. Mannos and D.J. Sakrison. The eects of a visual delity criterion on the
encoding of images. IEEE Transactions on Information Th eory, 20:525536,
July 1974.
[MSO90] Y. Morimoto, Y. Seguchi, and M. Okada. Screening and moire suppression in
printing and its analysis by fourier transform. Systems and Computers in Japan,
21:98105, 1990.
[Nus82] H.J. Nussbaumer. Fast Fourier Transform and Convolution Algorith ms, chap-
ter 5. Springer-Verlag, 1982.
191
REFERENCES
[OS89] A. V. Oppenheim and R.W. Schafer. Discrete-Time Signal Processing, pages
109111. Prentice-Hall, 1989.
[OYT
+
86] N. Ohyama, M. Yamaguchi, J. Tsujiuchi, T. Honda, and S. Hiratsuka. Suppres-
sion of moire fringes due to sampling of screened images. Optics Communica-
tions, 60:364368, 1986.
[Pel94] Adar Pelah. Inverting the perceptual transform. In Proceedings of SPIE, volume
2179, pages 29, San Jose, February 1994.
[PN92] T.N. Pappas and D.L. Neuho. Least-squares model-based halftoning. In Pro-
ceedings of SPIE, volume 1666, pages 165176, San Jose, CA, February 1992.
[RL94] P.G. Roetling and R.P. Loce. Digital halftoning. In E. Dougherty, editor, Digital
Image Processing Meth ods, chapter 10. Marcel Dekker, 1994.
[Roe76] P.G. Roetling. Halftone method with enhancement and moire suppression. Jour-
nal of Optical Society of America, 66:985989, 1976.
[Roe77] P.G. Roetling. Analysis of detail and spurious signals in halftone images. Journal
of Applied Ph otograph ic Engineering, 3:1217, 1977.
[Ros69] A. Rosenfeld. Picture Processing by Computer, page 1. Academic Press, 1969.
[RR93] D.W. Robinson and G.T. Reid, editors. Interferogram Analysis, chapter 2. In-
stitute of Physics Publishing, Britol and Philadelphia, 1993.
[Sch89] R.J. Schalko. Digital Image Processing and Computer Vision, page 56. John
Wiley & Sons, 1989.
[SM81] J.C. Stoel and J.F. Moreland. A survey of electronic techniques for pictorial
image reproduction. IEEE Transactions on Commumnications, 29:18981925,
1981.
[SSY89] J.S. Shu, R. Springer, and C.L. Yeh. Moire factors and visibility in scanned and
printed halftone images. Optical Engineering, 28:805812, 1989.
[Sto80] J.C. Stoel. Automatic multimode (continuous tone, halftone, linecopy) repro-
duction. U.S. Patent 4194221, 1980.
[SW82] A. Steinbach and K.Y. Wong. Moire patterns in scanned halftone pictures.
Journal of Optical Society of America, 72:11901198, 1982.
[Tex83] Texas Instruments Incorporated. Optoelectronics Data Book, 1983-1984. Texas
Instruments Incorporated, 1983.
[Uli87] R. Ulichney. Digital Halftoning. MIT Press, 1987.
192
REFERENCES
[Vai93] P.P. Vaidyanathan. Multirate Systems and Filter Banks. Prentice-Hall, 1993.
[Won77] K.Y. Wong. Suppression of moire patterns by error feedback. IBM Tech nical
Disclosure Bulletin, 29:19511952, 1977.
[YOH
+
89] M. Yamaguchi, N. Ohyama, T. Honda, J. Tsujiuchi, and S. Hiratsuka. A color
image sampling method with suppression of moire fringes. Optics Communica-
tions, 69:349352, 1989.
193
VITA
Xiangdong Liu was born on May 5, 1963 in Beijing, China. He received a bachelars
degree in computer science from Fudan University, Shanghai, China in 1983 and received a
masters degree in computer science from Shanghai Teachers University, Shanghai, China
in 1987. He also received a masters degree in applied mathematical sciences from the
University of Georgia, Athens, Georgia, USA, in 1991. He received a Ph.D. degree in
computer science from Virginia Polytechnic Institute and State University (Virginia Tech),
Blacksburg, Virginia, USA, in 1996. He worked as a teaching assistant at Fudan University
from 1983 to 1985 and as an instructor at Shanghai Teachers University from 1987 to 1989.
194

También podría gustarte