Download as pdf or txt
Download as pdf or txt
You are on page 1of 58

756 Vol. 7, No.

4 / December 2015 / Advances in Optics and Photonics Review

Compressive tomography
DAVID J. BRADY,* ALEX MROZACK, KEN MACCABE, AND
PATRICK LLULL
Department of Electrical and Computer Engineering, Box 90291, Duke University, Durham, North Carolina
27708, USA
*Corresponding author: david.brady@duke.edu

Received May 29, 2015; revised September 17, 2015; accepted September 22, 2015; published
November 19, 2015 (Doc. ID 241997)

Compressive tomography consists of estimation of high-dimensional objects from lower


dimensional measurements. We review compressive tomography using radiation fields,
including geometric, wave, and statistical field models. We find coded apertures most
appropriate for compressive coding with geometric models, spatial and frequency sub-
sampling most appropriate for wave models, and temporal modulation most appropriate
for statistical models. In each case, we show that compressive measurement enables
snapshot 3D imaging, eliminating or reducing the need to sacrifice temporal resolution
in multidimensional imaging. © 2015 Optical Society of America
OCIS codes: (110.1758) Computational imaging; (110.6880) Three-dimensional
image acquisition; (110.6955) Tomographic imaging; (110.7348) Wavefront
encoding
http://dx.doi.org/10.1364/AOP.7.000756

1. Background. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 757
2. Models and Strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 761
3. Geometric Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 763
4. Diffraction Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 778
4.1. Compressive Holography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 780
4.2. Frequency Scanned Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785
4.3. Dispersive Imagers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 789
5. Focal Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 792
5.1. Depth-Encoding Focal Sweeps. . . . . . . . . . . . . . . . . . . . . . . . . . . 797
5.2. Snapshot Focal Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . 801
6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 805
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 805
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 805
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 757

Compressive tomography
DAVID J. BRADY, ALEX MROZACK, KEN MACCABE, AND
PATRICK LLULL

1. BACKGROUND
Compressive sampling (CS) is the estimation of N signal values from M discrete mea-
surements, where M < N . Computed tomography (CT) is the estimation of a
D-dimensional signal from a set of d-dimensional measurements, where d < D.
Both CS and CT implement dimensionality transformations on measurements to es-
timate images [1]. The “dimension” of an image may be defined in two ways. First, the
sample dimension is the number of different data values needed to describe an image.
Under this definition, a megapixel image is a 106 -D object. Second, the object dimen-
sion is the dimension of the physical space over which the image is distributed. The
image has a well-defined value at every point in this space. Under this definition, a
megapixel image is 2, 3, or N-dimensional, matched to the dimension of its underlying
object space.
CS is a methodology for minimizing the sample dimension of an image. CT, in
contrast, focuses on transformation of the object dimension. Historically, CT relies
on measurements distributed over a second physical space. The measurement dimen-
sion of this second space may correspond, for example, to the 2D plane of a detector
array used to reconstruct a 3D image in cone beam x-ray CT. From this perspective,
CT is a methodology for combining a sequence of images distributed over a D − 1-
dimensional object space to produce an image distributed over D-dimensional object
space. With the addition of time, in the form of a sequence, the total physical dimen-
sion of the CT measurement dimension typically matches the object dimension. For
example, 3D bodies are imaged in x-ray cone beam tomography by combining se-
quences of 2D x-ray projection imagines. Over the past decade, however, various
strategies have emerged that reduce or eliminate the need to use a temporal sequence
in CT. Using these strategies, one may estimate a 3D object from a single 2D snapshot.
Compressive tomography is the estimation of D-dimensional objects from measure-
ments distributed over a d < D-dimensional space.
This paper reviews physical coding and measurement strategies for compressive
tomography. “Compressive” in this context implies both that the number of measure-
ments is less than the number of object voxels estimated and that the measurements
are embedded in a space of lower dimension than the object space. Compressive
tomography lies at the intersection of digital signal processing and physical system
design. The goal of this review is to enable the signal processing specialist to under-
stand physical constraints in implementation of CS kernels and to enable the physical
system designer to understand strategies for compressive coding in tomographic in-
struments. We review recent demonstrations of compressive tomography in the three
main radiation field models and we discuss opportunities and challenges for future
development.
Concepts of compressive sampling and tomography begin with an understanding
of noncompressive systems. Shannon proved that the sampling dimension of a band-
limited signal is upper-bounded by the space–bandwidth product S  X B, where X is
the spatial extent of the signal and B is the maximum spatial frequency [2]. Shannon’s
sampling theorem provides an interpolation algorithm that exactly recovers the sam-
pled function at every point in continuous space. CS extends Shannon sampling in
758 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

showing that the sampling dimension of a signal described by additional prior con-
straints may be less than the Shannon value. With this in mind, we define CS in this
paper to be “estimation of images with sub-Shannon sampling dimension.” In making
this definition, we recognize that sample models for continuous signals may play a
role in system performance and that aperiodic sampling [3,4] in particular may prove
advantageous. We also recognize that CS of continuous signals may involve more
sophisticated priors than band limits or sparsity.
While there is some subtlety in the definition of sampling dimension, object and mea-
surement dimension are usually physically obvious quantities. Most often, objects are
three dimensional, and measurements are distributed on 2D boundaries. Tomographic
systems measure physical properties of the object, such as density or scatter cross
section, distributed over the embedding dimension by sampling radiation fields dis-
tributed over the measurement dimension. Noncompressive systems assume that the
measurement dimension must match the object dimension. This match is achieved
by adding a temporal dimension to the measurements and by operating under the
assumption that the object does not vary over the measurement time. The most sig-
nificant contribution of compressive tomography to date has been to challenge the idea
that one must sacrifice temporal resolution to acquire tomographic images [5,6]. We
define compressive tomography in this paper to be “estimation of images from mea-
surements embedded in a space of lower dimensionality than the object space.” Note
that this definition allows continuous as well as discrete measurement and object dis-
tributions. We also include the possibility that a compressive tomography system does
not use CS (e.g., the sampling dimension of the measurements may exceed the number
of image values estimated).
Sampling theory has evolved quickly over the past quarter-century with the emergence
of multiscale [7] and data-dependent [8] approaches to discrete signal representation.
Beginning with early multiband sampling strategies [9], these approaches demonstrate
that dimensionality need not be preserved in signal representation. Wavelet theory
expands the object dimension of images using time-frequency representations. In con-
ventional sampling theory, object dimension translates into the dimension of data in-
dices. The increased dimensionality of a time-frequency representation translates into
an increased index dimension in wavelet sampling based on position in the continuous
object space and scale in a hierarchy of variable resolution spaces. This approach
enables reduction in sample dimension by dynamically or algorithmically assigning
multiscale resolution to regions of interest. These approaches are used in diverse im-
age compression algorithms and standards [7].
Whereas sampling theory focuses on discrete mathematical representation of con-
tinuous signals, “measurement theory” focuses on optimal strategies for taking data.
Canonical measurement challenges include group testing and weighing design. Group
testing seeks to identify unusual signal components using a minimum number of mea-
surements [10]. Weighing design seeks to maximize the accuracy with which signal
components can be estimated in the presence of noise [11]. Measurements in each case
consist of weighted sums of discrete signal components. The design challenge con-
sists of selecting weights within constraints to optimize system performance.
In contrast with sampling theory, measurement theory does not assume that signal
components are embedded in a continuous object space. Rather, the signal is repre-
sented as a discrete set of N variables, each corresponding to an explicitly defined
component. The object dimension is N . Group tests of N components with just a
single variant value typically reduce the sample dimension from the naïve value of
N to Olog N . More generally, the sample dimension of signal with K variant signal
values is less than or equal to K log N . It is not necessarily the case, however, that a
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 759

signal with K variant components can be uniquely specified from just K log N mea-
surements. Historically, most studies focused on “combinatorial group testing,” which
enables deterministic signal retrieval under strong prior constraints. More recently,
however, considerable attention has focused on “probabilistic group testing,” which
assumes that the measurement dimension matches the number of samples necessary to
describe the signal with high probability rather than with certainty. Seminal reports by
Donoho [12] and Candes and co-workers [13,14], showing that the measurement di-
mension is proportional to the sample dimension limit when K ≪ N , triggered a
multitude of theoretical and experimental studies.
Measurement theory for imaging and tomography has historically focused on noise
reduction rather than signal dimension compression. Multiplex measurement has long
been utilized in imaging and spectroscopy to improve photon efficiency and
increase the signal-to-noise ratio (SNR) in the presence of additive noise [1,11].
Multiplexing consists of deliberately measuring weighted sums of image pixel values
rather than directly sampling image data. Examples include Fourier transform spectros-
copy [15] and imaging [16] and coded-aperture spectroscopy [1,11] and imaging [17].
Multiplexing is used in tomographic imaging by necessity rather than by choice. In
most cases, it is physically impossible to independently measure object points distrib-
uted over a volume. Tomography captures the state of fields influenced by multiple
points in the volume. Tomographic multiplexing has generally focused on simple and
naturally implemented Radon or Fourier transforms. As more advanced signal analy-
sis and inversion strategies have emerged over the past quarter-century, system design-
ers began to consider more abstract multiplexing strategies. An early example arose in
the deliberate aberrations Dowski and Cathey introduced to optical imaging systems
to improve depth of field [18]. Our group utilized the Dowski and Cathey approach
[19], as well as other multiplexing strategies [20], to demonstrate 3D optical projec-
tion tomography. In analyzing these multiplexing strategies, one must, of course, also
consider physical plausibility; in particular, thermodynamic requirements for conser-
vation of energy and entropy must be satisfied [21].
Physical constraints enter more directly in the structure of measurement space. While
the object space for a tomographic system necessarily corresponds to the physical
distribution of the object, the measurement space may be selected (to some extent)
by the system designer. In the case of the single pixel camera [22], for example, each
measurement consists of an independent linear projection of every object point. Since
there is no natural order to these projections, the dimension of the measurement em-
bedding space is equal to the number of measurements. The lack of a natural order
implies independence between measurements consistent with maximal data compres-
sion. Measurements with an order parameter, as in Fourier transform or circulant cod-
ing, may tend to be less compressive. Independence from one measurement to the next
is also common in coded-aperture spectroscopy.
Most imaging systems distribute measurement resources in parallel over a continuous
measurement space, typically a 1D or 2D detector array. Measurements may also be
ordered by continuous temporal variation of a physical sampling parameter. In con-
ventional tomographic systems, the measurement space is homeomorphic to the object
space. When reconstructing a 2D object for example, one assumes the 2D Radon or
Fourier transform is captured, meaning that both the object and measurement space
correspond to R2 .
Compressive sampling strategies proposed in [12–14] did not include concepts of
object or measurement dimension. The success of these strategies in compressive im-
aging is thus surprising, given that the embedding dimension and implicit correlation
760 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

between adjacent pixels is the most significant constraint on image data. This ap-
proach introduces object dimension through basis selection. Approaching the problem
from a geometric perspective, Brady et al. proposed diverse multiplexing strategies
for feature specific and compressive measurement in [23]. Our approach built on the
feature-specific measurement strategy proposed by Neifeld and Shankar [24]. We
were also motivated by geometric coding strategies using reference structures to ran-
domize mappings between measurement and object spaces [25]. Reference structures
effectively disorder measurement space data, thereby eliminating the need for dimen-
sion conservation in tomography.
More recently, strategies to account for the spatial structure of image data include
Bayesian [26] and model-based systems [27]. As the theoretical and physical plat-
forms for compressive measurement have developed, data models specific to signals
embedded in continuous object spaces have become increasingly important. Current
approaches include Gaussian mixture [28], union-of-subspace [29], and manifold
models [30]. Models to combine CS sparsity models with learned or constrained im-
age properties to enable decompressive inference without prior knowledge of the
underlying sparse basis are also emerging [31]. One anticipates that novel models
may be proposed for some time as increasingly image-and task-specific data structures
are developed.
Multiplex measurement often produces drastic differences between the topologies of
object and measurement space. The design of measurement space for tomographic
imagers is a central theme of this review. In optimizing both mathematical condition-
ing and physical implementation, the complete lack of a measurement embedding
space is undesirable. On the other hand, discontinuous transformations increase cod-
ing power and compression. In balancing these constraints, measurements are likely to
be structured with compact localized support in object and measurement space but
with some degree of independence over that support.
This review focuses on tomographic imaging using radiation fields. Radiation may be
modeled using geometrical optics, wave optics, or statistical optics [1]. The geometric
field is described by the radiance, which propagates using ray projection. Waves are
described by the electromagnetic field and propagate according to the Maxwell equa-
tions. The statistical field is described by coherence functions, which propagate ac-
cording to multidimensional wave equations derived from the Maxwell equations. In
subsequent sections, we review physical system design strategies for compressive
tomography based on each of these models.
In focusing on radiation fields we do not consider magnetic resonance imaging, which
has been an important platform in the development of CS [32]. Indeed, adapted wave-
form encoding for compressive MRI [33] is a primary antecedent to current CS mod-
els. Antecedents may also be found in other branches of tomography, with particular
emphasis on the use of convex optimization and other strategies to recover missing
data [34]. Healy’s MRI studies are of particular significance, however, as early com-
binations of multiscale bases and deliberate system design to optimize sampling ef-
ficiency. We also neglect a fourth radiation field model in not considering diffusion
tomography, which has received attention elsewhere from the perspective of CS
theory [35,36].
This review does not focus on the inverse problem of image estimation from coded
data. Compressive measurement is an integrated sensing and processing design chal-
lenge requiring outstanding signal processing and physical system design. One purpose
of this review is to show that the basic strategies of physical design can be independent
of the inversion algorithm. Successful construction of compressive tomography systems
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 761

ultimately requires (1) well-understood and expressed prior knowledge about the ob-
ject, (2) accurate physical forward models for measurement systems targeting feature-
specific object information, and (3) inversion algorithms capable of combining prior
and measured information to accurately estimate the object. These three components
require state-of-the-art machine learning, physical design, and signal processing, ul-
timately including integrated design of all three components. The goal of this review is
to describe the tool set and low-level strategy for the physical design component.
It is important to emphasize that physical layer compression itself is not necessarily a
useful metric in analysis of measurement systems. The designer is more likely to value
such traditional metrics as system size, weight, and power, as well as cost, resolution,
SNR, and image or feature fidelity. Compressive sampling is best understood as a tool
for improving system performance against these metrics. In considering the utility of
CS, we find it useful to abstractly consider strategies for compressive tomography in
Section 2. Successive sections consider specific applications to ray, wave, and coher-
ence field models.

2. MODELS AND STRATEGIES


Compressed sensing theory typically begins with a linear forward model between a
discrete array of object features f ∈ RN and a discrete array of measurements g ∈ RM
such that
g  Hf  n; (1)

where n ∈ RM represents noise. The measurement system is said to be compressive


if M < N .
While this definition has been highly useful and successful in many sensing and im-
aging applications, it is helpful to reach a bit further back into the physics of the mea-
surement system and consider details of the transformation from continuous object
and field distributions to discrete representations. Tomography most commonly con-
sists of imaging 3D objects from measurements distributed over 1D or 2D sensor arrays.
Typical tomographic systems may be described by integral equations of the form
Z
gy  f xhx; ydx  ny; (2)

where x ∈ RN and y ∈ RM . We define compressive tomography as the estimation of


f x from gy in the case that M < N .
Tomographic systems typically use sensor arrays embedded on the boundary or sur-
face of a volume under observation. In fan beam tomography, for example, a linear
detector array measures the attenuation of rays through a 2D object space. In cone
beam tomography, a planar detector array measures rays projected through a 3D vol-
ume. Conventional tomography overcomes the dimensional mismatch between the
object and measurement spaces by varying illumination and sensor geometry as a
function of time, thereby increasing the dimension of the measurement space by
1. Thus, in conventional systems, M  N − 1 for measurements taken at a fixed time,
but M  N when time is taken into account.
The most unfortunate aspect of the conventional approach is that it requires that the
object remain static as measurements are collected over time. Over the past several
years, our group has applied CS theory to implement snapshot compressive tomog-
raphy. For example, we have shown that 3D hyperspectral [5], diffraction [6], and
x-ray scatter [37] images may be reconstructed from 2D data. We have also analyzed
CS for reconstruction of 3D objects with conventional optics [38]. Most recently, we
762 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

have shown that 3D video data cubes may be constructed from 2D frames [39], thus
using compressive tomography to reconstruct time itself.
Complete elimination of temporal sampling is often overly compressive in tomo-
graphic systems. Multiframe [40] and adaptive [41] data acquisition may be used
to adjust compression levels to match object characteristics. With such systems in
mind, one may consider compressive tomography as a generalized strategy for bal-
ancing trade-offs between continuous embedding dimensions and discrete sampling
dimensions, including the possibility of fractal and multiscale approaches to dimen-
sional transformations.
While the object distribution f x is by definition distributed over continuous
space, measurements ultimately consist of discrete digital data. There is no fundamen-
tal requirement that discrete measurements be indexed by a continuous variable.
Standard CS models assume independent kernels for each measurement.
Unfortunately, completely independent kernels are difficult or impossible to imple-
ment on measurements embedded in continuous physical space. Due in part to this
challenge, Candes’ early analysis of compressive tomography focused on discrete
subsampling of the temporal portion of Radon space with continuous sampling in
each snapshot.
Our group’s initial theoretical studies of compressive tomography focused on the use
of multidimensional reference structures to enable random or decorrelated measure-
ment over a continuous space [25]. However, most subsequent efforts to implement
practical compressive tomography may be described in the context of the following
three basic coding strategies.
1. Measurement space coding. The standard model for increasing measurement di-
mensionality with time involves varying the measurement kernel to obtain
Z
gy; t  f xhx; y; tdx: (3)

Measurement space coding multiplexes diverse kernels in a snapshot to obtain


Z Z
g̃y  gy; tCy; tdt  Cy; tf xhx; y; tdxdt; (4)

where Cy; t is a code applied to each time slice in measurement space. Cy; t is
designed to allow “code division multiple access” (CDMA) such that gy; t can be
isolated from g̃y.
2. Object space coding modulates the object density prior to measurement to obtain
the forward model
Z
gy  f xCxhx; ydx: (5)

Again, Cx enables the use of CDMA to increase the effective dimensionality of
the measurements.
3. Transform subsampling expands the subsampling strategy of [14] to optimize
which portions of the transform space are measured.

