搜索
bottom↓
回复: 0

RAISR: Rapid and Accurate Image

[复制链接]

出0入0汤圆

发表于 2017-7-26 11:13:45 | 显示全部楼层 |阅读模式
1
RAISR: Rapid and Accurate Image
Super Resolution
Yaniv Romano, John Isidoro, and Peyman Milanfar, Fellow, IEEE
Abstract
Given an image, we wish to produce an image of larger size with significantly more pixels and higher image
quality. This is generally known as the Single Image Super-Resolution (SISR) problem. The idea is that with sufficient
training data (corresponding pairs of low and high resolution images) we can learn set of filters (i.e. a mapping) that
when applied to given image that is not in the training set, will produce a higher resolution version of it, where the
learning is preferably low complexity. In our proposed approach, the run-time is more than one to two orders of
magnitude faster than the best competing methods currently available, while producing results comparable or better
than state-of-the-art.
A closely related topic is image sharpening and contrast enhancement, i.e., improving the visual quality of a
blurry image by amplifying the underlying details (a wide range of frequencies). Our approach additionally includes
an extremely efficient way to produce an image that is significantly sharper than the input blurry one, without
introducing artifacts such as halos and noise amplification. We illustrate how this effective sharpening algorithm, in
addition to being of independent interest, can be used as a pre-processing step to induce the learning of more effective
upscaling filters with built-in sharpening and contrast enhancement effect.
I. INTRODUCTION
Single Image Super Resolution (SISR) is the process of estimating a High-Resolution (HR) version of a LowResolution (LR) input image. This is a well-studied problem, which comes up in practice in many applications,
such as zoom-in of still and text images, conversion of LR images/videos to high definition screens, and more. The
linear degradation model of the SISR problem is formulated by
z = DsHx; (1)
where z 2 RM×N is the input image, x 2 RMs×Ns is the unknown HR image, both are held in lexicographic
ordering. The linear operator H 2 RMNs2×MNs2 blurs the image x, followed by a decimation in a factor of s
in each axis, which is the outcome of the multiplication by Ds 2 RMN×MNs2. In the SR task, the goal is to
More results can be found in the supplementary material: https://goo.gl/D0ETxG. This work was performed while the first author was an
intern at Google Research. Y. Romano is with the Department of Electrical Engineering, Technion – Israel Institute of Technology, Technion
City, Haifa 32000, Israel. E-mail address: yromano@tx.technion.ac.il. J. Isidoro and P. Milanfar are with Google Research, 1600 Amphitheatre
Pkwy, Mountain View, CA 94043. E-mail addresses: fisidoro,milanfarg@google.com.
arXiv:1606.01299v3 [cs.CV] 4 Oct 20162
recover the unknown underlying image x from the known measurement z. Note that, in real world scenarios, the
degradation model can be non-linear (e.g. due to compression) or even unknown, and may also include noise.
The basic methods for upscaling a single image are the linear interpolators, including the nearest-neighbor, bilinear
and bicubic [1], [2]. These methods are widely used due to their simplicity and low complexity, as the interpolation
kernels (upscaling filters) are not adaptive to the content of the image. However, naturally, these linear methods are
limited in reconstructing complex structures, often times result in pronounced aliasing artifacts and over-smoothed
regions. In the last decade powerful image priors were developed, e.g., the self-similarity [3]–[6], sparsity [7]–[12],
and Gaussian Mixtures [13], leading to high quality restoration with the cost of increased complexity.
In this paper we concentrate on example-based methods [8], [9], [11], [14]–[18], which have drawn a lot of
attention in recent years. The core idea behind these methods is to utilize an external database of images and learn
a mapping from LR patches to their HR counterparts. In the learning stage, LR-HR pairs of image patches are
synthetically generated, e.g., for 2× upscaling, a typical size of the HR patch is 6×6 and the one of the synthetically
downscaled LR patch is 3 × 3. Then, the desired mapping is learned and regularized using various local image
priors.
The sparsity model is one such prior [8], [9], where the learning mechanism results in a compact (sparse)
representation of pairs of LR and HR patches over learned dictionaries. Put differently, per each LR patch, these
methods construct a non-linear adaptive filter (formulated as a projection matrix), which is a combination of a few
basis elements (the learned dictionary atoms) that best fit to the input patch. Applying the filter that is tailored to
the LR patch leads to the desired upscaling effect.
The Anchored Neighborhood Regression (ANR) [10] keeps the high quality reconstruction of [8] and [9] while
achieving a significant speed-up in runtime. This is done by replacing the sparse-coding step that computes the
compact representation of each patch over the learned dictionaries, with set of pre-computed projection matrices
(filters), which are the outcome of ridge regression problems. As such, at runtime, instead of applying sparse-coding,
ANR suggest searching for the nearest atom to the LR patch, followed by a multiplication by the corresponding
pre-computed projection matrix. A follow-up work, called A+ [11], improves the performance of ANR by learning
regressors not only from the nearest dictionary atoms, but also from the locally nearest training samples, leading
to state-of-the-art restoration.
SRCNN [16] is another efficient example-based approach that builds upon deep Convolutional Neural Network
(CNN) [19], and learns an end-to-end mapping from LR images to their HR counterparts. Note that, differently
from sparsity-based techniques, SRCNN does not explicitly learn the dictionaries for modeling the patches. In this
case, the model is implicitly learned by the hidden convolutional layers.
The above mentioned SISR methods result in impressive restoration, but with the cost of (relatively) high
computational complexity. In this paper we suggest a learning-based framework, called RAISR, which produces high
quality restoration while being two orders of magnitude faster than the current leading algorithms, with extremely
low memory requirements.
The core idea behind RAISR is to enhance the quality of a very cheap (e.g. bilinear) interpolation method by3
applying a set of pre-learned filters on the image patches, chosen by an efficient hashing mechanism. Note that the
filters are learned based on pairs of LR and HR training image patches, and the hashing is done by estimating the
local gradients’ statistics. As a final step, in order to avoid artifacts, the initial upscaled image and its filtered version
are locally blended by applying a weighted average, where the weights are a function of a structure descriptor. We
harness the Census Transform (CT) [20] for the blending task, as it is extremely fast and cheap descriptor of the
image structure which can be utilized to detect structure deformations that occur due to the filtering step.
A closely related topic to SISR is image sharpening, aiming to amplify the structures/details of a blurry image. The
basic sharpening techniques apply a linear filter on the image, as in the case of unsharp masking [21] or Difference
of Gaussians (DoG) [22], [23]. These techniques are highly effective in terms of complexity, but tend to introduce
artifacts such as over-sharpening, gradient reversals, noise amplification, and more. Similarly to SISR, improved
results can be obtained by relying on patch priors, where the sensitivity to the content/structure of the image is
the key for artifact-free enhancement [24]–[28]. For example, with the cost of increased complexity compared to
the linear approach, the edge-aware bilateral filter [29], [30], Non-Local Means [3] and guided filter [25] produce
impressive sharpening effect.
As a way to generate high-quality sharp images, one can learn a mapping from LR images to their sharpened
HR versions, thus achieving a built-in sharpening/contrast-enhancement effect ”for free”. Furthermore, the learning
stage is not limited to a linear degradation model (as in Eq. (1)), as such, learning a mapping from compressed
LR images to their sharpened HR versions can be easily done, leading to an ”all in one” mechanism that not only
increases the image resolution, but also reduces compression artifacts and enhances the contrast of the image.