CDMA is, of course, most commonly understood in the context of multiuser com-
munications. CDMA considers the case that a set of relatively low frequency signals
f i t must communicate over the same channel. Multiplication of each signal with an
independent high-frequency code C i t enables one to isolate each signal even when
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 763

the overall transmitted data is gt  Σi f i tC i t. This is achieved by assuming that
the codes are orthogonal over short time windows such that
Z t X Z
gt0 C j t0 dt 0  f i t C i t 0 C j t0 dt 0  f j t: (6)
t−T i

In effect, coding turns the 1D measurement over time into a 2D measurement over
time and transmitter index. In the same way, coding in tomography systems effectively
increases the dimensionality of the measurements. The question “What is the maxi-
mum bandwidth of f t relative to the bandwidth of Ct such that this dimensionality
increase can be achieved?” is a central issue in CS theory.
As mentioned in Section 1, demonstrating that one can effectively reconstruct f from
Eq. (1) or f x from Eq. (2) is insufficient justification for compressive measurement.
One could argue, for example, that one could match the snapshot reconstruction
advantage of compressive tomography by simply measuring a complete set of data
with faster sampling time. In some cases, this may in fact be the case. To justify com-
pressive measurement one must ultimately argue that system performance metrics
(e.g., resolution, SNR, or feature sensitivity) using CS exceed corresponding metrics
for noncompressive sampling under similar operating constraints (e.g., sampling time,
illumination energy, size, weight, power, and cost.) Our goal in subsequent sections is
thus not just to show that compressive measurement is feasible, but that by judicious
application of coding strategies 1–3, one can design systems where compressive mea-
surement is essential.

3. GEOMETRIC TOMOGRAPHY
Geometric tomography consists of imaging using fields that propagate as rays. The
geometric optical field is described by the radiance, Bx; k, which is the power den-
sity at point x propagating with wavevector k. Definition of radiance as a transforma-
tion of the cross-spectral density [42] allows precise analysis of approximations
necessary to allow geometric analysis. For present purposes, it suffices to note that
the radiance is well behaved if interferometric effects are not present and field and
object features of interest are much larger than the wavelength [1,43]. Geometric
analysis is the basis for x-ray, gamma-ray, and positron tomography; light-field
imaging; and coded-aperture imaging.
The modern concept of compressive measurement arose most directly from geometric
tomography. As noted above, image estimation from incomplete data has been part of
x-ray imaging since the beginning of computed tomography [34]. Convex optimiza-
tion strategies used in initial CS studies have been used in tomography for over 30
years [44,45] and the use of the l1 norm to invert sparse geometric projection data
predates modern CS terminology [46]. Numerous subsequent studies have considered
TV and l 1 constrained reconstruction of sparse projection data [47–53]. While these
techniques recover missing data in transform space, they are closely related to inpaint-
ing strategies that recover missing data in image space [54]. Similarly, light-field cam-
eras that reconstruct 3D scenes from periodic samples in ray space may be regarded as
decompressive inpainting systems [55].
A major focus of this review, however, is to note that CS is not simply a signal-
processing technique for restoration of missing data. While there is no harm in in-
cluding inpainting in the family of CS techniques, CS ideally integrates design
and coding of the physical sampling model with prior knowledge and with estimation
algorithms. Physical layer coding for geometric tomography includes the selection of
illumination and detection geometry, source and detector characteristics, and optical
764 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

elements. Illumination strategy and absorbing elements, such as coded apertures and
reference structures, are the main coding mechanisms for geometric tomography.
Absorbing apertures in the form of pinhole cameras have been used for centuries. With
the emergence of computational imaging, more sophisticated coded apertures were
proposed to improve optical throughput in spectroscopy [56] and imaging [57]. Over
several decades of development using linear inversion operators, most coded-aperture
spectroscopy systems have focused on Hadamard codes [11], and most imaging sys-
tems have focused on uniformly redundant arrays [17]. Neither approach is designed
or optimized for tomographic imaging. Amid a flurry of interest in the late 1970s and
early 1980s, one group considered coded-aperture tomography using language essen-
tially identical to modern CS studies [58,59]. Subsequent interest in coded-aperture
tomography has been sporadic, with one recent study concluding that a coded-aperture
tomographic system based on mask translation is equivalent to a poorly conditioned
limited angle Radon system [60]. Brady et al. entered this area through explorations
of 3D coding structures for geometric imaging [61,62]. We subsequently examined the
use of 3D codes for compressive tomography [25] and the complexity of codes imple-
mented by multidimensional obscurants [63]. As discussed in Section 5, we were
momentarily distracted by studies of coded apertures in spectral tomography. This
work came back to geometric tomography in a 2009 study of the mathematical rela-
tionship between coded-aperture spectral and x-ray imaging [64]. More recently, we
demonstrated a new approach to coded-aperture tomography using apertures orthogo-
nal in scale rather than translation [65].
The primary goal of this section is to describe how to use illumination and coded
apertures to achieve compressive tomography. Tomographic system design addresses
physical and mathematical challenges. The primary physical challenge for tomogra-
phy is that individual voxels cannot be independently measured. The primary math-
ematical issue is that, even were independent measurement possible, it would probably
not be the optimal approach. As noted in [24], sequential measurement of the principal
components of the image class returns information faster than a scan of voxel values.
More generally, information theoretic measures may be used to choose a mathemati-
cally optimal measurement strategy [66]. While both physical constraints and insuf-
ficient prior information make full implementation of these strategies difficult, they
are useful guides to design. Design seeks to balance optimization against practical
issues in physical layout, data transfer, and computational processing.
Design begins with analysis of the physical nature of the system. The physical system
includes models for field propagation, field–matter interactions, and signal detection.
In geometric systems, the field propagates by ray projection. The field interacts with
matter via ray attenuation, scattering or emission. Detectors capture the integrated
irradiance of all rays striking them. From these simple models, one may construct
surprisingly sophisticated measurement strategies. Figure 1 illustrates the basic geom-
etries for field propagation and coding. In the sketch at the upper left, the irradiance of
an illumination ray is detected after attenuation on propagation through an object.
Representing the object density as f x, the detected signal is based on Beer’s law
for exponential attenuation:
R
− f xdx
g  I o e ray ; (7)

where the integral is over the line the ray follows through the object. The set of all line
integrals through the object is the basic measurable quantity of this system.
Measurement and coding strategy consists of selecting a subset of these integrals
to measure, determining the order in which they are measured, and deciding whether
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 765

to measure them individually or in groups. Traditionally, these design decisions for


attenuation tomography have been limited to the small subset of pencil beam, fan
beam, and cone beam illumination systems. As discussed here, compressive tomog-
raphy suggests that performance metrics may be improved by broadening this
design space.
The second row of Fig. 1 illustrates the two basic coding strategies for geometric
tomography. The two sketches on the bottom left illustrate illumination ray coding,
which may be implemented using source geometry and apertures between the source
and the object. Scatter tomography differs from attenuation tomography in that the
combination of illumination and detector coding enables individual voxels to be inde-
pendently addressed. The radiance at the detector for scatter tomography is
Z
gx ; k   Bx  x0 − αk 0 ; kf x  x0 − αk 0 ; k; k 0 dα;
0 0
(8)

where Bx; k is the illumination radiance at x with wavevector k, f x; k; k 0  is the


density at x scattering wavevector k into wavevector k 0, and x0 is the detector position.
As illustrated in the top center sketch, it is possible to independently measure every
voxel in f x; k; k 0  by illuminating sequentially with every possible incident ray and
then filtering at the detector to measure each possible scatter ray. This strategy is,
however, unlikely to be efficient. The lower center sketch shows illumination with
multiple rays, which may jointly scatter into the same output ray. Group testing strat-
egies must then be applied to code illumination and detected rays for object estimation.
The sketch at lower right illustrates the use of coded apertures, collimation filters, or
pinholes to modulate the range of rays striking detector elements. The detected signal
for emission tomography, which measures signals generated by fluorescence or radio-
active decay, is
Z
gx0   tx; x0 f xdx; (9)

where tx; x0  is the transmittance of the coded aperture for the ray originating at x and
propagating toward x0 . Without the coded aperture, the detected signal would be in-
dependent of x0 and all detectors would measure the mean object emission with no
spatial selectivity. Thus, detection coding is essential to emission tomography. For
attenuation tomography, detection coding is useful only if the object is simultaneously

Figure 1

Interaction and coding models for geometric tomography.


766 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

illuminated with multiple rays incident on the same detector element. This is not the
case in fan or cone beam tomography systems, but, as discussed below, multiplexed
illumination may be useful in compressive tomography. Similarly, compressive scatter
tomography benefits from both illumination and detection coding.
The last few paragraphs have described the basic physical measures of geometric
tomography and have suggested that some strategies for taking these measures may
be better than others. For attenuation tomography, one may naturally wonder how one
strategy can be better than another given that it is physically possible to measure the
attenuation of each ray independently. The only available coding strategies consist of
deciding whether or not to measure all the rays, in selecting the order in which ray
projections are evaluated, and/or evaluating ray projections using group tests.
To compare measurement strategies we must augment our physical model with resource
constraints and performance metrics. Example resource constraints may include limits
on illumination energy, acquisition time, and detector system complexity. Performance
metrics may include image fidelity, SNR, and feature sensitivity and specificity. Rigorous
comparisons require end-to-end analysis of task-specific performance metrics under op-
erating constraints with realistic object models. End-to-end system comparisons require
choices in object modeling and reconstruction algorithms, however, that are not essential
to this review’s focus on physical sampling strategy. The goals of this section are to show
1. that incomplete ray sampling for attenuation tomography can outperform full sam-
pling when both strategies are confined to the same total exposure;
2. that incomplete ray sampling may be utilized to remove or reduce temporal res-
olution in exchange for tomographic data, and
3. that incomplete ray illumination enables novel strategies for rapid and compressive
scatter imaging.

We use singular system analysis to approach the first point. Short of end-to-end analy-
sis, singular system analysis is a powerful tool for understanding and comparing the
potential performance of imaging systems [67]. Each singular vector is an orthogonal
object feature. If we assume that such features appear with independent probability,
each singular vector corresponds to an independent channel for communicating object
information. The SNR, and thus the information capacity, for each channel is inversely
proportional to the corresponding singular value [1]. If we assume a minimum accept-
able SNR and if we assume a uniform and independent noise variance in each mea-
surement, singular value analysis allows us to identify the number of independent object
features obtained by a given measurement system and singular vector analysis allows us
to consider the structure of object features measured. The singular vectors may be used
to analyze the relative coherence of the sampling and object bases and the probability
that the sampling system satisfies the restricted isometry property (RIP) [68].
As an example, Fig. 2 shows reconstructions of the 64 × 64 Shepp–Logan phantom
from Radon projections at 64 angles spaced over 180°. Figure 2(a) at the upper left is a
reconstruction from the full set of projections with Tikhonov regularization and
Gaussian noise with parameters detailed in an accompanying MATLAB script.
Figures 2(b)–2(d) show reconstructions using only 1/6th of the Radon data sampled
under various scenarios. Figure 2(b) is a limited angle reconstruction using the full
data set drawn from just 180∕6  30°. As one might expect, the limited angle
reconstruction has little longitudinal discrimination. Figure 2(c) is a reconstruction
using 1013 projections drawn with uniform probability from the 6080 samples
MATLAB’s Radon function draws for these parameters. While this reconstruction
includes substantial artifacts, the basic outline of the phantom is clear and resolution
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 767

and artifacts are uniform in all dimensions. Finally, Fig. 2(d) is a reconstruction using
every sixth projection at each of the sampled angles. In general, one might expect
periodic undersampling to produce aliasing artifacts, which are indeed apparent in
Fig. 2(d). Since periodic in Radon space does not correspond directly to periodic
in Cartesian space, however, the strategy illustrated in Fig. 2(d) achieves results
roughly equivalent to random sampling.
Singular value spectra for the measurement strategies of Fig. 2 are shown in Fig. 3.
The singular values of the 2D Radon transform fall as n−4 in the order number n [69].
1

The numerical result in Fig. 3 eventually falls below this limit due to discrete sampling
and numerical effects. Each spectrum is normalized to its lowest order singular value,
so the net values for the subsampled systems are each 6× less than the corresponding
full Radon value if the illumination per ray is constant. Nevertheless, it is interesting to
note that the low order values for the subsampled systems lie above the Radon limit

Figure 2

(a) (b) (c) (d)

(e) (f) (g) (h)

Tikhonov regularized reconstruction from (a) full Radon data, (b) limited angle data,
(c) random projections, and (d) periodically subsampled projections, and (e)–(h) show
truncated SVD reconstructions using the first 256 singular vectors and Tikhonov
regularization for the corresponding upper row images.

Figure 3
0
10
−.25
n
Radon
limited angle
random
periodic

−1
10
0 1 2 3
10 10 10 10

Singular value spectra for the Radon sampling strategies of Fig. 2.


768 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

and that the singular values for the limited angle system appear more attractive than
the random and periodically sampled approaches.
As a first step to resolving this paradox, we note that the singular value spectrum for a
measurement system sampling all the rays illuminating the object from a single angle
is perfectly flat. Since these rays do not overlap, the object region (e.g., the line) ob-
served in each measurement is orthogonal to all other object regions. The sampling
rays themselves represent the singular vectors for this system. This basic structure is
illustrated in the singular vectors for the 30° limited angle approach shown in Fig. 4.
The singular vectors integrate object structure parallel to the observation direction and
produce relatively high frequency structure normal to that direction. This approach is
rich in information if we assume that each ray projection produces independent
information. This would, in fact, be the case if the state of each object pixel were
independent of all other pixels. In practice, of course, we expect adjacent pixels
to be correlated. This means that observation of adjacent ray projections yields less
information than observation of widely spaced rays. This concept may be formalized
in terms of the mutual information associated with diverse measurement strategies
given an object probability distribution [66]. For present purposes, however, it is suf-
ficient to merely compare the singular vectors of Fig. 4 with the singular vectors as-
sociated with randomly sampling an equivalent number of projections uniformly
distributed over the full Radon space, as shown in Fig. 5. To complete these compar-
isons, Figs. 6 and 7 show singular vectors for the fully sampled and periodically
sampled approaches.
The random high-spatial-frequency structure of the singular vectors corresponding to
randomly sampled Radon projections, even at low orders, enables sampling based on

Figure 4
n=1 n=6 n=11

n=14 n=19 n=24

n=27 n=32 n=37

n=40 n=45 n=50

Representative singular vectors for limited angle Radon observations.


Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 769

these functions to discriminate features that are ambiguous when sampled using lower
frequency limited angle or even full Radon data. This effect is illustrated particularly
well in the truncated singular value decomposition (SVD) reconstructions shown in
the second line of Fig. 2. While relatively little loss of reconstruction fidelity is ob-
served for the randomly sampled system, the reconstruction from the full Radon data
is blurred due to the low-frequency nature of the low-order singular vectors. In com-
bination with a sparsity prior, this result suggests that, for a fixed total exposure dos-
age, one may be wiser to expose longer with sparsely sampled Radon projections
rather than sampling the full Radon transform. Alternatively, of course, this result
clearly suggests that satisfactory reconstructions may be obtained from substantially
subsampled data. In practice, sparsity priors must be added to the linear regression to
obtain accurate image estimates. As an example, Fig. 8 shows reconstructions under a
total variation constraint for the randomly and periodically sampled data of Fig. 2.
While the results of Figs. 2–8 present a relatively simple example, they nevertheless
provide clear evidence that satisfactory reconstructions may be obtained while sub-
stantially subsampling ray projections, and that distribution of rays selected for mea-
surement is critical. Reference [70] confirms this assertion in a detailed study of ray
subsampling for transmission tomography. Potential advantages of the CS approach
include reducing radiation exposure, decreasing the number of detectors needed,
and decreasing signal acquisition time. More advanced strategies may choose to

Figure 5
n=1 n=6 n=11

n=14 n=19 n=24

n=27 n=32 n=37

n=40 n=45 n=50

Representative singular vectors for randomly distributed Radon observations.