Triggered by this observation, we develop a sharpener as well, which is of independent interest. The proposed
sharpener is highly efficient and able to enhance both fine details (high frequencies) and the overall contrast of the
image (mid-low frequencies). The proposed method has almost similar complexity to the linear sharpeners, while
being competitive with far more complex techniques. The suggested sharpener is based on applying DoG filters
[22], [23] on the image, which are capable to enhance a wide range of frequencies. Next, a CT-based structure-aware
blending step is applied as a way to prevent artifacts due to the added content-aware property (similar mechanism
to the one suggested in the context of SISR).
This paper is organized as follows: In Section II we describe the global learning and upscaling scheme, formulating
the core engine of RAISR. In Section III we refine the global approach by integrating the initial upscaling kernel
to the learning scheme. In Section IV we describe the overall learning and upscaling framework, including the
hashing and blending steps. The sharpening algorithm is detailed in Section V. Experiments are brought in Section
VI, comparing the proposed upscaling and sharpening algorithm with state-of-the-art methods. Conclusions and
future research directions are given in Section VII.
II. FIRST STEPS: GLOBAL FILTER LEARNING
Given an initial (e.g. bilinear in our case) upscaled versions of the training database images, yi 2 RM×N , with
i = 1; · · · ; L, we aim to learn a d × d filter h that minimizes the Euclidean distance between the collection fyig4
(a) Learning Stage
(b) Upscaling Stage
Fig. 1. The basic learning and application scheme of a global filter that maps LR images to their HR versions.
and the desired training HR images fxig. Formally, this is done by solving a least-squares minimization problem
min
h
LX i
=1
kAih − bik2 2 (2)
where h 2 Rd2 denotes the filter h 2 Rd×d in vector-notation; Ai 2 RMN×d2 is a matrix, composed of patches of
size d×d, extracted from the image yi, each patch forming a row in the matrix. The vector bi 2 RMN is composed
of pixels from xi, corresponding to the center coordinates of yi patches. The block diagram, demonstrating the
core idea of the learning process is given in Fig. 1a.
In practice, the matrix A can be very large, so we employ two separate approaches to control the computational
complexity of estimating the filter. First, in general not all available patches needs to be used in order to obtain a
reliable estimate. In fact, we typically construct Ai and bi by sampling K patches/pixels from the images on a
fixed grid, where K  MN. Second, the minimization of the least-squares problem, formulated in Eq. (2), can be
recast in a way that significantly reduces both memory and computational requirements. To simplify the exposition,
the following discussion is given in the context of filter learning based on just one image, but extending the idea
to several images and filters is trivial. The proposed approach results in an efficient solution for the learning stage
where the memory requirements are only on the order of the size of the learned filter. The solution is based on the
observation that instead of minimizing Eq. (2), we can minimize
min
h
kQh − Vk2 2; (3)
where Q = AT A and V = AT b.
Notice that Q is a small d2 × d2 matrix, thus requiring relatively little memory. The same observation is valid
for V that requires less memory than holding the vector b. Furthermore, based on the inherent definition of matrixmatrix and matrix-vector multiplications, we in fact avoid holding the whole matrix (and vector) in memory. More
specifically, Q can be computed cumulatively by summing chunks of rows (for example sub matrices Aj 2 Rq×d2,
q  MN), which can be multiplied independently, followed by an accumulation step; i.e.
Q = AT A = X
j
AT
j Aj (4)5
Fig. 2. Bilinear upscaling by a factor of 2 in each axis. There are four types of pixels, denoted by P1-P4, corresponding to the four kernels
that are applied during the bilinear interpolation.
The same observation is true for matrix-vector multiplication
V = AT b = X
j
AT
j bj; (5)
where bj 2 Rq is a portion of the vector b, corresponding to the matrix Aj. Thus, the complexity of the proposed
learning scheme in terms of memory is very low – it is in the order of the filter size. Moreover, using this observation
we can parallelize the computation of AT j Aj and AT j bj, leading to a speedup in the runtime. As for the least
squares solver itself, minimizing Eq. (3) can be done efficiently since Q is a positive semi-definite matrix, which
perfectly suits a fast conjugate gradients solver [31].
To summarize, the learning stage is efficient both in terms of the memory requirements and ability to parallelize. As
displayed in Fig. 1b, at run-time, given a LR image (that is not in the training set), we produce its HR approximation
by first interpolating it using the same cheap upscaling method (e.g. bilinear) that is used in the learning stage,
followed by a filtering step with the pre-learned filter.
III. REFINING THE CHEAP UPSCALING KERNEL: DEALING WITH ALIASING
The ”cheap” upscaling method we employ as a first step, can be any method, including a non-linear one. However,
in order to keep the low complexity of the proposed approach, we use the bilinear interpolator as the initial upscaling
method1. Inspired by the work in [15], whatever the choice of the initial upscaling method, we make the observation
that when aliasing is present the input LR image, the output of the initial upscaler will generally not be shift-invariant
to this aliasing.
As illustrated in Fig. 2, in the case of upscaling by a factor of 2 in each axis, the interpolation weights of the
bilinear kernel vary according to the pixel’s location. As can be seen, there are four possible kernels that are applied
1We also restrict the discussion mainly to the case of 2× upscaling to keep the discussion straightforward. Extensions will be discussed at
the end of this section.6
Fig. 3. Spatially varying learning scheme of four global filters, taking into consideration the internal structure of the bilinear.
(a) P1-Filter (b) P2-Filter (c) P3-Filter (d) P4-Filter
(e) P1-Magnitude spectrum (f) P2-Magnitude spectrum (g) P3-Magnitude spectrum (h) P4-Magnitude spectrum
Fig. 4. Visualization of the four global filters, corresponding to P1-P4 type of pixels, in the pixel domain (a-d), along with their magnitude
in the frequency domain (e-f), where the warmer the color, the larger the value. The filters are learned on Fig. 2 image.
on the LR image according to the type of the pixel, denoted by P1-P4. Since a convolution of two linear filters can
be unified into one filter (in our case, the first is the bilinear and the second is the pre-learned one)2, we should
learn four different filters, corresponding to the four possible types of pixels, as demonstrated in Fig. 3.
The importance of this observation is illustrated in Fig. 4, which plots examples of actual learned filters, along
with their magnitude in the frequency domain. The obtained filters act like bandpass filters, amplifying the midfrequencies, and suppressing the high-frequencies (which contain aliasing components) of the interpolated image.
The learned filters have similar magnitude response (Fig. 4e-4h), but different phase response (Fig. 4a-4d), standing
2This observation is a promising way to further speed up the algorithm and reduce the overall complexity.7
Fig. 5. Applying the four spatially varying pre-learned filters on a LR image.
in agreement with the four different shifted versions of the interpolation kernels.
On the application side, similarly to the core/naive upscaling idea, we first upscale the LR image using the bilinear
interpolator. Then, differently from the naive approach, we apply the pre-learned filters according to the type of
the pixel, followed by an aggregation step that simply combines the outcome of the filtered patches (resulting in a
pixel) to an image. This process is illustrated in Fig. 5.
Notice that a similar observation holds for upscaling by any other integer factor s. For example, upscaling by
a factor of 3 implies that we should learn 9 filters, one per each pixel-type. Similarly, when upscaling by a factor
of 4, there are 16 types of pixels. As already mentioned, in order to keep the flow of the explanations, we will
concentrate on the 2× scenario since the generalization to other scaling factors is straightforward.
IV. RAISR: HASHING-BASED LEARNING AND UPSCALING
Generally speaking, the global image filtering is fast and cheap, as it implies the application of one filter per
patch. Since the learning scheme reduces the Euclidean distance between the HR and the interpolated version
of the LR images, the global filtering has the ability to improve the restoration performance of various linear
upscaling methods. However, the global approach described so far is weaker than the state-of-the-art algorithms,
e.g., sparsity-based methods [8]–[11] or the neural networks based ones [16] that build upon large amount of
parameters, minimizing highly nonlinear cost functions. In contrast to these methods, the global approach is not
adaptive to the content of the image, and its learning stage estimates only a small amount of parameters.