770 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

adaptively add ray measurements based on image quality metrics during sequential
recording.
Multiplexed illumination, meaning the simultaneous illumination of the object using
more than one source, is the second coding strategy available for attenuation tomog-
raphy. As illustrated in Fig. 9, a coded aperture may be placed between a source array
and the object or between the object and the detector array to assist in source sep-
aration or disambiguation. In practice, the source side coded aperture is the most likely
mechanism for selecting illumination rays from a single source for use in the sparse
illumination strategy explored in Figs. 2–8. Given the attractions of sparse illumina-
tion identified above, it is natural to note that the coded aperture may be designed such
that additional sources may simultaneously illuminate the object and be captured on
detectors not illuminated by the first source. For example, a periodic coded aperture
combined with an appropriately periodically spaced source array allows independent
simultaneous observation from multiple sources. The advantage of this approach rel-
ative to conventional source scanning is that projections from multiple sources may be
acquired in a single time step. The disadvantage, of course, is the higher cost asso-
ciated with multiple illumination sources.
In principle, one may separate data from multiple source illumination even when no
coded aperture is used, in which case the detector data is the sum of the projection
images from the disparate sources. For example, Uttam et al. [71] and Marcia et al.
[72] describe multiple image recovery from superimposed measurements. In practice,
better results are achieved if the superimposed images can be modulated by codes
prior to multiplexing. This allows data plane separation by CDMA. This strategy arises

Figure 6
n=1 n=6 n=11

n=14 n=19 n=24

n=27 n=32 n=37

n=40 n=45 n=50

Representative singular vectors for the fully sampled 2D Radon transformation.


Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 771

with identical mathematical structure in diverse physical forms of compressive tomog-


raphy. Our group has applied it for spectral imaging [73,74] and video compression,
but it may be applied in any situation in which different image planes are modulated
prior to superposition. We have also considered coded apertures to separate superim-
posed images from multiple sources for projection tomography [64]. Given the attrac-
tiveness of sparse illumination and relatively modest value of multiplexing in photon
counting systems, however, one is most likely to choose a sparse coded aperture such
that each detector element captures rays from only one source for the system shown in

Figure 7
n=1 n=6 n=11

n=14 n=19 n=24

n=27 n=32 n=37

n=40 n=45 n=50

Representative singular vectors for periodically sampled Radon observations.

Figure 8

(a) (b)

Total variation constrained reconstruction of (a) randomly sampled projections and


(b) periodic in angle projections.
772 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

Fig. 9. Multiplexed measurement is more essential in scatter tomography, which we


consider next.
Scatter tomography presents interesting design challenges in view of the high dimen-
sionality of the target object and the need to distinguish scattered rays. In contrast with
attenuation, where each illumination ray generates exactly one detected ray, an illu-
mination ray generates many scattered rays from every voxel it crosses. Spatial or
temporal filtering or multiplexing must be applied to disambiguate the scattered rays.
A particularly simple example is illustrated in Fig. 10, which shows scattering in the
plane from a single ray incident along the z (horizontal) axis. A sensor arranged along
the x (vertical) axis at a distance d from the z axis origin detects the signal:
Z
gx  f z; θ  tan−1 x∕z  ddz; (10)

where propagation and detector response factors have been omitted for brevity. The
goal of the tomographic system designer is to select a strategy for recovery of the 2D
distribution f z; θ, where θ is the scatter angle, from the 1D measurement of gx.
One strategy, adopted in coherent scatter computed tomography (CSCT) [75], consists of
observing the 1D scatter over the 2D distribution of illumination rays for the 3D object
f x; z; θ. This approach yields Radon projections over the 3D object, which can be

Figure 9

Multiplexed illumination using a coded aperture for source separation.

Figure 10
g(x)

θ x

z d
Multiplexed detection of scatter from a single illumination ray.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 773

inverted by standard methods. One could, of course, subsample the set of observed rays
randomly or periodically, as described above, to implement compressive measurement.
A simpler strategy for reconstructing just f z; θ may be implemented by varying the
distance d of the detector array along the illumination beam. For each value of d, gx
consists of distinct line integrals through f z; θ. Figure 11 shows the curves integrated
for several values of x for a particular system geometry. gx measured over a range of
d is a limited angle Radon transformation of f z; θ, which may be inverted by stan-
dard methods. As above, subsampling in d or x allows for compressive measurement.
As with multiplexed illumination, coded apertures may be used to disambiguate multi-
plexed rays in scatter tomography. With a coded aperture at z  0 a distance d from
the detector plane, as in Fig. 12, Eq. (10) becomes
Z  
xz
gx  t f z; θ  tan−1 x∕z  ddz; (11)
zd

where tx is the transmittance of the coded aperture. The factor z∕z  d in the
argument to t produces a magnification of the aperture code that is one-to-one with z.
A scattering object at a specific z will produce a magnified shadow of the coded aper-
ture, suggesting that the ideal coded aperture would make these signals distinguishable
by minimizing its self-similarity under magnification. Such codes based on sinusoids
were analyzed in Ref. [37], and work best for isotropic scattering models where f z is
the target density at beam position z. With a sinusoidal code, the forward model in
Eq. (11) is invertible since f z is a simple transformation of the measured spatial
frequency distribution.

Figure 11
1.5
1.0
0.5 0.5
0.0 0.0
0.5 0.5
1.0 1.0
1.5 1.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
z

Curves of integration through f z; θ for x  2; 1; 0.25 and 0 for d  0.1 and
d  1. The range of z is [0, 1] and the range of θ is −π∕2; π∕2.

Figure 12

Coded-aperture modulation for scatter tomography.


774 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

The sinusoid in one possible scale code and some others are shown in Fig. 13. The five
examples shown include a sinusoid, a periodic binary code obtained by rounding the
sinusoid to 0 or 1, a random gray-scale code, a random binary code, and a tiled modi-
fied uniformly redundant array (MURA) of length 5 [17]. To evaluate the best code for
the isotropic pencil beam system, a numerical analysis was performed. The function
f z was used in place of f z; θ in the forward model [Eq. (11)] to model isotropic
scattering, and the forward matrix was computed by numerical integration over rec-
tangular pixels in x and z. The simulation assumed d  1 and the frequency of the
sinusoid was such that ud  64. The detector was modeled as 256 discrete pixels
covering x  0 to x  1. The object was assumed to lie in the domain z  0 to z  2,
partitioned into 64 pixels. Figure 14(a) shows the singular values for the pencil beam
system employing each coded aperture, and Fig. 14(b) shows the point spread func-
tions (PSFs) for an impulse at the center of the object domain. The singular values of
the sinusoids decay more slowly than the random codes. The binary versions of each
code maintain larger singular values than their gray-scale counterparts. The singular
values of the MURA aperture dominate the other curves. Another way to evaluate the
system is to examine the PSFs of Fig. 14(b). Each code produces a PSF with correct
peak location; however, the PSFs vary with respect to contrast and peak width, fol-
lowing a trend one expects from the singular value plots. The binary codes outperform
the gray-scale codes, and the random codes produce undesirable artifacts for locations
close to the aperture. The tiled MURA code shows the best contrast, noting, however,
it has average throughput of 40% in comparison with 50% for the rest. The length-5
MURA was chosen since it produces the best results compared with the other choices
for code length. While MURAs and other high-resolution shift codes have been used
extensively for transverse imaging, these results suggest that tiling a short MURA is a
good strategy for constructing a scale code for ranging. The PSF is a particularly
informative picture when considering objects with sparse support. A compressive
approach to the pencil beam system might subsample x with a few discrete pixels,

Figure 13

One-dimensional coded apertures, top to bottom: sinusoid code, binary sinusoid code,
random gray-scale code, random binary code, and tiled MURA of length 5. The
horizontal axis is x.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 775

assuming f z has limited support and/or minimal total variation. The fact that a signal
with sparsely populated coefficients may be reliably recovered from sparsely sampled
measurements is a classic result of compressed sensing [14].
While f z is a sufficient description for objects causing isotropic scattering, the den-
sity f z; θ is a more general scattering function that applies, for example, to coherent
scattering from liquids, powders, and amorphous solids. The dependence of f on θ
can provide valuable information about chemical structure. Reconstruction of the
2D f z; θ from 1D measurements of gx as in Eq. (10) is truly compressive, and
is achieved by decomposing f into a sparse number of basis functions. These could
include a dictionary of materials for the θ dependence, and wavelets or another basis
of choice in the z coordinate. However, if f z; θ is not compressible, 2D measure-
ments are required. Reference [37] presents coded apertures enabling reconstruction
of f z; θ when the measured scatter image is the 2D function gx; y, acquired by a 2D
detector array. Since the scatter signal depends only on the polar angle θ, the multi-
plicity of scattered rays can be exploited with a coded aperture breaking the symmetry
in the azimuth angle ϕ. For this approach, a 2D code tx; y is preferred, but not ab-
solutely necessary, as demonstrated in the experimental system of Ref. [65], which
is a straightforward extension of Fig. 12 to a 2D aperture and detector. That system
measured f z; q in a snapshot by acquiring scatter images with a coded aperture based
on a square wave in x. Here, q is the momentum transfer, which is one-to-one with θ
at each x-ray energy. For the experiment, two sample vials were placed at different
locations along the pencil beam, one containing NaCl and another containing Al pow-
der. By using a nonlinear estimator with an accurate physical model, the scatter image
was processed to reconstruct the scattering density f z; q. Cross sections of the re-
constructed f z; q are shown along space (z) and momentum transfer (q) in Fig. 15.
The reconstruction correctly separates the two objects in z and recovers the dominant
peaks in q for each sample. These results show the capability of the coded-aperture
system to measure tomographic electron density and momentum transfer distributions.
The concepts behind the pencil beam system may be extended to a fan beam, as in
Ref. [76], which is an example of a system that can measure tomographic images of
homogeneous objects. The task is to measure Fy; z; θ  f y; zbθ for some un-
known density image f y; z and scattering function bθ. The fan beam geometry
is an extension of Fig. 12 to planar illumination within the y–z plane and a 2D aperture
and detector parallel to the x–y plane. Figure 16(a) shows a test object consisting of

Figure 14
0 Sinusoid (grayscale) Sinusoid (grayscale)
10 Sinusoid (binary)
1 Sinusoid (binary)
Random (grayscale) Random (grayscale)
Random (binary) 0.9 Random (binary)
MURA (binary) MURA (binary)

0.8

0.7
−1
10
0.6

0.5

0.4
−2
10
0 10 20 30 40 50 60 70 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Index z/d
(a) (b)

(a) Singular values and (b) PSFs for each coded aperture in Fig. 13. For clarity, each
curve has been normalized according to its peak value.
776 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

plastic “DUKE” letters placed in the fan beam, and Fig. 16(b) shows the reconstructed
density image. These results demonstrate the capability of the fan beam system for
snapshot 2D tomography. In the discrete fan beam forward model, if f y; z is
expanded over N modes and bθ is expanded over M modes, the assumption of
homogeneity reduces the solution space’s sample dimensionality from M N to M  N .
The success of the fan beam reconstruction suggests that more general scattering den-
sities f y; z; θ may be reconstructed if they are compressible down to 1∕M  1∕N
percent.

Figure 15
(a)

−60 −55 −50 −45


z (cm)

(b) 1 (c) 1
Reference Al spectrum
Reference NaCl spectrum
0.9 0.9 Spectral estimate
Scattering probability F(q) ./ a.u

Scattering probability F(q) ./ a.u


Spectral estimate

0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
15 20 25 30 35 40 45 50 55 60 15 20 25 30 35 40 45 50 55 60
Momentum transfer q in inverse nm units Momentum transfer q in inverse nm units

Coherent scatter reconstruction results for the pencil beam system in Ref. [65] with
NaCl and Al vials simultaneously in the beam. (a) Spatial scattering density f z; q
integrated over q with NaCl and Al vials in the beam. (b) Scattering density f q for
NaCl at z  −59.3 cm. (c) Scattering density f q for Al at z  −52 cm. These re-
sults show the capability of the coded-aperture system to measure tomographic elec-
tron density and momentum transfer distributions.

Figure 16

(a) Photo and (b) reconstructed density image of the plastic “DUKE” letters measured
in the fan beam experiment of Ref. [76]. These results demonstrate snapshot 2D
tomography.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 777

Effective 2D measurements are also possible from a 1D array of energy sensitive pixels,
which measure the energy distribution of the detected x rays. The measurements are then
the 2D function gx; ν, where ν is the x-ray frequency. The energy sensitive approach
has also been applied to CSCT [77] to recover f y; z; q using Radon methods when
gx; ν; τ is measured, where τ is a time coordinate as the object is rotated relative
to the apparatus. Reference [78] demonstrates recovery of f z; q from a “snapshot” mea-
surement of gx; ν. Figure 17(a) shows the background-subtracted scatter signal for a
10 mm sample of high-density polyethylene (HDPE). The reconstructed f z; q is shown
in Figs. 17(b) and 17(c) is a correlation map with the material’s known form factor, and
Fig. 17(d) is the form factor at the location of peak density. A compressive version of the
energy-sensitive system was also demonstrated in Ref. [79], where a single pixel was
used to measure gν and recover f z; q. The single-pixel approach relied on an aper-
ture with subpixel features, producing a z-dependent modulation of the measured energy
spectrum. Each of these energy-sensitive experiments demonstrates that energy sensi-
tivity increases both the sample and measurement dimensions of the scatter signal.
Raster scanning a pencil or fan beam is a straightforward approach to characterizing a
complete volume—however, based on the analysis for compressive transmission tomog-
raphy, a sparse, distributed sampling of the illuminating ray space is expected to yield
better measurement efficiency for compressible objects. Given that the object dimen-
sions are the four coordinates x; y; z; q for coherent scatter imaging of static 3D ob-
jects, the object dimensionality would be matched by a 2D array of energy sensitive
pixels acquiring video-rate data as the object is scanned linearly or rotated, providing

Figure 17

Results for the pencil beam system of Ref. [78], employing a linear array of energy-
sensitive detectors. (a) The background-subtracted scatter measurements for 10 mm
HDPE, (b) the reconstructed f z; q, (c) spatial correlation map with the known sam-
ple form factor, and (d) recovered momentum transfer profile at the location of peak
density, compared with the known form factor. These results demonstrate the ability of
energy sensitivity to increase the measurement dimensionality.
778 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

measurements gx; y; E; t. If the scattering material is known and only the density
f x; y; z is desired, an energy-sensitive detector may not be necessary, as the series
gx; y; t would match the dimensionality. Alternatively, a snapshot of an energy sen-
sitive array would provide the same dimensionality through gx; y; E. The experiments
described above show that compressive scatter tomography is also possible, where the
object dimensionality exceeds the measurement dimensionality. Ultimately, design of a
compressive tomography system for a particular task will involve joint optimization of
the coded apertures, acquisition geometry, simplifying assumptions about the target
objects and noise statistics, and use of available x-ray source and detector resources.

4. DIFFRACTION TOMOGRAPHY
Diffraction tomography consists of imaging using fields that propagate as coherent
waves [80,81]. In electromagnetic systems, the vector electric field Ex; t or the mag-
netic field Hx; t describes wave propagation. The electric and magnetic fields and
their propagation properties are determined by boundary conditions, the Maxwell
equations, and wave–material interactions. Since few objects radiate coherent wave
fields, diffraction tomography analyzes the field scattered by coherently illuminated
objects. Diffraction tomography arises in many contexts, including radar [82] and
holography [83] using electromagnetic fields, and ultrasound [84] and seismology
[85] using mechanical fields.
Wave models are necessary for diffraction-based imaging systems with wavelength-
scale resolution. Aside from the obvious distinction that diffraction tomography uses
waves and geometric tomography uses rays, many more subtle differences arise from
differences in source, object, and detector characteristics. The most profound differ-
ence is that time and phase are explicitly tracked in wave forward models, which
enables range imaging based on time of flight or phase delay. The fact that the wave
field is complex valued also offers benefits in that complex valued forward models
tend to be better conditioned than the real and nonnegative forward models of geo-
metric and coherence tomography. The complex valued field offers serious challenges,
however, in the form of speckle artifacts. In addition to its impact on image quality,
speckle is not compressible. The high sensitivity of the phase of the coherent field to
scatter and refractive variation mean that volume imaging of densely populated vol-
umes is challenging and less common for coherent systems. More commonly, diffrac-
tion tomography is used to image surfaces [86], a finite sequence of interfaces [87], or
a sparse collection of point scatterers [88].
As with geometric tomography, the nonlinear regularization and image estimation
from incomplete diffraction data have long been applied in diffraction tomography.
Rather than attempt a summary of this history here, we refer the reader to the com-
prehensive review by Potter et al. [89]. The first demonstration of deliberate coding
for compressive imaging with coherent fields was presented by Chan et al. [90]. A
theoretical discussion of compressive radar imaging using current language is pre-
sented in [91]. To our knowledge, the compressive holography experiments presented
in [87] are the first demonstration of compressive diffraction tomography using the
modern framework. Subsequent work on compressive imaging using coherent fields is
overwhelming, so we list only a few of the > 350 citations to [91] here [92–96]. While
most of these studies have focused on 2D image formation, several have considered
tomographic imaging [97]. Li et al.’s study of diffraction tomography [98] is particu-
larly interesting in its consideration of random reference structures reminiscent of
similar structures used in geometric tomography [25,61].
While analysis of experimental data is relatively rare in radar studies, many exper-
imental studies of compressive holography have appeared. As described in the recent
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 779

review by Rivenson et al. [99], most studies focus on subsampling conventional mea-
surements for Fresnel holograms [100,101], or phase or frequency [102] shifting.
Extensions of compressive holography include superresolution [103], multispectral
imaging [104], imaging through occlusions [105], and video-rate tomography [6].
Compressive diffraction tomography has also been demonstrated in millimeter wave
systems using holography [106] and coded illumination [88].
In contrast with the diversity of field–matter interactions explored by geometric tomo-
graphy, diffraction tomography almost exclusively images scatter. Assuming mono-
chromatic illumination ψ o r with wavevector k  2π∕λ, the scattered signal under
the Born approximations is
Z
ψ s r0 ; k  f rψ o r; kGr − r0 dr; (12)

where G is the solution to the inhomogeneous wave equation

∇2  k 2o Gr − r0   −δr − r0 : (13)

The simplest geometry for obtaining a set of measurements about f from ψ s is to


measure ψ s from a plane wave (ψ o r  expik · r) and with g sampling ψ s point-
by-point. In this case, the measurement set g becomes gr0 ; k  ψ s r0 ; k, and the
information obtained about f from g can be seen clearly in the Fourier space of
the measurement. Taking the 3D Fourier transform of g with respect to r0 allows ap-
plication of the convolution theorem, yielding

ĝu ∝ fˆ u − uo Ĝu; (14)

where ^ denotes a Fourier transform and u is the angular spatial frequency vector
ux;y ; uz . (Proportionalities will be used to ignore normalization differences between
2D and 3D imaging scenarios.) Ĝu can be obtained by Fourier transforming both
sides of Eq. (13), yielding
1
Ĝu ∝ : (15)
juj − k 2o
2

For the sake of simplicity of notation, we will work in only two dimensions temporarily,
and assume a single spatial frequency propagating in z incident on a line of detectors
located at z  zd . This yields the following expression for the measured field:
ZZ ˆ
f ux ; uz − k o 
gx; zd  ∝ expjux x  uz zd dux duz : (16)
u2x  u2z − k 2o
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Integrating with respect to uz first, and around the poles at uz   k 2o − u2x ,
yields
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
Z fˆ u ; k 2 − u2 − k  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
x o x o
gx; zd  ∝ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp j k 2o − u2x zd dux
k 2o − u2x
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
Z fˆ u ; − k 2 − u2 − k  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
x o x o
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp j k 2o − u2x zd dux : (17)
− k o − ux
2 2
780 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

One term has only positive traveling components and one term has only negative trav-
eling components, which correspond to the forward and backward scattered waves.
If the line of detectors is on the source side, only the negative traveling term is de-
tected. If the object is between the detectors and source, only the forward traveling
wave is detected. In general, the backscatter is more commonly detected in radar
[107], ultrasound [108], and in optics [86], but occasionally the forward scattered
wave is detected for phase objects [87]. In both cases, the detected signal is a pro-
jection of the object that only contains information about the object’s scattering den-
sity along a circular arc in Fourier space. The length of the arc is determined by the
effective numerical aperture of the sensing system [80]. In other words, the maximum
acceptance angle of the detector determines the transverse resolution measured
(Fig. 18). The curvature of the projection also yields a small amount of axial resolution
in most cases. Finally, other limitations arise from finite sampling, which leads to
aliasing artifacts, and the limitations imposed by the Born approximation, which as-
sumes weak scattering. We shall assume that the scatterers form a weakly scattering
scene with support small enough to be covered by the sampling period of the detector.
The other limitations we seek to overcome by encoding the measurement space or
exploiting structure in the signal.
For a system of numerical aperture much less than 1, the system can be well approxi-
mated as encoding a single axial spatial frequency (Fig. 18). This poses a severe prob-
lem for classic holographic techniques. Classically, holography backpropagates a field
measured at a single wavelength. Since only one axial spatial frequency is measured,
the out-of-focus planes will be present in the estimate unless one of two options is
pursued. The first option is to use a nonlinear estimation technique. This option has
been explored in [87], and has been shown to work well for, total-variation (TV)
sparse scenes. The second option is to fill in the axial space through translation, ro-
tation, or pulsed/swept sources by brute force. This, of course, increases the complex-
ity of the imaging hardware and therefore increases other costs, such as acquisition
time and monetary cost. The cost increase has led to the exploration of adaptive and
compressive techniques [88,109].

4.1. Compressive Holography


To break the range ambiguity and increase the signal-to-background ratio, Ref. [87]
first introduced the concept of compressive holography. This can be achieved by pos-
ing the coherent estimation problem introduced previously into a linear estimation

Figure 18

(a) (b)

(a) Simple setup for diffraction tomography and (b) the corresponding spatial fre-
quency space. The transverse resolution is dominated by acceptance angle θ, and
the axial resolution is dominated by wavelength range Δk .
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 781

problem of the type posed in Eq. (1). Holography is used most often at optical frequen-
cies when measuring the field directly is difficult due to the fact that phase-sensitive
detectors do not exist. To measure the field such that diffraction tomography may be
performed, an interferometric system, as in Fig. 19, must be used. Isolating the field is
then performed by simply filtering out the reference wave. For the sake of brevity in
our discussions, we refer the reader to [110] for discussion of the basic holographic
principle and assume we measure the field directly. A reasonable assumption for the
schematic shown in Fig. 19 is that the illuminating field is a plane wave. Under this
assumption and the first Born approximation previously discussed, the field in the
plane of the detector may be represented using the angular spectrum fields propagator
in place of convolution with G, resulting in
Z   qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
zmax
gx  F −1 F f x exp −jz k 2o − u2x dz  n; (18)
zmin

where F represents a Fourier transform [1]. If we discretize this equation (observe dis-
crete pixel measurements), then we obtain the following linear systems of equations:

g0i  F −1 Qi F f i ; (19)

Qi  diagvecSi ; (20)


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vecSi m;n  exp −jzi k 2o − u2xm ; (21)

X
g g0i  n; (22)
i

H  F −1 Q1 F ; F −1 Q2 F ; …; F −1 QN z F ; (23)

Figure 19

Basic holographic recording geometry. A Mach–Zehnder interferometer allows for a


reference beam to interfere with a reference beam for coherent detection using an
incoherent detector. The illumination of the object is assumed to be a plane wave,
and the reference beam is often assumed to be a tilted plane wave to allow for easy
filtering. Adapted from “Recording a hologram” (https://en.wikipedia.org/wiki/
Holography/) in Wikipedia, Bob Mellish. CC BY-SA 3.0.
782 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review
2 3
f1
6 f2 7
6 7
f6 .. 7; (24)
4 . 5
fNz

g  Hf  n; (25)

where i is the index of planes of different depths. Using Eq. (25), the problem of es-
timating f from g is clearly seen to be underdetermined. Since the detector measures a
1D signal (2D in reality, but we will limit ourselves to one transverse dimension to make
the figures and mathematics more clear), even if each plane is critically sampled, the
matrix H becomes a compressive matrix, becoming wider with each additional axial
slice in the estimation space. An example of the sampling matrix for a small scene
is shown in Fig. 20. (The matrices in this section assume no aliasing artifacts. In reality,
a large amount of zero-padding for the scene must be employed to make use of the
angular spectrum method. We neglect this in this overview discussion for simplicity
of simulation and illustration, but it should be noted for completeness and for exper-
imental implementation.)
A backpropagated field result for simulated measurements using the matrix shown in
Fig. 20 on the object shown in Fig. 21(a) is shown in Fig. 21(b). For the sake of
simplicity of argument and demonstration, circular artifacts are neglected in this dem-
onstration. The simulated field has a wavelength of 1 μm and was sampled by a 1 cm
wide detector with a 100 μm pixel pitch. The backpropagated field is exactly the cor-
rect field at each location from the first Born approximation as a result of Green’s
theorem. We, however, wish to estimate the scattering density, not the resulting scat-
tered field. This is an important distinction that means solving the aforementioned
underdetermined problem. The underdeterminancy can be overcome if the source
is sparse in some basis. Many scattering densities will likely fit this model. For in-
stance, some common conceivable scattering densities can be thought of as a sparse
array of point scatterers, or as a scattering surface with smoothly varying contours.
The holographic field f̂ may be reconstructed by solving

Figure 20
(a) z(cm)
76 84 92 100 108 116 124
−4
−2
x(mm)

0
2
4
−2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5
x(mm)

(b) z(cm)
76 84 92 100 108 116 124
−4
−2
x(mm)

0
2
4
−2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5 −2.5 2.5
x(mm)

(a) Absolute value and (b) angle of a holography sensing matrix.


Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 783

˜ 2  τ‖f‖
f̂  argmin‖g − Hf‖ ˜ 1; (26)

and the contour model can be reconstructed by solving

˜ 2  τ‖f‖
f̂  argmin‖g − Hf‖ ˜ TV : (27)

The ‖ · ‖1 term indicates the l1 norm and enforces canonical sparsity; the ‖ · ‖TV term
indicates the TV pseudo-norm and enforces sparse gradients, and τ is a weighting term
between the two terms of the objective function. For simplicity, we present the results
using the l1 norm on a small set of point sources in Fig. 21(c). See [87] for various
examples using the TV pseudo-norm.
While considering holography as a medium for diffraction tomography, we make a
brief aside to observe the effect of a type of corruption present in all forms of coherent
imaging: speckle. Speckle is the effect of rough surfaces creating interference patterns
that cannot be resolved by the spatial bandwidth of the sensing system. This results in
an illuminated scattering density x, which can be approximated to having a complex
circular Gaussian amplitude distribution [111]:

Figure 21

(a) Object composed of point scatterers, (b) the backpropagation of the measured field,
and (c) the compressed sensing reconstruction of the object.
784 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

x ∼ CN 0; Rx ; (28)

Rx  diagf: (29)

The complex distribution from which x is drawn has a pixel variance equal to the
band-limited scattering density f we wish to estimate. Since the true scattering density
is measured through a nonunitary system, the statistics of the measurement are af-
fected by the system:

g  Hx  n; (30)

g ∼ CN 0; HRx H†  n: (31)

The problem is now not estimating f from some generalized inverse of H applied to g,
but to estimate the diagonal covariance matrix of x from g. From a single measurement
set this is difficult, since the statistics of g are not well developed. There have been
several attempts in the literature to estimate the covariance of x directly from the
covariance of g via maximum-likelihood or similar estimators [112,113]. These meth-
ods, while effective, can be computationally costly. A method for directly applying the
compressed sensing techniques presented throughout this paper was initially shown in
[86], and has recently been extended to more complicated measurement schemes in
[114]. We will sketch the strategy of [86] here, and refer the reader to the inspiring
work for the implementation details.
Given many measurements fggi1:M , a set of minimum variance estimates of fx̂gi1:M
may be created from

x̂i  H† HH† −1 gi ; (32)

 H† gi : (33)

A good empirical estimate of the pixelwise variance of x would be the empirical aver-
age of jxi j2 . This empirical estimate can be related to the underlying covariance of x,
and therefore f, by
xi ∼ CN 0; H† HRx H† H  σ 2 H† H; (34)

1X M
s jx j2 ; (35)
M i1 i

≈ diagExx† ; (36)

≈ diagH† HRx H† H  σ 2 H† H; (37)


X
sn ≈ jhhp ; hn ij2 f p  σ 2 ‖hn ‖2 ; (38)
p

s  Bf  σ 2 w; (39)

bn;p  jhhn ; hp ij2 ; (40)

wn  ‖hn ‖2 ; (41)
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 785

where s is the empirical estimate, h denotes a column of H, and B and w define an


affine transformation between the empirical estimate and the parameters of interest f.
Given knowledge of the transformation and s, the bias term can be subtracted, and a
generalized form of the inverse of B can be applied using compressed sensing, as
throughout this paper. In general, however, B will be ill-conditioned due to the under-
determined H from which it was built. For this reason, [86] developed a precondition-
ing method to suppress the weak singular values of B. We refer the reader to that paper
for the details of the method. For an intuitive understanding, the preconditioner can be
thought of as applying a Wiener filter to the data, and then treating the effective matrix
as the product
B̃  B† B  εI−1 B† B; (42)

and the effective data as the product

s̃  B† B  εI−1 B† s − σ 2 w: (43)

This better conditioned forward model is then solved by

f̂  argmin‖B̃ f˜ −s̃‖  τ‖f‖


˜ 1: (44)

Figure 22 shows an imaging example using the same simulated system as before, but
now the point sources are actually diffuse diffraction-limited patches. Backpropagating a
single speckle realization [Fig. 22(b)] yields results that are difficult to interpret due to the
unreliable phase. Averaging many such backpropagations (generating s) yields a better
result [Fig. 22(c)], but contrast is still low. Applying the preconditioner [Fig. 22(d)]
removes nearly all of the background, and the rest is removed via the compressed sensing
convex optimization procedure [Fig. 22(e)]. For the remainder of the diffraction tomog-
raphy discussion, we will neglect the effect of speckle, though it will always be present in
real coherent systems.

4.2. Frequency Scanned Systems


Direct coding of the axial space requires either a scanned or pulsed waveform, or some
form of system rotation. For the purposes of this paper, we will focus on systems that
measure in a single plane, necessitating waveform coding. Waveform coding gives
rise to a diversity of wavenumbers that, in turn, yield a continuum of wave-normal
surfaces: a spherical shell of sampled spatial frequencies. With some rare exceptions
[115], waveform encoded systems are usually spatially scanned systems. This is be-
cause they are often used as focused systems in highly scattering media, as in OCT
[116], or at wavelengths where arrays of detectors are prohibitively expensive, such as
at millimeter-wave and microwave wavelengths [107]. Scanning systems come with
their own set of problems. The scanning hardware is expensive, bulky, and trades off
temporal resolution for spatial resolution.
To limit the trade-off of temporal resolution for spatial resolution, two cases will
be examined. This section will examine limiting the number of scanning
positions necessary to reconstruct an image. The next section will introduce a meth-
odology for multiplexing the range and cross-range information such that an
intelligent encoding and decoding can demultiplex the signal and achieve resolution
in both.
The scanning-based systems we will address in this section are synthetic aperture radar
(SAR) or interferometric synthetic aperture microscopy (ISAM) systems [107,116].
786 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

Figure 23 shows schematics for these two mathematically equivalent systems. Both sys-
tems emit radiation from a spatially band-limited source: a single antenna for SAR
[Fig. 23(a)] and a focal spot for ISAM [Fig. 23(b)], and collect the radiation at the same
point. As the source is swept laterally, a frequency sweep is collected at every point. This
builds an indexed set of measurements gx; k in for a 1D lateral scan defined by

Figure 22

(a) Object composed of point scatterers, (b) the backpropagation of the a single
speckle realization measured field, (c) the average of the fields from 100 backpropa-
gated speckle realizations, (d) the preconditioning estimate from the incoherent esti-
mation model from 100 speckle realizations, and (e) the compressive reconstruction
from the preconditioned model from 100 speckle realizations.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 787

ZZ    0 2
W0 x0 − x2 0 x0 − x2 −1 z
gx; k exp − − jkz − jk  jtan f x0 ; z0 dx0 dz0 ;
W z0  W z0 2 2Rz0  zR
(45)

where W 0 is the beam waist in the focal (antenna) plane defining the band limit, W 
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0
W 0 1  zzR 2 is the beam waist a distance z0 from the focal plane, zR is the Rayleigh
range or the range over which the beam remains focused, and R  z0 1  zzR0 2  is the
radius of curvature of the beam phase fronts. All the parameters are a function of the
wavelength and the numerical aperture of the system. As z0 becomes much greater than
zR , the beam waist and radius of curvature grow linearly with z0 , and the system can be
easily seen to be approximated as a truncated spherical wave. The Gaussian beam is
squared since both the sensing mode and transmitting mode are the same. Because
of the roughly spherical nature of the Gaussian beam, simple Fourier relationships
can be derived between a remapping of gx; k and f x; z, as in the previous section
[116]. For simplicity of understanding, however, we will simply assume a trivial dis-
cretization of the space in x0 and z0 from Eq. (45), allowing for arrival at the equation
g  Hf again. A complete set of x; k data gives rise to a fully determined estimation
problem for a band-limited object. If the scanning time required for a full dataset is too
high, however, the system will need to estimate the object from an underdetermined
measurement set. In compressive scanning broadband systems, instead of extrapolating
from a single k-surface as in compressive holography, holes are left in a k-shell by not
sampling every point in x; k space.
An early example of this was shown in [117], which used random sampling in x and
decompressive inference to estimate a tomographic image from optical coherence
tomography (OCT). OCT is a synthetic aperture (scanning) method akin to synthetic
aperture radar. The main difference lies in the assumption that the beam is focused into
a sample using low-numerical-aperture optics and is, therefore, assumed to be non-
diverging. ISAM links SAR and OCT by solving the diffraction tomography problem
for optical fields. The low numerical aperture necessary to make the OCT assumption
valid has a drastically negative effect on (1) the ability to measure extended axial
objects; because all fields diverge eventually, (2) on transverse resolution because
the beam focus is not as tight, and (3) on the ability to measure compressively on
extended objects, because points separated by less than a beam width interfere coher-
ently. The coherent interference requires that either a Gaussian beam deconvolution
model be used exactly or a Fourier equivalent, such as the u2x  u2z  4k 2 remapping,