Adaptivity to the image content can be achieved by dividing the image patches into clusters, and constructing
an appropriate filter per each cluster (e.g. as done in [10], [11]). However, the clustering implies the increase
of the overall complexity of the algorithm, which is an outgrowth that we want to avoid. Therefore, instead of
applying ”expensive” clustering (e.g. K-means [32], GMM [33], [34], dictionary learning [8], [9], [11], [14]–[18]),
we suggest using an efficient hashing approach, leading to adaptive filtering that keeps the low complexity of the
linear filtering. More specifically, the local adaptivity is achieved by dividing the image patches into groups (called
”buckets”) based on an informative and ”cheap” geometry measures, which utilize the statistics of the gradients (a8
(a) Learning Stage
(b) Upscaling Stage
Fig. 6. Hashing based learning and upscaling schemes. We suggest dividing the patches into ”buckets”, where each bucket contains patches
with similar geometry (can be considered as a cheap clustering method). Then, a least squares fitting is applied per each bucket and possible
shift. At run-time the hash-table key is computed per each patch, leading to the corresponding pre-learned locally adaptive filters.
detailed description is given in Section IV-A). Then, similarly to the global approach, we also learn four filters,
but this time per each bucket. As a consequence, the proposed learning scheme results in a hash-table of filters,
where the hash-table keys are a function of the local gradients, and the hash-table entries are the corresponding
pre-learned filters. An overview of the proposed hashing-based learning is shown in Fig. 6a.
Given the hash-table, containing filters per quantized edge-statistic descriptor (more details in Section IV-A), the
upscaling procedure becomes very effective. Following Fig. 6b, we compute the hash-table key per each patch of
the initial interpolated image, pointing to the relevant filters (four filters, one per patch-type), to be applied on the
corresponding patch.
Similarly to the global learning process (see Section II), we utilize the matrix-matrix and matrix-vector multipli-9
cations once again. Per each bucket q, we learn a filter hq by minimizing the following cost function
min
h
q
kAT q Aqhq − AT q bqk2 2; (6)
where A
q and bq are the patches and pixels that belong to the q-th bucket. In this case, the low memory requirements
of the proposed learning process are crucial, especially for large hash-table that requires millions of examples to
produce a reliable estimate for the filters. As a consequence, by utilizing the observation described in Section II,
we perform a sub-matrix accumulation on a sub-image block basis, leading to a learning process that can handle
any desired number of examples.
A. Hash-Table Keys: Local Gradient Statistics (Angle, Strength, Coherence)
Naturally, there are many possible local geometry measures that can be used as the hash-table keys, whereas
the statistics of the gradients has a major influence on the proposed approach. We suggest evaluating the local
gradient characteristics via eigenanalysis [35], which yields the gradient’s angle and information about the strength
and coherence of the nearby gradients. Eigenanalysis also helps in cases of thin lines, stripes and other scenarios
that the mean gradient might be zero, yet the neighborhood exhibits a strong directionality.
The direction, strength and coherence are computed by utilizing the pn×pn surroundings of each pixel, i.e., for
the k-th pixel we consider all the pixels that are located at k1; :::; kn. The basic approach starts with a computation
of 2 × n matrix, composed from the horizontal and vertical gradients, gx and gy, of the surroundings of the k-th
pixel, expressed by
Gk =
26664
gxk1 gyk1
...
...
gxkn gykn
37775
: (7)
As stated in [35], the local gradient statistics can be computed using the Singular Value Decomposition (SVD) of
this matrix. The right vector corresponds to the gradient orientation, and the two singular values indicate the strength
and spread of the gradients. Since the work is being performed per-pixel, we hereby focus on efficiency. We can
compute those characteristics more efficiently using an eigen-decomposition of GT k Gk which is a 2 × 2 matrix,
which can be computed conveniently in a closed form. Moreover, in order to incorporate a small neighborhood of
gradient samples per pixel, we employ a diagonal weighting matrix Wk, constructed using a separable normalized
Gaussian kernel.
Following [35], the eigenvector φk 1, corresponding to the largest eigenvalue of GT k WkGk, can be used to derive
the angle of the gradient θk, given by
θk = arctan(φk 1;y; φk 1;x): (8)
Notice that due to the symmetry, a filter that corresponds to the angle θk is identical to the one corresponding to
θk + 180◦.
As shown in [35], the square root of the largest eigenvalue λk 1 is analogous to the ”strength” of the gradient.
The square root of the smaller eigenvalue λk 2 can be considered as the ”spread” of the local gradients, or rather10
(a) 2× upscaling filters
(b) 3× upscaling filters
(c) 4× upscaling filters
Fig. 7. Visualization of the learned filter sets for (a) 2×, (b) 3× and (c) 4× upscaling, learned from using an angle, strength, and coherence
based hashing scheme. Per each subset of filters, the angle varies from left to right; the top, middle, and bottom 3 rows correspond to low,
medium and high coherence. Within each set of 3 rows, gradient strength increases from top to bottom. As can be inferred, the general trend
is that as coherence increases, the directionality of the filter increases. Also, as strength increases the intensity of the filter increases. Notice
how the 3× and 4× upscaling filters are not simply scaled versions of the 2× filters, but also have extracted additional information from the
training data.11
how much they vary in direction. Both of these can be measured in units of intensity. The two eigenvalues can be
combined into a unitless measure known as ”coherence” [35]. The coherence value µk ranges from 0 to 1, and
formulated as
µk =
pλk 1 − pλk 2
pλk 1 + pλk 2 : (9)
Strength and coherence are very useful for detecting a variety of different local image properties. A low strength
and low coherence often signifies a lack of image structure, and usually corresponds to noise or compression
artifacts. High strength, but low coherence often indicates corners or other multi-directional structure. Having a
high coherence is generally an edge, or series of stripes in the same direction, with the strength measuring the
relative intensity of the stripes. Intuitively, strength and coherence allow us to detect semantically different local
image properties, so by using them as part of a hash enables the filter learning process to adapt to these conditions.
As such, combining angle θk, strength λk 1, and coherence µk into a hash function, as detailed in Algorithm 1, can
produce a family of learned filters that are able to handle a variety of situations.
Algorithm 1: Computing the hash-table keys.
Inputs
1: Initial interpolated version of the LR image.
2: Qθ – Quantization factor for angle (e.g. 24).
3: Qs – Quantization factor for strength (e.g. 3).
4: Qµ – Quantization factor for coherence (e.g. 3).
Output
1: Hash-table keys per pixel, denoted by θk, λk 1, and µk.
Process
• Compute the image gradients
• Construct the matrix GT
k WkGk, and obtain the gradients’ angle θk, strength λk 1, and coherence µk
• Quantize: θi l Qθiθ m λi 1 l Qλi 1s m µi l Qµµi m, where d·e is the ceiling function
In Fig. 7, one can see that bucketing by angle, coherence and strength produces a wide variety of filters. The
ones that correspond to low coherency and strength tend to be bandpass and directionally invariant in nature. As
coherency increases, so does the directionality of the filter, smoothing orthogonal to the gradient, but strongly
sharpening in the direction of the gradient.
B. Using Patch Symmetry for Nearly-Free 8× More Learning Examples
The amount of data needed for effective and stable learning of filter sets can be large. For instance, in practice,
it takes at least 105 patches to reliably learn a given filter of size 9 × 9 or 11 × 11. Say we decompose the patches12
Fig. 8. Rebucketing for x-flips, y-flips, and xy-swaps. For example bucket ’B’ accounts for gradient angles between 45◦ and 90◦. When the
patch is flipped on the x-axis, the bucketing also needs to be reflected, as shown in the second graph. The third graph shows y-flips, for which
the subsequent modulo 180◦ operation maps back to the 0◦ to 180◦ range. The final diagram shows xy-swaps (flip around the x=y line). The
different possible combinations of x-flip, y-flip, and xy-swaps account for the 8 different patch transformations.
into B buckets, when using a hashed set of filters, this implies that we need 105 patches per bucket. However,
reaching this amount using real world training data is not as simple as using 105 × B patches.
The problem emerges from the observation that certain hash values occur much more commonly than other
hashes. There are often many more horizontal and vertical structures in imagery, and flat regions (such as sky, and
painted surfaces) are common. Intuitively, these common structures result in the more common hashes.
In order to make the patches that hash to uncommon hash values more effective, we leverage the patch symmetry
and increase the learning power of each patch. More specifically, we can generate 8 different example patches;
four 90◦ rotations, and four mirrored 90◦ rotations. Since 8 patches are generated for each original patch, we are
effectively incorporating 8 times as much information for learning. Since each transformation is a rotation and
mirroring, the transformed patches often belong to a different hash bucket and shift. For example, a patch that is
rotated by 90◦ changes the hash angle bucket by 90◦. A visual demonstration of the proposed idea is given in
Fig. 8, which demonstrates a simplified hashing scheme with 4 angular buckets on a polar graph, and how they get
rebucketed from x-flips, y-flips, and xy-swaps.
Moreover, the patch transformations do not need to be performed as actual image transformations of each incoming
patch, which would be expensive. If the gradient angle dependent hash bucket boundaries are symmetric to flips
in x, y and xy-swaps, the accumulation for transformed patches can be performed. Achieving this symmetry is a
function of the hashing function, and, in our case, having a number of angle buckets that is divisible by 4 satisfies
this requirement. As such, we can accumulate the per bucket and per pixel-type matrices AT q Aq and AT q b across all
training samples, as suggested in Eq. (6). Then, the symmetry can be applied as one last accumulation of permuted
matrices, which act as a set of symmetry augmented matrices. In practice, the additional accumulation step to enable
symmetry takes less than 0.1% of the learning runtime (only a few additional seconds on a 3.4GHz 6-Core Xeon
desktop computer).13
C. Built-in Suppression of Compression Artifacts and Sharpening Effect
The linear degradation model that assumes blur and decimation, as expressed in Eq. (1), is very common in the
literature, but less so in the real world. For example, often times the measured images are blurred with unknown
kernel, compressed, post-processed (e.g. by applying gamma correction), contaminated by noise, and more.
Learning a mapping that is capable to handle highly non-linear degradation model can be done by RAISR. We
found that an effective suppression of compression artifacts is achieved by learning a mapping from compressed LR
images to their uncompressed HR versions. Notice that an inherent compression parameter is the bit-rate/compression
quality, affecting the outcome of the learning scheme. For example, JPEG encoders use a quality level parameter,
which varies from 0 (the worst quality) up to 100 (the best quality). Our experiments show that an aggressive
compression (e.g. 80) indeed suppresses the compression artifacts, but can lead to smoothed result. We also found
that a moderate compression level (e.g. 95) in training helps to suppress the aliasing in addition to alleviating
moderate compression artifacts.
In the same spirit, gaining sharpening effect can be done by learning a mapping from LR training images to their
sharpened HR versions. The stronger the sharpening during the training phase, the sharper the outcome of RAISR
upscaling. We should emphasize that at runtime, we simply apply the pre-learned filters (with possibly built-in
sharpening effect); we do not apply a separate sharpening step.
To conclude, by applying compression and sharpening as pre-processing steps, the learned filters are capable to
map input compressed LR image to a sharpened HR output. As such, by choice, RAISR not only estimates the
missing spatial information, but also suppresses the compression artifacts and amplifies the underlying signal.
D. Blending: An Efficient Structure-Preserving Solution
The proposed learning scheme results in adaptive upscaling filters, having built-in suppression of compression
artifacts and sharpening effect. Concentrating on the sharpening property, two well known side-effects that exist
are halos that appear along edges and the noise amplification. Put differently, applying the pre-learned filters on
the initial interpolated image can lead to structure deformations due to the sharpening property.
As a way to avoid a significant modification in structure, we suggest measuring the local change in structure
that occurred due to the filtering step, and blend accordingly. In areas that the structure of the initial interpolated
image and the filtered one is somewhat similar – we choose the filtered version, while in areas that a major change
is occurred by the filtering step – we choose the initial upscaled image. This suggestion relies on the observation
that the cheap interpolated image typically does an adequate job on areas of the image that contain low spatial
frequencies (e.g. flat regions). On the other hand, the higher spatial frequencies that need to be reconstructed require
the careful treatment of the estimated filters. The blending therefore combines the most appropriate contributions
from the cheap upscaled and the RAISR filtered image to yield the final result. One could have identified these
regions ahead of time by clustering and apply different treatment, however this would have resulted in slower
execution. In what follows, we present a fast alternative that works on two output images for point-wise blending.14
a d g
b e h
c f i
(a)
e > a e > d e > g
e > b – e > h
e > c e > f e > i
(b)
100 80 90
105 95 85
110 120 170
(c)
0 1 1
0 − 1
0 0 0
(d)
8 bit string
a b c d f g h i
0 0 0 1 0 1 1 0
(e)
Fig. 9. Census transform. (a) 3x3 window of pixels, (b) Boolean comparisons between the center pixel and its neighbors, (c)-(e) Numerical
example: (c) Intensity values, (d) The outcome of the boolean comparisons, (e) Census result, an 8 bit string that measures the local structure.
Fig. 10. RAISR upscaling that allows the amplification of high-frequencies only. Applying the pre-leaned filters and avoiding halos and noise
amplification. The blending select the filtered pixels in structured areas, and the cheap upscaled pixels in flat areas.
Inspired by the CT descriptor [20], we suggest using its outcome as an engine that detect structure deformations
and revert the errors of the upscaler. In order to understand the blending mechanism, a brief overview about the CT
is given here. This transform maps the intensity values of the pixels within a small squared region (e.g. of size 3×3)
to a bit string that captures the image structure. The CT is based on the relative ordering of local intensity values,
and not on the intensity values themselves. Following Fig. 9, the center pixel’s intensity value (central cell in Fig.
9a and 9c) is replaced by a bit string descriptor of length 8 (Fig. 9e), composed of a set of boolean comparisons
between the center pixel and its 3 × 3 neighborhood pixels (Fig. 9b, 9d). Note that, in practice, when applying the
comparisons we allow small variations in the intensity values, controlled by a threshold.
Back to RAISR, we suggest two different blending schemes, given in Fig. 10 and Fig. 11, which result in
different enhancement effects (as illustrated in Sections VI-A and VI-B, respectively). The first scheme allows the
pre-learned filters to enhance only the high-frequencies of the image, leading to HR images that look natural, as
required by the conventional SISR problem (a demonstration is given in Fig. 14). While the second scheme allows
the enhancement of a wide range of frequencies, leading to better looking images due to the contrast enhancement15
Fig. 11. RAISR upscaling that allows the amplification of wide range of frequencies, enabling a contrast enhancement effect. Applying the
pre-leaned filters and avoiding halos and noise amplification. The blending mechanism enables efficient structure and contrast enhancement, i.e.,
the amplification of a wide range of frequencies.
effect (as shown in Fig. 17). Note that both versions aim at avoiding artifacts and structure deformations.