Figure 23

Schematics of (a) a SAR system and (b) an ISAM system for sampling a full band
volume by sweeping frequency and moving the detector. SAR systems use RF hard-
ware and measure a coherent field directly. ISAM measures the field through an
interferometer.
788 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

be used. Both methods will yield nearly equivalent results, though the Fourier method
is likely much more computationally efficient.
Another consequence of the confocal OCT assumption is that, even if the object does
lie in the valid region, it will not be sampled well by a compressive scanner if the
object is sparse in the lateral dimension. This is a direct consequence of the RIP
[118]. The RIP states that all objects of equivalent sparsity should be projected with
nearly equivalent energy for CS to be guaranteed to work. Clearly this will not be
satisfied in the case that an object with transverse sparsity is sampled near the focus
of a compressively sampled OCT system. If the beam scan places the scanner over a
point source, it will receive a return signal, whereas if the beam scan skips the point
source, no signal will be returned. In contrast, if a point source is far from the confocal
region, it will be sampled by roughly the same intensity everywhere. Therefore, re-
gardless of where the point source is located and what lateral samples are removed for
temporal sampling efficiency, the RIP will be obeyed (at least for simple point source
targets).
An example of imaging a set of point sources is shown in Fig. 24 for an illumination
beam centered at 1 THz with a bandwidth of 77 GHz focused by a 0.25 NA lens. The
point sources are located roughly 10 Rayleigh ranges behind the focus. The image is
reconstructed in three ways: using the OCT approximation, using the transpose of the
exact model (roughly equivalent to only applying the Fourier remapping without com-
pensating for the bandpass), and using the exact complex Gaussian deconvolution in
each depth. From the figure we can deduce two things. The first is that the transpose
model is almost equivalent to the complete inverse—both produce depth-independent
resolution assuming high SNR. The second is that the OCT approximation fails as
expected for this defocused beam. For depths at which only a single point is present,

Figure 24
Original image OCT estimate
−0.03 −0.03

−0.02 −0.02

−0.01 −0.01

0 0

0.01 0.01

0.02 0.02

0.1 0.15 0.2 0.25 0.1 0.15 0.2 0.25


x (m)

(a) (b)
Transpose image with no bandpass compensation Full inverse estimate with l 1 regularization
−0.03 −0.03

−0.02 −0.02

−0.01 −0.01

0 0

0.01 0.01

0.02 0.02

0.1 0.15 0.2 0.25 0.1 0.15 0.2 0.25


(c) z (m) (d)

Reconstructions of point sources using a synthetic aperture imager centered at 1 THz


with a 77 GHz bandwidth. The beam is focused by a. 25 NA lens. The Rayleigh
range for the configuration is 1.5 cm. (a) is the original image, (b) an OCT approxi-
mation reconstruction, (c) a transpose model estimate, and (d) a full inverse with l1
regularization.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 789

the point may be recovered by naive deconvolution. If that is not the case, the inter-
ference from neighboring points will render naive deconvolution useless.
The problem is exacerbated for the compressive case. In Fig. 25 the same three re-
constructions of the same sample image are shown with the lateral sampling locations
reduced by 50% at random. In this case, even the points isolated in range cannot be
naively deconvolved because the interference makes the blur appear to come from many
phantom points. On the other hand, the measurements are still very well-conditioned if
the correct estimation procedures are used. Using the transpose and a full l1 regularized
inverse as in Eq. (26) both lead to the points being reconstructed with high fidelity,
though clearly with some loss of quality relative to the full measurement example.

4.3. Dispersive Imagers


In the previous subsections we mentioned some specific cases of measurement sys-
tems. These systems all, however, obey the same single measurement property that the
measurement g is a function of the transmitted wave function ψ o , the scattering den-
sity of the object f , and the equivalent transmitted wave of the receiver ψ r. This
amounts to neglecting proportionality constants:
Z
grT X ; rRX ; k  ψ o r − rT X ; kψ r r − rRX ; kf rdr (46)
R

[119]. In classical diffraction tomography and in holography, the receiving detector is


often modeled by a delta function. This makes ψ r r; k a simple spherical wave for
each individual measurement. In SAR/ISAM the transmit and receive modes are
the same so the equation is a function of ψ 2o r; k. Both of these classes of measure-
ment systems rely on independent changes of rT X , rRX , and k to build a complete

Figure 25
Original image OCT estimate
−0.03 −0.03

−0.02 −0.02

−0.01 −0.01

0 0

0.01 0.01

0.02 0.02

0.1 0.15 0.2 0.25 0.1 0.15 0.2 0.25


x (m)

(a) (b)
Transpose image with no bandpass compensation Full inverse estimate with l1 regularization
−0.03 −0.03

−0.02 −0.02

−0.01 −0.01

0 0

0.01 0.01

0.02 0.02

0.1 0.15 0.2 0.25 0.1 0.15 0.2 0.25

(c) z (m) (d)

Reconstructions of point sources using a synthetic-aperture imager centered at 1 THz


with a 77 GHz bandwidth. The beam is focused by a 0.25 NA lens. 50% of the lateral
samples are discarded, making the estimation a compressed sensing problem. The
Rayleigh range for the configuration is 1.5 cm. (a) is the original image, (b) an OCT
approximation reconstruction, (c) a transpose model estimate, and (d) a full inverse
with l1 regularization.
790 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

measurement set. In holography, multidetector apertures are inexpensive, and the


rT X ∕rRX measurements are made in parallel. For systems that do not have inexpensive
detectors, this requires mechanical scanning in general. While compressive techniques
can greatly reduce scan times, it would certainly be beneficial in terms of cost and time
savings to remove mechanical scanning from the system. In the far field of the sensing
system ψ o r; k can be well-approximated as

ψ o r; k ≈ ψ o ϕ; θ; ro  exp−jkr − ro : (47)

Under this approximation, one can intuitively decouple range and cross-range mea-
surements into frequency sweeps for range, and angular or lateral sweeps for cross
range. In other words, it is the cross-range resolution that requires either arrays of
detectors or mechanical scanning in general. To remove mechanical scanning from
systems that require single-pixel detectors, the range and cross-range data must be-
come coupled. This will result in an underdetermined measurements system, and,
once again, we will rely on decompressive inference to decouple the measurements
and return a reliable estimate of the scene.
The simplest implementation of a scheme for coupling range and cross range is a
leaky-wave antenna. If a series of periodic holes is placed in the top of a waveguide,
then a beam will couple out with a transverse wavenumber conserved between the
leaked wave and the difference between the transverse wavenumber in the waveguide
and the wavenumber of the periodic boundary condition, i.e., the grating vector
[Fig. 26(a)] [120]. Neglecting harmonics other than the fundamental, the free-space
transverse wavenumber is

k fsx  k wg
x − : (48)
p
If the driving frequency is changed, the ratio between the transverse and axial com-
ponents of the wavenumber will change, and the beam can be swept across a scene to
perform imaging. In this case, a complete set of transverse information is obtained, but
no axial information is obtained, because there is no finite bandwidth in any given
diffraction-limited angular spot. In order to achieve finite bandwidth over each spot in
the imaging volume, one might consider using a more complicated pattern of holes on
top of the waveguide. This would create a more complicated grating vector and allow
for multiple beams to couple out of the waveguide at each frequency. This philosophy
is flawed for two reasons. The first is that this limits the coding in transverse space to be
convolution codes, which are known to be poorly conditioned. The second is that since
more complicated patterns of holes effectively require more holes, eventually there will
not be enough room on the waveguide for the desired encoding to be performed.
To overcome these limitations, Ref. [88] used metamaterial elements patterned on top
of the waveguide instead of holes. This comes with the advantages of allowing for
phase coding, because the metamaterials have a phase and amplitude response, as well
as having the grating vector change with driving frequency, because the metamaterials
are resonant elements. Explicitly, the radiated field can be built as follows. Assume a
field distribution inside the waveguide ψ wg , which is the exciting field. The dipole
moment of each metamaterial element is the product

mrT X   αrψ wg r; (49)

ω2
αr ∝ ; (50)
ω2 − ω2o r  jω ω2Q
o r
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 791

where ωo r is the resonant frequency of the metamaterial element, and Q is the afore-
mentioned Q-factor. Given this array of excited dipole moments, the radiated field can
be calculated by taking the discrete sum
X
ψ T X r ∝ mrA Gr − rA ; (51)
rA

where G is the Green’s function from Eq. (13). This field composition allows for a
quasi-random encoding in transverse and axial space, which is much more likely to
obey RIP (Fig. 26). If the receiver is also a complicated antenna of this type, the en-
coding based off of Eq. (46) can become better conditioned. In general, however,
higher detection SNRs can be obtained using low-gain antennas; therefore, a balance
is struck between detection efficiency and measurement conditioning by using one
metamaterial antenna and one low-gain receiver.
An example of a set of fields from a 1D antenna with metamaterial elements patterned
on it is shown in Fig. 27. The antenna is 50 cm wide, and measures over a frequency
range of 18–25 GHz. Metamaterial elements are simulated as resonant dipoles with
a Q-factor, a measure of sharpness of frequency response [119], of 200. The high
Q-factor creates an antenna response, and the corresponding radiation pattern, that
rapidly varies with frequency, allowing for imaging to be performed. Figure 28 shows
a reconstruction of a set of point sources measured by a simulated setup with a

Figure 26

(a) Leaky-wave antenna illuminating an object and (b) a metaimager illuminating an


object. The leaky-wave antenna has a periodic excitation that results in a single spatial
frequency, and therefore a narrow beam, to be illuminated from the antenna. The meta-
imager has a complicated excitation that creates a chaotic beam pattern. Since
the excitation is due to resonant elements, the beam pattern changes with frequency
in a complicated fashion as opposed to just sweeping the beam pattern.
792 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

metamaterial aperture of the dimensions described and a single low-gain receiver. The
reconstructed points have resolution equal to the resolution of a fully sampled system
due to the decompressive inference used in the reconstruction: 1 cm range resolution
and 3 cm cross-range resolution.

5. FOCAL TOMOGRAPHY
Focal tomography consists of imaging using fields that propagate as statistical coher-
ence functions. The statistical field is described by the cross spectral density, W x1 ; x2 ; ν,
which is the temporal Fourier transform of the (assumed stationary) mutual coherence
of the electromagnetic fields at points x1 and x2 [1,121]. We use the term “focal
tomography” to describe imaging with this field model because, for reasons discussed
below, focal elements are universally attractive in this context and because the term
“coherence tomography” is commonly used in a different context. Optical coherence
tomography (OCT) uses spatially coherent light to probe 3D structures [122]. OCT
originally used coherence functions to model the detected field in time domain sys-
tems. Since OCT systems have complete control over the coherence state of the illu-
mination source, however, modern systems use tunable narrowband lasers and are
more properly understood as examples of diffraction tomography. Coherence field
models are more essential to the analysis of ambient fields emitted or scattered by
natural objects. Such fields are spatially incoherent at the point of emission or scatter-
ing, meaning that W x1 ; x2 ; ν  0 for x1 ≠ x2 on the surface of imaged objects.
In parallel with previous sections, in this section we review the history of compressive
tomography for systems utilizing statistical field models, we describe the forward
model for such systems, and we present example coding strategies. We begin by
considering the basic physical observables of the coherence field. The basic

Figure 27
10 −10
(a) 1.8 x 10 x 10
4

2 3
Frequency (Hz)

2.2 2

2.4 1

2.6
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
x(m)

10
(b) x 10
1.8 3
2
2
Frequency (Hz)

1
2.2 0
−1
2.4
−2

2.6 −3
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
x(m)

(a) Absolute value and (b) phase of the fields from the metaimager at one plane a
distance of 85 cm from the source.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 793

observable for geometric tomography is the intensity in a single ray. The electric field
itself is the minimal observable for the coherent field. In contrast, the cross spectral
density is not directly observable. Rather, one must use interferometry to measure stat-
istical functions. The most common strategies are two-beam interference and focal
transformations [1]. Two-beam interferometry is commonly used in radio astronomy
[123], where background-free detection of the cross spectral density is possible. It
would be interesting to consider compressive coherence tomography using sub-
sampled interferometry in such systems, but of course astronomical systems work
at ranges where volume imaging is not possible. One might consider compressive
interferometric coherence tomography for passive millimeter wave or terahertz imag-
ing. At optical frequencies, the fact that only the irradiance of the field is observable
means that interferometric measurement of coherence functions is poorly conditioned
for complex objects [124]. Despite this challenge, 4D spatio-spectral tomographic
imaging has been demonstrated using interferometric coherence sampling [20] and
compressive subsampling has been demonstrated for interferometric imaging through
turbulence [125].
Cross spectral densities describe both the emission properties of 3D sources and the
state of the electromagnetic field [121]. For radiation fields, the cross spectral density
on a bounding surface is sufficient to describe the cross spectral density over the
bounded volume. This means that the field cross spectral density reduces to 4 spatial
degrees and 1 spectral degree of freedom. The source cross spectral density, in con-
trast, may in principle be independently distributed over six spatial and one spectral
dimensions. Such a distribution, however, implies relative coherence between fields
emitted or scattered at diverse source points. Natural sources lack such coherence and
are typically modeled as spatially incoherent, in which case the source irradiance at
each point in three spatial dimensions fully describes the radiated cross spectral

Figure 28
(a) 0.85
0.9

0.95
z(m)

1.05

1.1

1.15
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
x(m)

(b) 0.85
0.9

0.95
z(m)

1.05

1.1
1.15
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
x(m)

(a) Sample image of point scatterers and (b) reconstruction from the metaimager.
794 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

density, which may be parameterized in three spatial dimensions and one spectral
dimension [1]. It is possible to use focal transformations to measure the 5D cross
spectral density as described in [126], but more commonly focus is used to transform
the 4D cross spectral density radiated by spatially incoherent objects back into a 3D
spatial and 1D spectral irradiance distribution.
Practical imaging of complex natural objects requires measurements at or near a
physical focus. This means that the basic observable for focal tomography is the ir-
radiance at a focal point. While one could theoretically directly measure the electro-
magnetic field, even for incoherent or partially coherent sources, and use diffraction
tomography to form an image, such measurements are physically impossible. The
goal of a focal tomography system is to estimate the irradiance or spectral density
at each point on the imaged object (or, better, the density of the object itself at each
point in a volume.) The irradiance at a given point is the sum of the power in each
electromagnetic mode focused at that point. The number of electromagnetic modes
focused at a given point in space is equal to the time–bandwidth product of the field.
The time–bandwidth product at optical frequencies for imagers working at 30 frames
per second with an optical bandwidth of 100 nanometers is greater than 1012 . The
number of photons measured in a typical imaging pixel is generally less than 105 .
This means that typical natural imaging systems measure less than one photon per
million optical modes. With such low photon flux, the electromagnetic field (e.g., the
coherent mode amplitude) cannot be estimated with any reasonable SNR. The irra-
diance at the focal spot, in contrast, is a statistical measure integrated over all the
modes, which can be estimated with good SNR. This statistic is best measured at
the focal point; any optical transformation that blurs the field distribution over space
can only make the target distribution less orthogonal to other statistics and harder to
accurately measure.
The magical and wonderful power of lenses to map all of the light radiated at a point in
the object volume to a single point in the image volume is thus central to focal tomog-
raphy. As with other tomographic systems, the primary challenge is that the focused
light spans three spatial dimensions, but the image sensor typically spans only two.
Stereo matching, structured illumination, and projection tomography are the most
common forms of 3D tomography using focal cameras [127–129]. These methods
are forms of geometric tomography, however, and are best understood in the context
of Section 3. Conventional solutions to the focal tomography challenge are to sweep
the sensor through the focal volume as a function of time (image space coding) or to
zoom the lens to change F as a function of time (pupil plane coding) on a fixed sensor
location. Focal tomography then consists of estimating a fused 3D image from a series
of frames. This strategy is closely related to “depth from defocus” [130]. While early
studies of depth from defocus assumed fairly coarse image estimation in pursuit of
3D object shapes, more recent studies use sophisticated estimation algorithms to
obtain high depth-of-field images [131–134]. A recent study applies compressive-
sensing-inspired estimation to defocus image fusion [135].
Rather than give up temporal resolution to acquire the full 3D focal stack, various
methods of compressive focal tomography have been utilized. Several such methods
are discussed in our 2011 review [38]; we briefly review those methods here before
focusing on strategies that have emerged since 2011. For example, snapshot focal
tomography using integral imaging was demonstrated by Hong et al. [136]. Plenoptic
[137] and light-field [138] systems introduce a limiting stop to balance transverse
resolution and radiance characterization. The plenoptic camera re-images the image
volume through a lenslet array in an attempt to characterize the radiance at each point
in a focal image. Light-field imaging may be regarded as the focal tomography analog
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 795

of pixelated RGB-filter-based color imaging. Both approaches assign local pixels to


different ranges in periodic blocks. In contrast to RGB color coding, which aliases
color images without reducing transverse resolution [1], lenslet-based plenoptic sens-
ing low-pass filters the light field. While compressive strategies for light-field imaging
have been considered [139,140], truly compressive focal tomography would ideally
use a coding scheme that uniformly distributes samples over low- and high-
spatial-frequency image components.
Extended depth-of-field imaging using pupil coding and lens design attempts to
achieve snapshot range invariant transverse imaging [18,141–143]. These systems
do not attempt to reconstruct the 3D object or the focal volume; they may be regarded
as strategies to remove the effects of diffraction such that focal imaging is equivalent
to x-ray projection imaging. As such, they may be combined with projection methods
for tomographic purposes [19]. Unfortunately, extended depth of focus (EDOF)
fundamentally blurs focus, which inevitably reduces transverse resolution [38].
Lensless interferometric and coded-aperture systems represent the far extreme of this
approach with infinite depth of focus but also unbounded impulse response support.
Various lens and pupil design strategies seek a balanced approach to loss of resolution
versus extended depth of field. One may alternatively seek to maintain some level of
focus [e.g., to obtain a depth-invariant modulation transfer function (MTF)] while still
allowing a range-variant impulse response [144], but this approach will also ultimately
balance loss of transverse resolution against depth of focus.
As we have seen in several examples in this paper, imager performance per unit mea-
surement is improved if samples can be distributed quasi-randomly over nonconvex
subspaces. In the case of projection tomography, we use masks to distribute samples
over radiance space. In diffraction tomography, we use frequency diversity to distrib-
ute illumination over modal space. This form of quasi-random sample distribution is
relatively challenging for focal tomography, however, because the best sample points
are clustered in image space. Attempts to randomly sample pupil space lead to data
clustered at low frequencies, and systematic sampling in image space as with plenop-
tic cameras leads to severe aliasing and low spatial resolution. To address this chal-
lenge, we proposed 3D coding structures in [38] and showed that such structures could
create high-frequency compressed sampling modes. Unfortunately, the challenge of
fully modeling and accounting for diffraction from such structures would obviate their
theoretical advantages. Alternatively, one might use fiber optics or volume-distributed
sensor elements to randomly sample the focal volume.
In the context of these previous efforts, it is important to note a critical difference
between studies of compressive focal tomography and those of geometric and dif-
fraction tomography. Ray coding and compressive holography strategies outlined in
previous sections have enormous advantages with no substantial cost to the overall
measurement system. It is clearly advantageous to structure the order in which x-ray
projections are measured to minimize dose and maximize information acquisition
rates. It is clearly advantageous to begin estimation of holographic or radar scatterers
prior to acquiring complete diffraction signatures. In contrast, cost–benefit analyses
for light-field imaging and extended depth-of-field lens designs have yet to show that
gains in longitudinal resolution justify the loss of transverse resolution inherent in
these methods. In short, no compelling method for snapshot focal tomography is
known at present. The challenge of capturing the 3D focal field remains, however,
and conventional focal sweeps are unlikely to be the optimal solution.
Array cameras provide a promising platform for compressive focal tomography. To
date, most demonstrations of 3D imaging using array cameras have relied on
796 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

projection tomography under the assumption that the entire object is in focus.
Interesting recent examples include TOMBO [145] and PiCam [146]. Of more rel-
evance to focal tomography, array cameras with diverse focal states have also been
demonstrated [147,148]. More generally, array cameras have been used to compres-
sively estimate video and spectral data cubes [149,150]. Array cameras are more ef-
fective than single-aperture light-field cameras because they operate at the full
resolution of their entrance apertures, allowing capture of similar data with much
lower system cost and volume. They can also easily be designed to uniformly sample
the focal stack, in contrast with relatively irregular sampling in light-field cameras. To
date, however, array cameras have not achieved significant practical deployment. An
array that fully samples the focal volume would be unreasonably large; it still makes
more sense to allow a lens to scan 100 focal positions rather than building an array of
100 lenses. It is reasonable, however, to imagine increasing the speed of the focal
sweep or allowing tracking of targets at several different ranges using arrays of
3–10 cameras. The greatest advantages of arrays are achieved with a large baseline
to allow geometric tomography, with the challenge of obtaining the focal stack left to
each individual camera.
Given that some form of focus adjustment remains the most attractive focal tomog-
raphy strategy, we are tempted to relax our definition of compressive tomography for
this case. Sampling the 3D focal volume is mathematically similar to the sampling the
3D hyperspectral data cube or the 3D video data cube, both challenges that we have
previously addressed using coded-aperture modulation to create CDMA-style patterns
on data cube slices. Coded-aperture snapshot spectral imaging (CASSI) [73] consists
of an imaging diffractive spectrometer with a 2D mask at the first image plane with
spectrally dispersive elements in the relay path to a second image plane. Hyperspectral
imaging generally and CASSI specifically have served as model systems for compres-
sive measurement in numerous studies, including demonstrations of compressive
tomography in video-rate spectral imaging [5] and of various extensions to adaptive
sensing [151] and variable compression ratio [40] More recently, we noted in [39] that
the CASSI approach could be adapted to measure the space–time video data cube.
From a compressive tomography perspective, one can understand this approach as
measuring a finite time 3D data cube from a single 2D snapshot. From a focal tomog-
raphy perspective, it takes only a small step to realize that, if snapshot capture of video
is possible, then that video can be a focal sweep. This focal tomography strategy
effectively achieves the quasi-random focal volume modulation envisioned in [38]
and it was demonstrated in [152,153].
Though focal sweeping shows promise over fixed lens designs and pupil codes, it
performs poorly in terms of depth estimation because the focal sweep kernels are
nearly range-invariant [154]. To gain depth sensitivity, the system’s impulse response
must somehow change as a function of depth. Reference [155] proposes a PSF that
varies helically when defocused; the helical rotation encodes the object’s depth.
Captured images may be deconvolved and localized with nanometer-scale precision
for 3D superresolution microscopy [156]. This pupil encoding may be implemented
directly by a phase profile applied to a lens or indirectly via an spatial light modulator
(SLM). The former case results in a permanent degradation to the imager’s perfor-
mance; the latter case increases system complexity and is limited to a finite spectral
bandwidth due to the SLM.
Alternatively, simultaneous image translation and focal sweeping may be used
to multiplex each focal channel within the object volume into a single measure-
ment. Translating the image in a known way during the exposure period yields a
shift-invariant, depth-variant PSF. Like focal sweeping, this image translation may
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 797

be disabled whenever high-transverse-resolution image or video capture is desired.


Additionally, it does not inherently add substantial complexity to the optical system;
it can even be accomplished within the form factor of a cell phone camera through use
of the optical image stabilization module. Below, we motivate image translation as a
solution to the focal tomography problem and describe the theory behind it.

5.1. Depth-Encoding Focal Sweeps


Let unprimed and primed variables, respectively, denote object- and image-space co-
ordinates. Let f x; y; z; t denote the four-dimensional (3D space and time) data cubes.
The irradiance of an incoherent image formed at the optical system’s detector plane
focused at a distance conjugate to zo is given by
ZZZZ
0 0
gx ; y ; zo   f x; y; z; thzo x − x0 ; y − y0 ; z − zo px x; ypt tdxdydzdt; (52)

where h is the depth-dependent system PSF of the optical system; px and pt are the
spatial and temporal pixel sampling functions, respectively. These are presumably
rectangular windows with respective supports of the pixel pitch Δ and the integration
time Δt .
We assume h takes a Gaussian distribution with standard deviations that vary with
defocus error, since Hermite–Gaussian beams are eigenfunctions of the Fresnel propa-
gation transformation. Under this assumption, the defocused blur kernel h for an object
located at zo conjugate to an image-space distance z from the camera varies according to
the focused beam waist wo and the optical defocus relative to the object plane:
 
x2 y2
1 −π
hzo x; y; z  2 e w2 σ 2 zo ;z
0 : (53)
w0  σ zo ; z
2

Here, the defocus affects the blur kernel’s standard deviation σzo ; z for a wavelength λ
according to the paraxial focus error:

Az 1 1 1
σzo ; z  − − ; (54)
2λ F zo z

where F and A denote the system focal length and entrance pupil diameter. The standard
deviation of the blur grows approximately linearly as a function of the dioptric focus
error. Ten of these blur kernels are shown in Fig. 29 for an optical system with F 
5 mm and A  1 mm, focused at a distance of 1.27 m. The kernels correspond to 10
equispaced dioptric distances within a depth range of 0.04–6.66 diopters (25 m–15 cm);
each kernel is spaced by 0.74 diopters.
We’ve seen that EDOF strategies modulate the pupil (via a deformable optic or a phase
mask) directly or via a sweep of the lens’ back focal distance as a function of time.
Herein, we consider the latter because it is a more commonplace functionality within
existing cameras. Linearly sweeping the focus a distance Δz during the integration
period results in an EDOF kernel given by
Z Δ
T
hs1 ;zo x; y; z  hzo x; y; z − νtdt; (55)
0

where ν  ΔΔzt represents the focal sweep velocity. In the experiments, Δz  172 μm,
meaning the back focal distance translates between 5 and 5.172 mm; this corresponds
798 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

to sweeping an object volume of 0.04–6.6 diopters. The swept focus kernels for this
range are shown in Fig. 30.
Under focal sweeping, Eq. (52) is modified for images with the detector at a position
zo to
ZZZZ
0 0
g z0 x ; y   f x; y; z; ths1 ;z0 x − x0 ; y − y0 ; zpx x; ypt tdxdydzdt: (56)

We next look at the resolution of the captured data g. The 2D space and time Fourier
transform of g is given by

ĝ zo u; v; w  sincΔu; ΔvsincΔt wfˆ u; v; z; wĥs1 ;zo u; v; z: (57)

Equivalently to pupil coding [38], the support of ĥs1 ;zo may approach that of the
stationary focused optical system; however, the modulation transfer efficiency is
reduced regardless of the object’s range, which results in a loss of spatial resolution
and contrast.
As previously mentioned, the focal sweep kernels hs1 ;zo are range-invariant, so Eq. (57)
is approximately constant along the z dimension. If the goal of image translation is to
encode depth into the measurements, the swept focus kernel needs to differ in shape as
a function of depth; a deconvolution with only the kernel coinciding with the origi-
nating depth z0 should yield a feasible image. This difference should be balanced with
SNR and modulation transfer efficiency in mind; nulls or very low values of the result-
ant swept-focus transfer function will facilitate depth estimation at the cost of spatial
resolution [157].

Figure 29

Blur kernels for an F/5, 5 mm conventional camera focused at a distance of 1.27 m.


Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 799

As a metric to design PSFs for depth from defocus and motion deblurring, [157,158],
use the Kullback–Leibler (KL) divergence of two PSF irradiance profiles to define
a discriminability metric between them. According to information theory, a larger
KL-divergence value between two measurements corresponds to a greater information
rate obtained by taking both. Here, larger discriminability metrics correspond to
greater depth estimation capability. The discriminability between two PSFs i and j
is given by
  
1 X σ i u; v σ i u; v
Dij  − log − 1; (58)
N u;v σ j u; v σ j u; v

where N is the number of discrete frequency components and σ i u; v; σ j u; v denote
the variances of PSFs i and j at spatial frequencies u; v. PSFs with smaller discrim-
inability values are more similar to each other than those with larger values; the ideal
set of PSFs maximizes the discriminability while maintaining high optical transfer
function values over the spatial frequency range of interest.
Transverse image motion can translate the PSF such that it changes shape based on its
origin in depth. For example, consider a kernel that, in addition to a focal sweep,
translates the sensor in the x dimension with motion ηt:
Z Δ
T
hs2 ;zo x; y; z  hzo x − ηt; y; z − νtdt: (59)
0

Figure 31 shows how this kernel’s shape changes as a function of originating depth zo
with linear translation: ηt  αt, where α  10 pixels per frame. The Fourier trans-
form of data captured using this kernel is given by

Figure 30

Swept blur kernels for an F/5, 5 mm conventional camera sweeping the object volume
for an object located at 1.27 m. Note the invariance of the kernels as a function of depth.
800 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

ĝ zo u; v; w  sincΔusincΔT wfˆ αu; v; z; αu  wĥs2 ;zo u; v; z: (60)

Note that the temporal passband is shifted by the transverse motion αt; this form of
shifting improves the sampling efficiency for objects exhibiting motion in the trans-
lation direction. More subtly, an object that changes shape or phase during Δt is
smeared onto the detector using this technique: the shape or phase changes may
be recovered through deconvolution. If the overall volume of the spatiotemporal pass-
band is a cube containing 8umax vmax wmax elements, image translation effectively
shears this passband volume; the temporal cutoff frequency increases by a factor
of α while the spatial cutoff frequency is sheared at an angle tan−1 αu
umax
max
. If one as-
sumes that samples are distributed uniformly within this passband, there is no effect on
the overall information rate. However, if the low-frequency content is compressible or
structured, the higher temporal bandpass may be appealing.
We can see that the this kernel’s peak irradiance position is displaced as a function
of zo . Though the kernel changes shape during the integration time, in practice this
is not substantial except at the midpoint and the extremities of the depth range; the
discriminability of kernels hs2 ;zo is greater than that of hs1 ;zo (see Fig. 33) but may be
suboptimal, particularly for kernels near the dioptric midpoint, whereby the kernel has
significant lobes on both sides of the peak intensity point.
We can translate the image in two dimensions to improve the discriminability between
kernel candidates. Two-dimensional image translation modifies the focal sweep kernel
to be
Z Δ
T
hs3 ;zo x; y; z  hzo x − ηt; y − ξt; z − νtdt: (61)
0

Figure 31

Horizontal linear translation focus kernel shapes hs2 ;z0 as a function of object distance z0 .
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 801

This kernel is shown with circular motion of radius a  2.3w0 (ηt  a sin3Δ
2t
t
,
ξt  a cos3Δ 2t
t
) in Fig. 32 such that the kernel is rotated an angle π3 during the in-
tegration time Δt . Note that now the angle, rather than the position, of the kernel’s
peak irradiance encodes its depth; the discriminability gains of hs3 ;zo relative to
hs2 ;zo for tomographic reconstruction may be seen in Fig. 33. This increase is attri-
buted to a distribution of the transfer function’s nulls over a greater region in trans-
verse space.
Rotational image translation is analogous to engineering the PSF as a single rotating
lobe [159] but may be implemented solely by means of a camera’s optical image sta-
bilization module and is configurable at any time. Additionally, similar to Eq. (60),
temporal resolution for motion in each direction improves by a factor of the arc length
projected onto that direction: the spatial resolution in each direction degrades by the
same factor.
Another important fidelity metric for depth estimation is the Cramer–Rao lower bound
(CRLB). This number, which represents the minimum variance an unbiased range
estimator is capable of, is proportional to the system’s information rate according
to the likelihood function of the sensing process and is hence related to the discrim-
inability among the PSFs. The CRLB of depth estimation with helical PSFs has been
discussed in [159,160]. Image translation yields PSFs that behave similarly to the
single-helix PSF; thus, the same analysis applies here.

5.2. Snapshot Focal Tomography


Because of the reduced transverse spatial resolution, we focus on the tomogra-
phic ranging capabilities of these PSFs under the assumption that the sweeping

Figure 32

Circular translation swept-focus kernel shapes hs3 ;z0 as a function of object distance zo .
802 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

may be turned off at any time. This global snapshot ranging technique could lead to
faster and more intelligent autofocus than conventional contrast or phase detection
algorithms.
We illustrate the focal tomography problem with a toy example shown in Fig. 34.
The objects are placed within the same 10 equispaced range bins used in the
analysis presented in Section 5.1. We first simulate this scene as taken by the same
F  5 mm, A  1 mm conventional camera focused on subimage 2 (located 1.27 m
from the camera) with added Gaussian noise; the simulated image is shown on the left
side of Fig. 35. Deconvolving each subimage with its perpetrating kernel yields the
result on the right. These results assume that the kernels are estimated and
known perfectly a priori; in practice, they are unknown beforehand and are
estimated using blind deconvolution algorithms. We use the deconvolution algorithm
proposed by [161] for its quality and speed; a 1 megapixel image takes ∼3 s to
deconvolve.
As expected in the conventional capture case, the high-frequency components within
the images are low-pass filtered by the Gaussian kernels of Eq. (53): they are irre-
coverable by deconvolution. The system’s condition number is poor; amplified noise
is present within the results.
Employing the focal sweep kernels hs1 at a focal sweep velocity ν  170 μm∕Δt
results in the captured and reconstructed images shown in Figs. 36(b) and 36(e). The
hazy appearance is characteristic of all focal sweep data; the resulting MTF is reduced
due to the Fourier power distribution across the focal volume.
Translating the image in one dimension with velocity α  10 pixels per integration
period Δt (i.e., convolving each discrete x; y object plane that is located a distance zo
away from the camera with its corresponding kernel hs2 ;zo ) while sweeping the focus
results in the captured and reconstructed imagery in Figs. 36(c) and 36(f). To generate
the all-in-focus image in Fig. 36(g), each subimage of the captured data is
deconvolved with each of the library’s 10 blur kernels (Fig. 31).

Figure 33
Kernel discriminability (true depth of 3 diopters)
3.5
No sensor motion
1D linear motion
3
2D circular rotation

2.5
Discriminability

1.5

0.5

−0.5
0 1 2 3 4 5 6 7
Depth (diopters)

Discriminability of focal sweep kernel candidates relative to the true kernel located a
distance zo  333 mm away from the camera. Curves for hs1 ;zo , hs2 ;zo , and hs3 ;zo are
plotted in red, green, and blue, respectively.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 803

Rotating the image in a circle of radius a  2.3w0 through an angle π3 while sweeping
the focus results in the captured and reconstructed images shown in Figs. 36(d) and
36(g). Note that both rotational motion and horizontal motion have high-frequency
directional ringing artifacts, as seen, for example, in Fig. 36(f), subimage 1 (candidate

Figure 34

Simulated objects for the image translation focal tomography method. Each object is
assumed planar and is located a different distance away from the camera. Subimages
1–4 are, respectively, modeled at distances 25 m, 1.27 m, 22 cm, and 15 cm away from
the lens. The camera is focused on subimage 2.