We start with the first blending mechanism, which enables the amplification of the high-frequencies only without
modifying the low/mid frequencies, as usually done by the conventional SISR algorithms. Following Fig. 10, our
suggestion is based on the observation that in flat areas (or generally low frequency areas), a linear upscaler produces
good results because there are no fine details to be recovered or aliasing to be suppressed, thus there is no need
to further improve the results in these areas. On the other hand, the linear interpolation fails to recover structured
areas, where the proposed pre-learned filters play the key role. Moreover, within the structured areas, especially
along strong edges, the pre-learned filters may introduce halos due to their size (11×11 or 9×9) and the sharpening
property.
For the sake of completeness, let us explain briefly how the CT, which is indeed blind to the illumination, can be
formulated as a mechanism for edge/structure-detection, thus allowing us to amplify only the high-frequencies of the
image. In this case, the blending weights are the outcome of the so called ”randomness” measure, which indicates
how likely a pixel is in a structured area. Specifically, the size of the Least Connected Component (LCC) of the
CT descriptor (in the case of Fig. 9d the LCC size is 3) is translated to a weight, determining the strength/amount
of the structure within the descriptor window. In general, the larger the size of LCC the higher the weight. Put
differently, by measuring the ”randomness” of the bit string we can infer whether the pixel is a part of an edge or not,
forming the blending weights map. A block diagram of the proposed upscaling scheme, which allows sharpening
of high-frequencies only, is given in Fig. 10.
Although the conventional SISR algorithms result in HR images that look natural, contrast enhancement (i.e.
amplification of low, mid, and high frequencies) often times can lead to better looking images (a visual demonstration16
is given in Fig. 17). This observation leads us to propose the second CT-based blending scheme. According to Fig.
11, in order to measure the local change in structure, we
(i) Compute the CT of the initial upscaled image and the filtered image, and then
(ii) Per each pixel – count the number of bits that were changed, i.e., evaluate the Hamming distance, where the
larger the distance, the larger the change in structure.
As a consequence, by translating the number of bits that were changed to weights, we form the desired blending
map. Notice that the CT is blind to the intensity value itself. Therefore, differently from the randomness strategy
(see Fig. 11), in this case, the obtained blending map allows a local change in the intensity (or contrast), while
avoiding major changes in the structure.
In the learning stage, the target HR images are pre-processed by the proposed DoG sharpener (a detailed
explanation about the sharpener is given in Section V), which sharpens the structures and improves the overall
contrast of the image. As a result, a built-in enhancement of both details (high-frequencies) and contrast (mid/midlow frequencies) is achieved when applying the pre-learned filters. A block diagram of the upscaling scheme that
allows contrast enhancement (i.e. amplification of a wide range of frequencies) is shown in Fig. 11.
Our experiments show that when enhancing only the high-frequencies we obtain images that look natural, having
the same contrast of the LR ones (see Fig. 14). When enhancing a wider range of frequencies (i.e. allowing contrast
change), RAISR produces better looking images (as illustrated in Fig. 17), but there is no guarantee that the output
will have the same nature of the one of the LR. If we use the versions that clean compression artifacts, do sharpening
and contrast enhancement, the effect on the PSNR or SSIM comparisons will not be very clear anymore. So in
terms of this quantitative measure, we might observe deterioration, even if the images look excellent (even better
than the originals!).
To conclude, we introduced the RAISR algorithm that turns a LR image into a HR image. The process is carried
out in several steps:
(i) A very cheap (e.g. bilinear) interpolation method is used to upscale the LR image.
(ii) A hash-table, containing set of filters, is learned from a training database, where the hash-table keys are a
function of gradient properties. The filters are applied on the output of step (i) to improve its quality.
(iii) The outputs of steps (i) and (ii) are selectively blended (with different weights at each pixel) to produce the
final result.
As a closing remark, the whole learning phase is summarized in Algorithm 2 and the upscaling stage is detailed
in Algorithm 3.
V. CT-BASED DOG SHARPENER
In this section we briefly revisit the Difference of Gaussian (DoG) operator, which is widely used for edge
enhancement [23]. Then, we introduce a very efficient way to sharpen an image using this operator while eliminating
halos, noise amplification, and other similar common sharpening artifacts.17
Algorithm 2: RAISR: Hashing-based learning.
Inputs
1: s – Upscaling factor.
2: fxigL i=1 – The ground truth HR images (optionally sharpened/enhanced).
3: fzigL i=1 – The LR (optionally compressed) versions of fxigL i=1.
Output
1: Ht : (j) 7! hj;t – Hash-table that maps each key j = (θ; λ1; µ) to the corresponding d2-dimensional filter
hj;t, where 1 ≤ t ≤ s2 is the pixel/patch type.
Process
• Qj;t 0; Vj;t 0 for all hash-table keys denoted by j and pixel-type t
for i = 1 to L do
• Compute yi, an initial interpolated version of zi
• Initialize Aj;t and bj;t to be empty matrices for all possible hash-table keys denoted by j and pixel-type t
for each pixel k in yi do
• Denote by j = (θk; λk 1; µk) the hash-table key of the pixel k
• Denote by pk 2 Rd2 the patch extracted from yi centered at k
• Denote by xi(k) the ground truth HR pixel of xi located at k
• Denote by t the type of the pixel k
• Aj;t [Aj;t ; pT k ], i.e., concatinate pT k to the end of Aj;t
• bj;t [bj;t ; xi(k)], i.e., concatinate xi(k) to the end of bj;t
end
for each key j and pixel-type t do
• Qj;t Qj;t + AT j;tAj;t
• Vj;t Vj;t + AT j;tbj;t
end
end
for each key j and pixel-type t do
• hj;t arg minh kQj;th − Vj;tk2 2
• Ht (j) hj;t
end18
Algorithm 3: RAISR: Hashing-based upscaling.
Inputs
1: s – Upscaling factor.
2: z – LR image.
3: Ht : (j) 7! hj;t – Hash-table that maps each key j = (θ; λ1; µ) to the corresponding d2-dimensional filter
hj;t, where 1 ≤ t ≤ s2 is the pixel/patch type.
Output
1: x^ – HR estimate of z.
Process
• Compute y, an initial interpolated version of z
for each pixel k in y do
• Denote by j = (θk; λk 1; µk) the hash-table key of the pixel k
• Denote by pk 2 Rd2 the patch extracted from y centered at k
• Denote by t the type of the pixel k
• hj;t Ht (j)
• x^(k) pT k ht
end
• x^ CT-Blending(x^; y)
The DoG filter is formulated as a subtraction of two low-pass filters – Gaussians with different standard-deviation
σ. In general, a Gaussian filter attenuates the high frequencies of the image, where the parameter σ controls the
blurring effect, i.e., modifies the cut-off point of the filter. As such, when subtracting two Gaussians with different
standard-deviations, we manually design a bandpass filter that reduces the amplitude of all frequencies between
two cut-off points. More formally, a DoG filter Dσ;α;ρ(z) can be expressed by
D
σ;α;ρ(z) = (1 + ρ)Gσ(z) − ρGασ(z); (10)
where Gσ(z) is a Gaussian filter with standard deviation σ, the scalar ρ controls the amplification factor, and α is
a constant that controls the range of the frequencies that we wish to pass. According to the chosen parameters, the
filter D
σ;α;ρ(z) captures different frequency-bands, thus being an efficient mechanism to amplify a wide range of
frequencies by choice [23].
Similarly to the explanation that is given in Section IV-D, when dealing with sharpening two main issues are
raised: (i) Noise amplification, and (ii) Halos artifacts. The noise may exist in the whole frequency domain, thus
amplified by the naive DoG sharpener along with the underlying signal. In addition, since this sharpener is not
adaptive to the content/structure of the image it tends to produce sharpening artifacts, such as halos, over-sharpening,19
Fig. 12. CT-based DoG sharpener: Applying the DoG filters on the image, followed by blending steps that utilize the outcome of the CT.
gradient-reversals and more. Differently from the linear DoG filter, the content-aware non-linear filtering methods
[24]–[28] successfully avoid these common artifacts. However, computationally, they are more complex than the
linear approach.