Figure 35

Left: simulated images captured by the camera when focused on subimage 2 (upper
right image). Right: the same images after deconvolution with their respective kernels.
The noise has been amplified, and the high-frequency details are lost.
804 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

kernel number 1 located at a distance of 25 m). This effect is attributed to the reduced
MTF of these kernels in the dominant direction of translation and may be overcome
through use of image and kernel priors.
Figure 36

Measurements and deconvolution results for various focal sweeping strategies.


(a) Target object for comparison. Each subimage is modeled from the corresponding
focal distances shown in Fig. 34. (b)–(d) Images simulated with the focal sweep cam-
era convolved with hs1 ; hs2 , and hs3 . The corresponding kernels from the known zo
values are shown in the lower-right corners. (e)–(g) Deconvolution results from
the respective images.

Figure 37
PSNR vs. candidate kernel, correct one is: 1 PSNR vs. candidate kernel, correct one is: 2
35 35

30 30

25 25

20 20

15 15
0 2 4 6 8 10 0 2 4 6 8 10

PSNR vs. candidate kernel, correct one is: 6 PSNR vs. candidate kernel, correct one is: 10
35 35

30 30

25 25

20 20

15 15
0 2 4 6 8 10 0 2 4 6 8 10

PSNR plots versus candidate kernel number for the three different swept focus
kernels. Each subplot corresponds to its subimage. The candidate kernel numbers are
of the order of their descending distances zo . Both hs2 ;zo and hs3 ;zo prove more accurate
for ranging than hs1 ;zo , as can be seen by their more pronounced peaks.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 805

The deconvolution of each subimage that yields the highest peak SNR (PSNR)
is shown in Figs. 36(e)–36(g); the kernel candidate number corresponding to
these results for each subimage is shown in Fig. 37. Note that the PSNR plots
have distinct peaks at the correct image candidate number for the 1D and 2D image
translation kernels; the focal tomography architect wishes to make these peaks as
pronounced as possible. Since one does not have access to the ground truth images,
PSNR may not be used as an evaluation metric. Blind [162,163] or guided decon-
volution algorithms may be used instead of PSNR to find the kernel within the li-
brary that best matches each patch or sliding window of the image. Upon completion
of this process, a tomographic image may be rendered from the single captured
image.

6. CONCLUSION
Compressive sampling has exploded as an academic discipline over the past decade.
While compressive tomography is a more limited subset of this topic, several of the
more practically important implications of CS arise from the ability to reconstruct 3D
and 4D scenes from 2D measurements. As indicated in this review, compressive strat-
egies for x-ray, holographic, radio wave, and focal tomography demonstrate clear
practical advantages over conventional temporal sweeps. While considerable research
in optimal and adaptive strategies remains, it is clear that consideration of tomographic
measurement as a coding problem is here to stay.

ACKNOWLEDGMENT
This project was supported by DARPA DSO through the Knowledge Enhanced
Compressive Measurement program, grant number N66001-11-1-4002.

REFERENCES
1. D. Brady, Optical Imaging and Spectroscopy (Wiley, 2009).
2. C. E. Shannon, “Communications in the presence of noise,” Proc. IRE 37, 10–21
(1949).
3. Y. Chen and Y. Chi, “Robust spectral compressed sensing via structured matrix
completion,” IEEE Trans. Inf. Theory 60, 6576–6601 (2014).
4. G. Tang, B. Bhaskar, P. Shah, and B. Recht, “Compressed sensing off the grid,”
IEEE Trans. Inf. Theory 59, 7465–7490 (2013).
5. A. Wagadarikar, N. Pitsianis, X. Sun, and D. J. Brady, “Video rate spectral
imaging using a coded aperture snapshot spectral imager,” Opt. Express 17,
6368–6388 (2009).
6. J. Hahn, S. Lim, K. Choi, R. Horisaki, and D. J. Brady, “Video-rate compressive
holographic microscopic tomography,” Opt. Express 19, 7289–7298 (2011).
7. S. Mallat, A Wavelet Tour of Signal Processing (Elsevier/Academic, 2009).
8. M. Elad, Sparse and Redundant Representations: From Theory to Applications
in Signal and Image Processing (Springer-Verlag, 2010).
9. A. Papoulis, “Generalized sampling expansion,” IEEE Trans. Circuits Syst. 24,
652–654 (1977).
10. D. Z. Du and F. K. Hwang, Combinatorial Group Testing and Its Applications
(World Scientific, 2000).
11. M. Harwit and N. Sloane, Hadamard Transform Optics (Academic, 1979).
12. D. L. Donoho, “Compressed sensing,” IEEE Trans. Inf. Theory 52, 1289–1306
(2006).
13. E. J. Candes and T. Tao, “Near-optimal signal recovery from random projections:
universal encoding strategies?” IEEE Trans. Inf. Theory 52, 5406–5425 (2006).
806 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

14. E. Candes, J. Romberg, and T. Tao, “Robust uncertainty principles: exact signal
reconstruction from highly incomplete frequency information,” IEEE Trans. Inf.
Theory 52, 489–509 (2006).
15. P. Griffiths and J. De Haseth, “Fourier Transform Infrared Spectrometry,” in
Chemical Analysis: a Series of Monographs on Analytical Chemistry and Its
Applications (Wiley, 2007), Vol. 171.
16. D. L. Marks, R. A. Stack, and D. J. Brady, “Three-dimensional coherence im-
aging in the Fresnel domain,” Appl. Opt. 38, 1332–1342 (1999).
17. S. R. Gottesman and E. E. Fenimore, “New family of binary arrays for coded
aperture imaging,” Appl. Opt. 28, 4344–4352 (1989).
18. E. R. Dowski and W. T. Cathey, “Extended depth of field through wave-front
coding,” Appl. Opt. 34, 1859–1866 (1995).
19. D. L. Marks, R. A. Stack, D. J. Brady, and J. van der Gracht, “Three-dimensional
tomography using a cubic-phase plate extended depth-of-field system,” Opt.
Lett. 24, 253–255 (1999).
20. D. L. Marks, R. A. Stack, D. J. Brady, D. C. Munson, and R. B. Brady, “Visible
cone-beam tomography with a lensless interferometric camera,” Science 284,
2164–2166 (1999).
21. D. J. Brady, “Multiplex sensors and the constant radiance theorem,” Opt. Lett.
27, 16–18 (2002).
22. D. Takhar, J. N. Laska, M. B. Wakin, M. F. Duarte, D. Baron, S. Sarvotham, K. F.
Kelly, and R. G. Baraniuk, “A new compressive imaging camera architecture
using optical-domain compression,” Proc. SPIE 6065, 606509 (2006).
23. D. J. Brady, N. Pitsianis, X. Sun, and P. Potuluri, “Compressive measurement
and signal inference,” U.S. patent 7,283,231 (Oct. 16, 2007).
24. M. Neifeld and P. Shankar, “Feature-specific imaging,” Appl. Opt. 42,
3379–3389 (2003).
25. D. J. Brady, N. Pitsianis, and X. Sun, “Reference structure tomography,” J. Opt.
Soc. Am. A 21, 1140–1147 (2004).
26. S. Ji, Y. Xue, and L. Carin, “Bayesian compressive sensing,” IEEE Trans. Signal
Process. 56, 2346–2356 (2008).
27. R. Baraniuk, V. Cevher, M. Duarte, and C. Hegde, “Model-based compressive
sensing,” IEEE Trans. Inf. Theory 56, 1982–2001 (2010).
28. M. Chen, J. Silva, J. Paisley, C. Wang, D. Dunson, and L. Carin, “Compressive
sensing on manifolds using a nonparametric mixture of factor analyzers: algorithm
and performance bounds,” IEEE Trans. Signal Process. 58, 6140–6155 (2010).
29. Y. Eldar, P. Kuppinger, and H. Bolcskei, “Block-sparse signals: uncertainty rela-
tions and efficient recovery,” IEEE Trans. Signal Process. 58, 3042–3054 (2010).
30. R. Baraniuk and M. Wakin, “Random projections of smooth manifolds,” Found.
Comput. Math. 9, 51–77 (2009).
31. S. Gleichman and Y. C. Eldar, “Blind compressed sensing,” IEEE Trans. Inf.
Theory 57, 6958–6975 (2011).
32. M. Lustig, D. Donoho, and J. M. Pauly, “Sparse MRI: the application of
compressed sensing for rapid MR imaging,” Magn. Reson. Med. 58, 1182–1195
(2007).
33. D. M. Healy and J. B. Weaver, “Adapted wave-form encoding for magnetic-
resonance-imaging,” IEEE Eng. Med. Biol. Mag. 14(5), 621–638 (1995).
34. R. Rangayyan, A. P. Dhawan, and R. Gordon, “Algorithms for limited-view
computed-tomography—an annotated-bibliography and a challenge,” Appl.
Opt. 24, 4000–4012 (1985).
35. O. Lee, J. M. Kim, Y. Bresler, and J. C. Ye, “Compressive diffuse optical tomog-
raphy: noniterative exact reconstruction using joint sparsity,” IEEE Trans. Med.
Imag. 30, 1129–1142 (2011).
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 807

36. M. Süzen, A. Giannoula, and T. Durduran, “Compressed sensing in diffuse op-


tical tomography,” Opt. Express 18, 23676–23690 (2010).
37. D. J. Brady, D. L. Marks, K. P. MacCabe, and J. A. O’Sullivan, “Coded apertures
for x-ray scatter imaging,” Appl. Opt. 52, 7745–7754 (2013).
38. D. J. Brady and D. L. Marks, “Coding for compressive focal tomography,”
Appl. Opt. 50, 4436–4449 (2011).
39. P. Llull, X. Liao, X. Yuan, J. Yang, D. Kittle, L. Carin, G. Sapiro, and D. J. Brady,
“Coded aperture compressive temporal imaging,” Opt. Express 21, 10526–10545
(2013).
40. D. Kittle, K. Choi, A. Wagadarikar, and D. J. Brady, “Multiframe image estimation
for coded aperture snapshot spectral imagers,” Appl. Opt. 49, 6824–6833 (2010).
41. X. Yuan, J. Yang, P. Llull, X. Liao, G. Sapiro, D. J. Brady, and L. Carin,
“Adaptive temporal compressive sensing for video,” arXiv:1302.3446 (2013).
42. A. Walther, “Radiometry and coherence,” J. Opt. Soc. Am. 58, 1256–1259 (1968).
43. A. T. Friberg, “Existence of a radiance function for finite planar sources of ar-
bitrary states of coherence,” J. Opt. Soc. Am. 69, 192–198 (1979).
44. D. C. Youla and H. Webb, “Image restoration by the method of convex projec-
tions: part 1 theory,” IEEE Trans. Med. Imag. 1, 81–94 (1982).
45. M. I. Sezan and H. Stark, “Image restoration by the method of convex projec-
tions: part 2 applications and numerical results,” IEEE Trans. Med. Imag. 1,
95–101 (1982).
46. M. H. Li, H. Q. Yang, and H. Kudo, “An accurate iterative reconstruction algo-
rithm for sparse objects: application to 3D blood vessel reconstruction from a
limited number of projections,” Phys. Med. Biol. 47, 2599–2609 (2002).
47. J. W. Stayman, W. Zbijewski, Y. Otake, A. Uneri, S. Schafer, J. Lee, J. L. Prince,
and J. H. Siewerdsen, “Penalized-likelihood reconstruction for sparse data ac-
quisitions with unregistered prior images and compressed sensing penalties,”
Proc. SPIE 7961, 79611L (2011).
48. K. Choi, J. Wang, L. Zhu, T. S. Suh, S. Boyd, and L. Xing, “Compressed sensing
based cone-beam computed tomography reconstruction with a first-order
method,” Med. Phys. 37, 5113–5125 (2010).
49. E. Y. Sidky and X. C. Pan, “Image reconstruction in circular cone-beam com-
puted tomography by constrained, total-variation minimization,” Phys. Med.
Biol. 53, 4777–4807 (2008).
50. E. Y. Sidky, C. M. Kao, and X. H. Pan, “Accurate image reconstruction from
few-views and limited-angle data in divergent-beam CT,” J. X-Ray Sci. Technol.
14, 119–139 (2006).
51. E. Hansis, D. Schafer, O. Dossel, and M. Grass, “Evaluation of iterative sparse
object reconstruction from few projections for 3-D rotational coronary angiog-
raphy,” IEEE Trans. Med. Imag. 27, 1548–1555 (2008).
52. G. H. Chen, J. Tang, and S. H. Leng, “Prior image constrained compressed sens-
ing: a method to accurately reconstruct dynamic CT images from highly under-
sampled projection data sets,” Med. Phys. 35, 660–663 (2008).
53. J. Song, Q. H. Liu, G. A. Johnson, and C. T. Badea, “Sparseness prior based
iterative image reconstruction for retrospectively gated cardiac micro-CT,”
Med. Phys. 34, 4476–4483 (2007).
54. M. Bertalmio, G. Sapiro, V. Caselles, and C. Ballester, “Image inpainting,” in
Proceedings of the 27th Annual Conference on Computer Graphics and
Interactive Techniques (ACM/Addison-Wesley, 2000), pp. 417–424.
55. R. Ng, M. Levoy, M. Brédif, G. Duval, M. Horowitz, and P. Hanrahan, “Light
field photography with a hand-held plenoptic camera,” Computer Science
Technical Report CSTR 2 (Stanford University, 2005).
56. M. Golay, “Multislit spectroscopy,” J. Opt. Soc. Am. 39, 437–444 (1949).
808 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

57. L. Mertz, Transformations in Optics (Wiley, 1965).


58. W. E. Smith, R. G. Paxman, and H. H. Barrett, “Image-reconstruction from
coded data. 1. Reconstruction algorithms and experimental results,” J. Opt.
Soc. Am. A 2, 491–500 (1985).
59. R. G. Paxman, H. H. Barrett, W. E. Smith, and T. D. Milster, “Image-
reconstruction from coded data. 2. Code design,” J. Opt. Soc. Am. A 2,
501–509 (1985).
60. J. F. Crouzet, “3D coded aperture imaging, ill-posedness and link with incom-
plete data Radon transform,” Inverse Probl. Imag. 5, 341–353 (2011).
61. P. Potuluri, M. B. Xu, and D. J. Brady, “Imaging with random 3D reference
structures,” Opt. Express 11, 2134–2141 (2003).
62. A. Sinha and D. J. Brady, “Size and shape recognition using measurement sta-
tistics and random 3D reference structures,” Opt. Express 11, 2606–2618 (2003).
63. P. K. Agarwal, D. Brady, and J. Matousek, “Segmenting object space by geo-
metric reference structures,” ACM Trans. Sens. Netw. 2, 455–465 (2006).
64. K. Choi and D. J. Brady, “Coded aperture computed tomography,” Proc. SPIE
7468, 74680B (2009).
65. K. MacCabe, K. Krishnamurthy, A. Chawla, D. Marks, E. Samei, and D. Brady,
“Pencil beam coded aperture x-ray scatter imaging,” Opt. Express 20, 16310–
16320 (2012).
66. W. Carson, M. Chen, M. Rodrigues, R. Calderbank, and L. Carin,
“Communications-inspired projection design with application to compressive
sensing,” arXiv:1206.1973 (2012).
67. H. Barrett and K. Myers, Foundations of Image Science (Wiley-Interscience,
2004).
68. E. Candes and J. Romberg, “Sparsity and incoherence in compressive sampling,”
Inverse Probl. 23, 969–985 (2007).
69. M. Bertero and P. Boccacci, Introduction to Inverse Problems in Imaging
(Taylor & Francis, 1998).
70. Y. Kaganovsky, D. Li, A. Holmgren, H. Jeon, K. P. MacCabe, D. G. Politte, J. A.
O’Sullivan, L. Carin, and D. J. Brady, “Compressed sampling strategies for
tomography,” J. Opt. Soc. Am. A 31, 1369–1394 (2014).
71. S. Uttam, N. A. Goodman, M. A. Neifeld, C. Kim, R. John, J. Kim, and
D. Brady, “Optically multiplexed imaging with superposition space tracking,”
Opt. Express 17, 1691–1713 (2009).
72. R. F. Marcia, C. Kim, C. Eldeniz, J. Kim, D. J. Brady, and R. M. Willett,
“Superimposed video disambiguation for increased field of view,” Opt.
Express 16, 16352–16363 (2008).
73. M. E. Gehm, R. John, D. J. Brady, R. M. Willett, and T. J. Schulz, “Single-shot
compressive spectral imaging with a dual-disperser architecture,” Opt. Express
15, 14013–14027 (2007).
74. A. Wagadarikar, R. John, R. Willett, and D. Brady, “Single disperser design for
coded aperture snapshot spectral imaging,” Appl. Opt. 47, B44–B51 (2008).
75. J.-P. Schlomka, A. Harding, U. van Stevendaal, M. Grass, and G. L. Harding,
“Coherent scatter computed tomography: a novel medical imaging technique,”
Proc. SPIE 5030, 256–265 (2003).
76. K. P. MacCabe, A. D. Holmgren, M. P. Tornai, and D. J. Brady, “Snapshot
2D tomography via coded aperture x-ray scatter imaging,” Appl. Opt. 52,
4582–4589 (2013).
77. J. Delfs and J. P. Schlomka, “Energy-dispersive coherent scatter computed
tomography,” Appl. Phys. Lett. 88, 243506 (2006).
78. J. A. Greenberg, K. Krishnamurthy, and D. Brady, “Snapshot molecular imaging
using coded energy-sensitive detection,” Opt. Express 21, 25480–25491 (2013).
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 809