We wish to keep the computational advantages of working with the linear DoG filter, while gaining adaptivity
to the content of the image. Similarly to RAISR, we use the ”blending trick” (see Section IV-D) once again. By
integrating the CT blender to the naive DoG sharpener scheme we achieve the desired content-aware property, in
an extremely efficient fashion. The blending is similar to the one that described in the context of RAISR, with the
exception that now there are several images that are fused together. Specifically, the input image is locally blended
with its different enhanced versions, obtained by applying naive DoG filters that amplifies different frequency-bands.
Moreover, in order to reduce computations, we suggest a cascade implementation that utilizes the already filtered
images between the levels (see Fig. 12). For example, instead of applying a wide separable Gaussian filter in order to
capture the mid-low frequency band, one can filter the already computed smoothed image, obtained in the previous
cascade level. As can be easily inferred, the complexity of the proposed sharpener is very low; it is equivalent to
the application of linear separable filters on the image, followed by a pixel-wise weighted average. Notice that the
weights are a function of very cheap descriptor that applies extremely basic manipulations on 3 × 3 neighborhood
of a pixel (e.g. boolean comparisons and evaluation of Hamming distance).
Following Fig. 12, notice that different blending mechanisms leads to different enhancement effect: When counting
the modified bits of the CT descriptor, we are capable to enhance the contrast of the image (even in relatively lowfrequencies). On the other hand, when choosing the randomness measure as a blending map, we enhance the content
only along edges and structures.
VI. EXPERIMENTS
In this section we test the proposed algorithm in various scenarios. First, in the context of the conventional
SISR, the effectiveness of RAISR is tested and compared to several state-of-the-art algorithms for 2×, 3×, and
4× upscaling. Then, the abilities of the proposed algorithm to tackle real-world scenarios are demonstrated by20
applying RAISR upscaling on arbitrary compressed images. In this case, we show that RAISR is able to produce
contrast-enhanced high quality images by learning filters from compressed LR images to their contrast-enhanced
HR versions. The pre-processing of the training images is done by applying the proposed CT-based DoG sharpener.
A. Single Image Super-Resolution
In this subsection we compare the performance of proposed algorithm with several state-of-the-art methods for
2×, 3×, and 4× upscaling factors, both quantitatively and visually. All the results are obtained by applying the
scheme that is given in Fig. 10 (and its variants for 3× and 4× upscaling), followed by a back-projection step (see
Appendix A). The filters are learned using a collection of 10,000 advertising banner images. This imagery type
was chosen for its wide variety of both synthetic and real world content.
For all upscaling factors, we learn filters of size 11 × 11. In the context of the hash-table, we consider a
neighborhood of size 9 × 9 for the computation of the angle, strength, and coherence of the gradients. Also, the
quantization factors of the angle Qθ, strength Qs, and coherence Qµ are set to result in a total of 216 total buckets
(24 angular bins by 3 strength bins by 3 coherence bins). In addition, we found that an amplification of the highfrequencies of the training-set HR images leads to an improved restoration. The proposed CT-based DoG sharpener
(see Fig. 12) is utilized for this task, where we use the randomness measure as the blending map; the number of
layers is set to 3, with σ = 0:85, α = p2 and ρ = 55. Note that a fast approximation of separable Gaussian filter
is suggested, obtained by applying one dimensional [1; 2; 1]=4 filter in a separable fashion (acts as a LP Gaussian
filter with σ ≈ 0:85).
As done by the leading algorithms [11], [16], [17], the proposed algorithm is tested on 2 widely used databases:
(i) Set5 [36], and (ii) Set14 [9], which are composed of 5 and 14 standard images, respectively. For all scaling
factors, the test LR images are generated by downscaling the original HR images using the bicubic interpolation.
The restoration performance is evaluated using the Peak Signal to Noise Ratio (PSNR), and Structural Similarity
(SSIM) metrics, measured between the luminance channel of the original and the estimated images. The higher
these measures the better the restoration. Note that upscaling an RGB image is done by converting it to the YCbCr
color space. Then, a bicubic interpolation is applied on the luminance channel, which is used as an initialization
for RAISR, while the chromatic channels are upscaled only by the bicubic interpolation. Finally, the estimated HR
channels are converted back to the RGB color space.
The proposed algorithm is compared to various methods, including the sparse-coding approach of Zeyde et al.
[9], along with the efficients GR and ANR [10]. Since we put emphasis on runtime, we test two versions of A+
[11], the first is a fast version that uses 16 atoms, while the second produces state-of-the-art results (with the cost of
increased runtime) by utilizing 1024 atoms. In addition, we compare our algorithm to the state-of-the-art SRCNN
[16], [17], which is based on a powerful convolutional neural network architecture. A neighbor embedding technique
is also included, called NE+LLE, which assumes that the LR-HR patches lie on low-dimensional manifolds, having21
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
32.2
32.4
32.6
32.8
33
33.2
33.4
33.6
PSNR
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(a) PSNR, 2× upscaling
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
0.908
0.91
0.912
0.914
0.916
0.918
0.92
SSIM
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(b) SSIM, 2× upscaling
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
29
29.2
29.4
29.6
29.8
30
30.2
30.4
PSNR
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(c) PSNR, 3× upscaling
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
0.82
0.825
0.83
0.835
0.84
0.845
SSIM
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(d) SSIM, 3× upscaling
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
27.3
27.4
27.5
27.6
27.7
27.8
27.9
28
28.1
28.2
28.3
PSNR
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(e) PSNR, 4× upscaling
10-2 10-1 100 101
<--faster-- runtime [sec] --slower-->
0.75
0.755
0.76
0.765
0.77
0.775
0.78
0.785
SSIM
Zeyde
GR
ANR
NE+LS
A+(1024)
A+(16)
SRCNN
RAISR
(f) SSIM, 4× upscaling
Fig. 13. Quantitative comparison between the restoration performance vs. runtime. The vertical coordinate of each point in the scatter plots
corresponds to the average PSNR (a,c,e) and SSIM (b,d,f) of each method, measured on the test images of Set5 and Set14. The horizontal
coordinate corresponds to the average runtime (the average size of the upscaled images is about 0:63 × 106 pixels). The size of each point
reflects the standard error of the PSNR/SSIM. Detailed quantitative results can be found in the supplementary material (website address can be
found at the bottom of the first page of this paper).22
(a) Bicubic (b) SRCNN (c) A+ (d) RAISR
(e) Bicubic (f) SRCNN (g) A+ (h) RAISR
(i) Bicubic (j) SRCNN (k) A+ (l) RAISR
Fig. 14. Visual comparison for upscaling by a factor of 2 for the images Zebra, Flowers and Baby. (a,e,i) Bicubic interpolation, (b,f,j) SRCNN,
(c,g,k) A+ with 1024 atoms, and (d,h,l) RAISR.
locally similar geometry [37]3. We should note that all the baseline methods significantly outperform the bicubic
interpolation.
A quantitative comparison between RAISR and the leading SISR methods for 2×, 3×, and 4× upscaling is given
in Fig. 13. Per each method and upscaling factor we measure the average PSNR and SSIM over the images of Set5
and Set14 (the horizontal coordinate in Fig. 13) vs. the average runtime (the vertical coordinate in Fig. 13). The
size of each point equals to 10 · σPSNR err and 103 · σSSIM err , where σerr is the standard error of the PSNR/SSIM. As can
3We would like to thank the authors of [9], [11], [16], [17] for providing the software that produces the results of NE+LLE, Zeyde et al.,
GR, ANR, A+, and SRCNN. In all cases we use the original codes, with default parameters.23
TABLE I
QUANTIFYING THE BENEFIT OF THE BUILT-IN SHARPENING EFFECT. WE MEASURE THE AVERAGE PSNR AND SSIM (HIGHER IS BETTER)
OVER SET5 AND SET14 IMAGES, ALONG WITH THE STANDARD-ERROR OF EACH QUALITY METRIC.
Dataset Scaling
Learning without sharpening Learning with sharpening Improvement
PSNR σerr
PSNR SSIM σSSIM err PSNR σPSNR err SSIM σSSIM err PSNR SSIM
Set5
2× 35.913 1.374 0.947 0.015 36.153 1.337 0.951 0.015 0.241 0.004
3× 32.061 1.392 0.895 0.019 32.211 1.363 0.901 0.019 0.150 0.006
4× 29.689 1.479 0.839 0.021 29.837 1.481 0.848 0.021 0.147 0.009
Set14
2× 31.980 0.998 0.900 0.015 32.127 1.024 0.902 0.015 0.147 0.003
3× 28.764 0.985 0.809 0.026 28.860 0.999 0.812 0.027 0.096 0.003
4× 26.912 0.908 0.732 0.032 27.002 0.921 0.738 0.033 0.090 0.006
be inferred from Fig. 13, RAISR is competitive with the state-of-the-art methods in terms of these quality measures
while being much faster. More specifically, for all upscaling factors, RAISR has similar restoration quality to the
fast version of A+ (using 16 atoms), and it outperforms NE+LS, Zeyde et al., GR, and ANR. Notice that the more
complex version of A+ (the one that uses 1024 atoms) and SRCNN performs better than RAISR, but with the cost
of increased computations.
In terms of runtime, RAISR is the fastest method by far, demonstrating that high-quality results can be achieved
without sacrificing the computational complexity. Following Fig. 13, our implementation is about one to two orders
of magnitude faster than the baseline methods (please refer to the supplementary material for detailed results). The
runtime is evaluated on a 3.4GHz 6-Core Xeon desktop computer4.
In the case of upscaling by a factor of 2, a visual comparison between RAISR, bicubic and the state-of-the-art
A+ and SRCNN is provided in Fig. 14. As can be seen, the restoration quality of RAISR for both Zebra, Flowers
and Baby images is competitive with much more complex algorithms. The ability of RAISR to restore continues
edges is shown on the Zebra image. In addition, the effective reconstruction of fine details is demonstrated via the
Flowers and Baby images.
Next, we test the effect of sharpening the HR training images. To this end, we learn a new set of filters using the
same collection of 10,000 images, however this time without applying the pre-sharpening step. Following Table I,
this step indeed leads to an improvement both in PSNR and SSIM, emphasizing its potential. Motivated by these
results, a promising future direction could be to test this concept as a way to improve various other state-of-the-art
methods, e.g. A+ and SRCNN.
Fig. 15 illustrates how the Butterfly image (taken from Set5) evolves throughout the different stages of RAISR,
demonstrating the effect of the sharpening and the need for the blending step. The LR image is first upscaled by a
factor of 2 using the bicubic interpolation (see Fig. 15b). Then, the pre-learned filters are applied: Fig. 15c shows
the result when the filters are learned without sharpening, while Fig. 15d illustrates the case where the filters have
4Notably, we implemented a version of RAISR running on a mobile GPU, which runs at speeds of over 200Mpix/s. The desktop GPU version
of the same can be expected to perform an order of magnitude faster.24
(a) Original (b) Bicubic
(c) Filtered image (learning without sharpening) (d) Filtered image (learning with sharpening)
(e) Blending of (b) and (c) (f) Blending of (b) and (d)
Fig. 15. Visual illustration of the evolution of Butterfly image throughout the different blocks of RAISR, along with a demonstration of the
built-in sharpening effect. (a) The original HR image, (b) Bicubic interpolation by a factor of 2 of the input LR image, (c) Filtered image – the
filters do not include a built-in sharpening, (d) Filtered image – the filters map the initial interpolated image to its sharpened HR version, (e)
blending result of the image in (b) with the one in (c), and (f) blending result of the image in (b) with the one in (d).25
a built-in sharpening effect. As can be seen, the image in Fig. 15d is indeed sharper than the one in Fig. 15c,
and both have noticeable halos along the edges in addition to an amplification of the noise especially in flat areas.
These artifacts are then reduced by the blending step, as depicted in Fig. 15e and 15f. In this case, the improvement
achieved by the pre-sharpening during training is 0.44dB and 0.009 in terms of PSNR and SSIM, respectively.
To conclude, we tested the upscaling performance of the proposed method on popular datasets and for various
upscaling factors. As demonstrated, RAISR is much faster than the leading algorithms, while achieving a competitive
restoration performance.
B. All in One Enhancement
The ability of the proposed algorithm to handle real-world scenarios is evaluated as well. More specifically,
we test the ability of RAISR to upscale an image while reducing compression artifacts and improving the overall
contrast of the image. The results are obtained using the scheme that is given in Fig. 11. Differently from the
conventional SISR problem (Section VI-A), in this case we do not apply the back-projection step. In addition, in
order to further reduce computations, we choose the bilinear as the initial interpolator. As a result, we manage to
reduce the runtime by about a factor of 2, i.e, this version of RAISR is more than two orders of magnitude faster
than the state-of-the-art methods A+ and SRCNN.
In the learning stage, we downscaled the training images (taken from BSDS300 database [38]) using the bicubic
interpolation, followed by a compression step using JPEG, with a quality parameter that is set to 85 out of 100. In
order to achieve contrast enhancement effect, the target HR images are pre-processed using the proposed CT-based
DoG sharpener (see Fig. 12), where the blending is done by counting the modified bits of the CT descriptor, which
allows the amplification of a wide range of frequencies. Naturally, the wider the Gaussian filters that are used in the
DoG scheme, the wider the range of frequencies that are amplified. As such, we suggest using 2 DoG layers: The
first layer applies a narrow Gaussian filter, with σ ≈ 0:85 (i.e., the separable one dimensional [1; 2; 1]=4 filter),
while the second is a much larger Gaussian filter of size 64 × 64, with σ = 8:5 (i.e., we set α = 10). In both cases
we use ρ = 55. Note that the hyper parameters of RAISR (filter-size, hashing parameters, etc.) are the same as in
Section VI-A.
Before testing RAISR in a real world scenario, where the degradation model is unknown, we show the benefits of
compressing the training LR images on a synthetic example. To this end, following Fig. 16, we degrade the Comic
image (taken from Set14) by downscaling it by a factor of 2 in each axis and then apply a JPEG compression on
the result. As can be seen in Fig. 16b, when the filters are learned without compressing the training LR images we
obtain a sharp result but with the cost of undesired amplification of compression artifacts. On the other hand, when
the learning stage includes compression, the upscaled outcome has less artifacts without the loss of sharpness, as
illustrated in Fig. 16c. This result should not surprise us as we take into account the degradation model that includes
both downscaling (by a factor of 2) and compression. Quantitatively, compared to a learning stage that does not
involve such a compression, in terms of PSNR we achieve an average improvement of 0.43dB and 0.34dB for Set5
and Set14, respectively. Similarly, we obtain higher SSIM score, where the average improvement is of 0.017 and26
(a) (b) (c)
Fig. 16. Visual comparison for upscaling by a factor of 2 the JPEG-compressed Comic image (taken from Set14). (a) Bilinear interpolation,
(b) RAISR with learning stage as in Section VI-A which does not include compression of the training LR images, (c) RAISR with learning
that involves compression of the training LR images.
0.012 for Set5 and Set14, respectively.