79. J. Greenberg, K. Krishnamurthy, and D. Brady, “Compressive single-pixel snap-


shot x-ray diffraction imaging,” Opt. Lett. 39, 111–114 (2014).
80. M. Slaney and A. Kak, Principles of Computerized Tomographic Imaging
(SIAM, 1988).
81. E. Wolf, “Principles and development of diffraction tomography,” in Trends in
Optics (Academic, 1996), Vol. 3, pp. 83–110.
82. D. C. Munson, Jr., J. D. O’Brien, and W. K. Jenkins, “A tomographic formu-
lation of spotlight-mode synthetic aperture radar,” Proc. IEEE 71, 917–925
(1983).
83. W. H. Carter, “Computational reconstruction of scattering objects from holo-
grams,” J. Opt. Soc. Am. 60, 306–314 (1970).
84. A. Devaney, “A filtered backpropagation algorithm for diffraction tomography,”
Ultrason. Imag. 4, 336–350 (1982).
85. A. Devaney, “Geophysical diffraction tomography,” IEEE Trans. Geosci.
Remote Sens. GE-22, 3–13 (1984).
86. K. Choi, R. Horisaki, J. Hahn, S. Lim, D. L. Marks, T. J. Schulz, and D. J. Brady,
“Compressive holography of diffuse objects,” Appl. Opt. 49, H1–H10 (2010).
87. D. J. Brady, K. Choi, D. L. Marks, R. Horisaki, and S. Lim, “Compressive holog-
raphy,” Opt. Express 17, 13040–13049 (2009).
88. J. Hunt, T. Driscoll, A. Mrozack, G. Lipworth, M. Reynolds, D. Brady, and D. R.
Smith, “Metamaterial apertures for computational imaging,” Science 339,
310–313 (2013).
89. L. C. Potter, E. Ertin, J. T. Parker, and M. Cetin, “Sparsity and compressed sens-
ing in radar imaging,” Proc. IEEE 98, 1006–1020 (2010).
90. W. L. Chan, M. L. Moravec, R. G. Baraniuk, and D. M. Mittleman, “Terahertz
imaging with compressed sensing and phase retrieval,” Opt. Lett. 33, 974–976
(2008).
91. R. Baraniuk and P. Steeghs, “Compressive radar imaging,” in IEEE Radar
Conference, April 2007, pp. 128–133.
92. V. M. Patel, G. R. Easley, D. Healy, and R. Chellappa, “Compressed synthetic
aperture radar,” IEEE J. Sel. Top. Quantum Electron. 4, 244–254 (2010).
93. M. A. Herman and T. Strohmer, “High-resolution radar via compressed sensing,”
IEEE Trans. Signal Process. 57, 2275–2284 (2009).
94. J. H. Ender, “On compressive sensing applied to radar,” Signal Process. 90,
1402–1414 (2010).
95. Y. Yu, A. P. Petropulu, and V. Poor, “Mimo radar using compressive sampling,”
IEEE J. Sel. Top. Quantum Electron. 4, 146–163 (2010).
96. A. C. Gurbuz, J. H. McClellan, and W. R. Scott, Jr., “Compressive sensing for
subsurface imaging using ground penetrating radar,” Signal Process. 89, 1959–
1972 (2009).
97. A. Budillon, A. Evangelista, and G. Schirinzi, “Three-dimensional SAR focus-
ing from multipass signals using compressive sampling,” IEEE Trans. Geosci.
Remote Sens. 49, 488–499 (2011).
98. L. Li, W. Zhang, and F. Li, “Compressive diffraction tomography for weakly
scattering,” arXiv:0904.2695 (2009).
99. Y. Rivenson, A. Stern, and B. Javidi, “Overview of compressive sensing tech-
niques applied in holography,” Appl. Opt. 52, A423–A432 (2013).
100. Y. Rivenson, A. Stern, and B. Javidi, “Compressive Fresnel holography,”
J. Display Technol. 6, 506–509 (2010).
101. M. Marim, E. Angelini, J.-C. Olivo-Marin, and M. Atlan, “Off-axis compressed
holographic microscopy in low-light conditions,” Opt. Lett. 36, 79–81 (2011).
102. M. M. Marim, M. Atlan, E. Angelini, and J.-C. Olivo-Marin, “Compressed sens-
ing with off-axis frequency-shifting holography,” Opt. Lett. 35, 871–873 (2010).
810 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

103. Y. Liu, L. Tian, J. W. Lee, H. Y. H. Huang, M. S. Triantafyllou, and G.


Barbastathis, “Scanning-free compressive holography for object localization
with subpixel accuracy,” Opt. Lett. 37, 3357–3359 (2012).
104. R. Horisaki, J. Tanida, A. Stern, and B. Javidi, “Multidimensional imaging using
compressive Fresnel holography,” Opt. Lett. 37, 2013–2015 (2012).
105. Y. Rivenson, A. Rot, S. Balber, A. Stern, and J. Rosen, “Recovery of partially
occluded objects by applying compressive Fresnel holography,” Opt. Lett. 37,
1757–1759 (2012).
106. C. F. Cull, D. A. Wikner, J. N. Mait, M. Mattheiss, and D. J. Brady, “Millimeter-
wave compressive holography,” Appl. Opt. 49, E67–E82 (2010).
107. D. Wehner, High-Resolution Radar, Artech House Radar Library (Artech House,
1995).
108. M. F. Insana and T. J. Hall, “Parametric ultrasound imaging from backscatter
coefficient measurements: image formation and interpretation,” Ultrason.
Imaging 12, 245–267 (1990).
109. A. Mrozack, M. Heimbeck, D. L. Marks, J. Richard, H. O. Everitt, and D. J.
Brady, “Adaptive millimeter-wave synthetic aperture imaging for compressive
sampling of sparse scenes,” Opt. Express 22, 13515–13530 (2014).
110. J. Goodman, Introduction to Fourier Optics, McGraw-Hill Series in Electrical
and Computer Engineering (McGraw-Hill, 1996).
111. J. W. Goodman, Speckle Phenomena in Optics: Theory and Applications
(Roberts, 2007).
112. A. D. Lanterman, “Statistical radar imaging of diffuse and specular targets
using an expectation-maximization algorithm,” Proc. SPIE 4053, 20–31
(2000).
113. T. Schulz, “Penalized maximum-likelihood estimation of covariance matrices
with linear structure,” IEEE Trans. Signal Process. 45, 3027–3038 (1997).
114. A. Mrozack, K. Krishnamurthy, G. Lipworth, D. Smith, and D. Brady, “Imaging
of diffuse objects with dispersive imagers,” in 38th International Conference on
Infrared, Millimeter, and Terahertz Waves (IRMMW-THz), September 2013.
115. D. L. Marks, T. S. Ralston, S. A. Boppart, and P. S. Carney, “Inverse scattering
for frequency-scanned full-field optical coherence tomography,” J. Opt. Soc.
Am. A 24, 1034–1041 (2007).
116. T. S. Ralston, D. L. Marks, P. S. Carney, and S. A. Boppart, “Inverse scattering
for optical coherence tomography,” J. Opt. Soc. Am. A 23, 1027–1037
(2006).
117. E. Lebed, P. J. Mackenzie, M. V. Sarunic, and F. M. Beg, “Rapid volumetric
OCT image acquisition using compressive sampling,” Opt. Express 18, 21003–
21012 (2010).
118. E. Candes, “The restricted isometry property and its implications for compressed
sensing,” C. R. Math. 346, 589–592 (2008).
119. G. Lipworth, A. Mrozack, J. Hunt, D. L. Marks, T. Driscoll, D. Brady, and D. R.
Smith, “Metamaterial apertures for coherent computational imaging on the
physical layer,” J. Opt. Soc. Am. A 30, 1603–1612 (2013).
120. J. Volakis, Antenna Engineering Handbook, 4th ed. (McGraw-Hill, 2007).
121. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge
University, 1995).
122. M. E. Brezinski, Optical Coherence Tomography: Principles and Applications
(Elsevier, 2006).
123. A. R. Thompson, J. M. Moran, and G. W. Swenson, Jr., Interferometry and
Synthesis in Radio Astronomy (Wiley, 2008).
124. S. Basty, M. A. Neifeld, D. Brady, and S. Kraut, “Nonlinear estimation for
interferometric imaging,” Opt. Commun. 228, 249–261 (2003).
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 811

125. A. A. Wagadarikar, D. L. Marks, K. Choi, R. Horisaki, and D. J. Brady, “Imaging


through turbulence using compressive coherence sensing,” Opt. Lett. 35, 838–
840 (2010).
126. D. L. Marks, R. A. Stack, and D. J. Brady, “Digital refraction distortion correc-
tion with an astigmatic coherence sensor,” Appl. Opt. 41, 6050–6054 (2002).
127. C. Wöhler, 3D Computer Vision (Springer, 2012).
128. D. L. Marks, R. Stack, A. J. Johnson, D. J. Brady, and D. C. Munson, “Cone-
beam tomography with a digital camera,” Appl. Opt. 40, 1795–1805 (2001).
129. J. Sharpe, U. Ahlgren, P. Perry, B. Hill, A. Ross, J. Hecksher-Sørensen, R.
Baldock, and D. Davidson, “Optical projection tomography as a tool for 3D
microscopy and gene expression studies,” Science 296, 541–545 (2002).
130. S. Chaudhuri and A. N. Rajagopalan, Depth from Defocus: A Real Aperture
Imaging Approach (Springer, 1999).
131. W. B. Seales and S. Dutta, “Everywhere-in-focus image fusion using control-
lable cameras,” Proc. SPIE 2905, 227–234 (1996).
132. H. A. Eltoukhy and S. Kavusi, “A computationally efficient algorithm for multi-
focus image reconstruction,” Proc. SPIE 5017, 332–341 (2003).
133. S. T. Li, J. T. Y. Kwok, I. W. H. Tsang, and Y. N. Wang, “Fusing images with
different focuses using support vector machines,” IEEE Trans. Neural Netw. 15,
1555–1561 (2004).
134. S. T. Li and B. Yang, “Multifocus image fusion using region segmentation and
spatial frequency,” Image Vis. Comput. 26, 971–979 (2008).
135. B. Yang and S. T. Li, “Multifocus image fusion and restoration with sparse
representation,” IEEE Trans. Instrum. Meas. 59, 884–892 (2010).
136. S. Hong, J. Jang, and B. Javidi, “Three-dimensional volumetric object
reconstruction using computational integral imaging,” Opt. Express 12, 483–
491 (2004).
137. E. H. Adelson and J. Y. A. Wang, “Single lens stereo with a plenoptic camera,”
IEEE Trans. Pattern Anal. Machine Intell. 14, 99–106 (1992).
138. M. Levoy, “Light fields and computational imaging,” Computer 39, 46–55
(2006).
139. A. Ashok and M. A. Neifeld, “Compressive light field imaging,” Proc. SPIE
7690, 76900Q (2010).
140. S. D. Babacan, R. Ansorge, M. Luessi, P. Ruiz Mataran, R. Molina, and A. K.
Katsaggelos, “Compressive light field sensing,” IEEE Trans. Image Process. 21,
4746–4757 (2012).
141. J. Ojeda-Castaneda, L. R. Berriel-Valdos, and E. L. Montes, “Line-spread func-
tion relatively insensitive to defocus,” Opt. Lett. 8, 458–460 (1983).
142. W. T. Welford, “Use of annular apertures to increase focal depth,” J. Opt. Soc.
Am. 50, 749–753 (1960).
143. W. Chi and N. George, “Electronic imaging using a logarithmic asphere,”
Opt. Lett. 26, 875–877 (2001).
144. A. Greengard, Y. Schechner, and R. Piestun, “Depth from diffracted rotation,”
Opt. Lett. 31, 181–183 (2006).
145. R. Horisaki, Y. Nakao, T. Toyoda, K. Kagawa, Y. Masaki, and J. Tanida,
“A compound-eye imaging system with irregular lens-array arrangement,”
Proc. SPIE 7072, 70720G (2008).
146. K. Venkataraman, D. Lelescu, J. Duparré, A. McMahon, G. Molina, P.
Chatterjee, R. Mullis, and S. Nayar, “PiCam: an ultra-thin high performance
monolithic camera array,” ACM Trans. Graph. 32, 1–13 (2013).
147. A. Kubota, K. Aizawa, and T. Chen, “Reconstructing dense light field from array
of multifocus images for novel view synthesis,” IEEE Trans. Electron. Comput.
16, 269–279 (2007).
812 Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics Review

148. C. Frese and I. Gheta, “Robust depth estimation by fusion of stereo and focus
series acquired with a camera array,” in IEEE International Conference on
Multisensor Fusion and Integration for Intelligent Systems,” September 2006,
pp. 243–248.
149. P. M. Shankar, W. C. Hasenplaugh, R. L. Morrison, R. A. Stack, and M. A.
Neifeld, “Multiaperture imaging,” Appl. Opt. 45, 2871–2883 (2006).
150. M. Shankar, N. P. Pitsianis, and D. J. Brady, “Compressive video sensors using
multichannel imagers,” Appl. Opt. 49, B9–B17 (2010).
151. M. E. Gehm and J. Kinast, “Adaptive spectroscopy: towards adaptive spectral
imaging,” Proc. SPIE 6978, 69780I (2008).
152. P. Llull, X. Yuan, X. Liao, J. Yang, L. Carin, G. Sapiro, and D. J. Brady,
“Compressive extended depth of field using image space coding,” in
Computational Optical Sensing and Imaging (Optical Society of America,
2014), paper CM2D–3.
153. X. Yuan, P. Llull, X. Liao, J. Yang, D. J. Brady, G. Sapiro, and L. Carin, “Low-
cost compressive sensing for color video and depth,” arXiv:1402.6932v1 (2014).
154. C. Zhou, D. Miau, and S. K. Nayar, “Focal sweep camera for space-time refo-
cusing,” Technical Report (Department of Computer Science, Columbia
University, 2012).
155. S. R. P. Pavani, M. A. Thompson, J. S. Biteen, S. J. Lord, N. Liu, R. J. Twieg,
R. Piestun, and W. Moerner, “Three-dimensional, single-molecule fluorescence
imaging beyond the diffraction limit by using a double-helix point spread func-
tion,” Proc. Natl. Acad. Sci. USA 106, 2995–2999 (2009).
156. G. Grover, S. Quirin, C. Fiedler, and R. Piestun, “Photon efficient double-helix
PSF microscopy with application to 3D photo-activation localization imaging,”
Biomed. Opt. Express 2, 3010–3020 (2011).
157. A. Levin, R. Fergus, F. Durand, and W. T. Freeman, “Image and depth from a
conventional camera with a coded aperture,” ACM Trans. Graph. 26, 70 (2007).
158. Y. Bando, B.-Y. Chen, and T. Nishita, “Motion deblurring from a single image
using circular sensor motion,” Comput. Graph. Forum 30, 1869–1878 (2011).
159. M. D. Lew, S. F. Lee, M. Badieirostami, and W. E. Moerner, “Corkscrew point
spread function for far-field three-dimensional nanoscale localization of point-
like objects,” Opt. Lett. 36, 202–204 (2011).
160. A. Agrawal, S. Quirin, G. Grover, and R. Piestun, “Limits of 3D dipole
localization and orientation estimation with application to single-molecule
imaging,” in Imaging and Applied Optics (Optical Society of America, 2011),
paper CWA4.
161. D. Krishnan and R. Fergus, “Fast image deconvolution using hyper-Laplacian
priors,” in Advances in Neural Information Processing Systems (2009),
pp. 1033–1041.
162. A. J. Bell and T. J. Sejnowski, “An information-maximization approach to blind
separation and blind deconvolution,” Neural Comput. 7, 1129–1159 (1995).
163. T. F. Chan and C.-K. Wong, “Total variation blind deconvolution,” IEEE Trans.
Image Process. 7, 370–375 (1998).

David J. Brady is the Fitzpatrick Professor of Photonics at Duke


University, where he leads the Duke Imaging and Spectroscopy
Program. Brady is a Fellow of The Optical Society (OSA), SPIE,
and IEEE and was awarded the 2013 SPIE Dennis Gabor Award.
He is the author of Optical Imaging and Spectroscopy (OSA/
Wiley, 2009). Brady’s research focuses on computational imaging.
He is a graduate of Macalester College and the California Institute
of Technology.
Review Vol. 7, No. 4 / December 2015 / Advances in Optics and Photonics 813

Alex Mrozack is a research scientist at Morpho Detection, Inc. He


earned his Ph.D. in the Duke Imaging and Spectroscopy Program
at Duke University, where he worked on millimeter wave imaging
systems. His undergraduate degree is from Rice University.

Kenneth MacCabe is a computational physicist with experience


in x-ray imaging, optics, precision orbit determination, and atomic
physics. He earned his Ph.D. in Physics from the University of
North Carolina (UNC) Chapel Hill and was a visiting graduate
student at Duke University, prior to which he worked and studied
at MIT. Ken also holds a MS Physics from UNC Chapel Hill and a
BS Physics and BS Mathematics from Washington State University.

Patrick Llull is a doctoral candidate in the Duke University


Imaging and Spectroscopy Program. His areas of interest are com-
putational and compressive imaging, with foci on compressive
video, multidimensional sensing and displays, and focal tomogra-
phy. Patrick graduated with his BS from the University of Arizona
College of Optical Sciences in 2012.

You might also like