Fig. 17 shows the superiority of handling compression artifacts along with the visually pleasant outcome of our
contrast-enhancement learning scheme. Upscaling results of a cropped and compressed version of the ResolutionChart image is given in Fig. 17a-17c, obtained by applying SRCNN, A+, and RAISR, respectively. As can be
seen, RAISR reduces some of the compression artifacts (especially around the digits and in between the circles,
located in the bottom-right and left parts of the image, respectively). Despite the reduction of compression artifacts,
RAISR successfully keeps the desired sharpness property (notice the effective restoration of the very fine details
of the circles in the left part of the image).
The ability of RAISR to handle compressed, blurred with an unknown (possibly motion-blur) kernel, and noisy
images is shown in Fig. 17d-17i. As can be inferred, the blending step and the built-in suppression of compression
artifacts increase the robustness of the algorithm, especially when tackling noisy images. Furthermore, the contrastenhancement effect leads to better looking images (notice the contrast-enhanced letters in the Newspaper image
and the amplification of the fine details in the Painting image). Additional examples on real images are shown in
supplementary material.
To summarize, we demonstrated the ability of RAISR to achieve a high-quality restoration in various scenarios.
Our experiments indicate that the proposed method stands in a line with the best algorithms that are currently
available, while being about 2 orders of magnitude faster.27
(a) SRCNN (b) A+ (c) RAISR
(d) SRCNN (e) A+ (f) RAISR
(g) SRCNN (h) A+ (i) RAISR
Fig. 17. Visual comparison for upscaling by a factor of 2 of Resolution-Chart, Newspaper and Painting images. (a,d,g) SRCNN, (b,e,h) A+
with 1024 atoms, and (c,f,i) RAISR. In this case, RAISR filters map compressed LR images to their contrast-enhanced HR versions.28
VII. CONCLUSION
In this paper we proposed a rapid and accurate learning-based approach for single image super-resolution (called
RAISR). The suggested algorithm requires a relatively small set of training images to produce a low-complexity
mechanism for increasing the resolution of any arbitrary image not seen before. The core idea behind RAISR is
to learn a mapping from LR images to their HR versions. The mapping is done by filtering a ”cheap” upscaled
version of the LR image with a set of filters, which are designed to minimize the Euclidean distance between the
input and ground-truth images.
More specifically, RAISR suggests a highly efficient locally adaptive filtering process, where the low-complexity
is kept thanks to an appealing hashing scheme: In the learning stage, a database of images is divided into buckets
(”cheap” clusters) of patches that share similar geometry, and a filter is learned per each bucket. In the application
side, based on the geometry of the LR patch (hash-table key), the relevant pre-learned filter (hash-table entry) is
chosen and applied on the patch.
Moreover, artifact-free results are obtained by blending the initial (cheap) estimation of the HR image with the
filtered one. This step is based on the observation that in flat areas the reconstruction of the cheap upscaler is
effective (there are no fine details or edges to be recovered in these areas). We harness the efficient CT descriptor
for this task. As such, with a negligible computational cost, an accurate reconstruction is achieved. Note that two
different blenders are suggested; the first allows the amplification of high frequencies only (using the ”randomness”
mask), while the second allows the enhancement of a wide range of frequencies. Differently from the randomness
mask, the latter blending mechanism counts the number of bits of the CT descriptors that were changed due to the
filtering step.
As such, the additional complexity of RAISR over a very basic interpolation method (e.g. bilinear interpolation)
is roughly the application of 2-3 linear filters on the image: The first filter leads to the hash-table-key, pointing to
the second filter that enhances the quality of the patch (the pre-learned filter), and the last filter (which is almost
negligible) is used for the blending step. Despite the extremely low computational cost, the restoration performance
of RAISR is competitive with much more complex state-of-the-art methods. For example, in the case of A+ (which
is one the fastest methods that currently available), the complexity of choosing the filter is linear with the size of the
dictionary, where the complexity of our approach for choosing the filter is constant due to the hashing mechanism.
Furthermore, sharpening and suppression of compression artifacts are achieved by pre-processing the training
images. A sharpening/contrast-enhancement effect is obtained by amplifying the details and contrast of the HR
training images. Similarly, compression artifacts are handled by compressing the LR training images. This results
in filters that are unique in profile, and perform the dual purpose of de-aliasing and sharpening in a single filter.
The pre-learned filters tend to have a smoothing effect in the center of the kernel to perform aliasing removal, while
containing ridges near the edges of the kernel which have the effect of sharpening that edge. As such, our learning
approach is able to learn an effective set of filters that would be nearly impossible to construct by hand.
Motivated by the sharpening effect that can be achieved, we suggested a novel sharpener, which is based on29
applying (separable) DoG filters on the image, which are capable of flexibly enhancing a wide range of frequencies.
As a way to reduce artifacts (e.g. halos and noise amplification), we suggest using once again the CT blender, leading
to an extremely efficient and effective sharpener.
We should note that the quality of the hashing mechanism is crucial. Therefore, a further study of the hash
function is needed, and could possibly lead to improved results. In addition, we use a simple least squares solver
to learn the filters, while improved results may be achieved by regularizing the learning with efficient priors. In a
wider perspective, we wish to explore the ability of RAISR to cast as a boosting mechanism [39]–[46], where in
this case the learned filters can be designed to map any input images (not only the ”cheap” interpolated images) to
the desired outputs.
APPENDIX
A. Back Projection
As a way to reduce artifacts and estimation errors, many methods (e.g., [14], [47]–[52]) harness the popular
Iterative Back Projection (IBP) [53] as a global post processing step. Following the degradation model in Eq. (1),
one can demand an equivalence between the LR image and the downscaled version of the estimated HR image.
More formally, given an estimation of the HR image x~ (i.e., the output of RAISR), we suggest minimizing the
following cost
min
x
kx − x~k2 2 s.t. z = DsHx; (11)
leading to the desired HR estimation x^. The solution of Eq. (11) can be obtained by applying several iterations
of gradient descent. We found it helpful to apply this post-processing step in the case of the conventional SISR
process (the results in Section VI-A).
REFERENCES
[1] R. Keys, “Cubic convolution interpolation for digital image processing,” IEEE Trans. on Acoustics, Speech and Signal Proc., vol. 29,
no. 6, pp. 1153–1160, 1981. 2
[2] H. Hou and H. Andrews, “Cubic splines for image interpolation and digital filtering,” IEEE Trans. on Acoustics, Speech and Signal Proc.,
vol. 26, no. 6, pp. 508–517, 1978. 2
[3] A. Buades, B. Coll, and J. M. Morel, “A review of image denoising algorithms, with a new one,” Multiscale Modeling and Simulation
(SIAM interdisciplinary journal), vol. 4, no. 2, pp. 490–530, 2005. 2, 3
[4] H. Takeda, S. Farsiu, and P. Milanfar, “Kernel regression for image processing and reconstruction,” IEEE Trans. on Image Proc., vol. 16,
no. 2, pp. 349–366, February 2007. 2
[5] K. Dabov, A. Foi, V. Katkovnik, and K. Egiazarian, “Image denoising by sparse 3-D transform-domain collaborative filtering,” IEEE Trans.
on Image Proc., vol. 16, no. 8, pp. 2080–2095, August 2007. 2
[6] H. Takeda, P. Milanfar, M. Protter, and M. Elad, “Super-resolution without explicit subpixel motion estimation,” IEEE Transactions on
Image Processing, vol. 18, no. 9, pp. 1958–1975, September 2009. 2
[7] M. Aharon, M. Elad, and A. Bruckstein, “The K-SVD: An algorithm for designing of overcomplete dictionaries for sparse representation,”
IEEE Transactions on Signal Processing, vol. 54, no. 11, pp. 4311–4322, November 2006. 2
[8] J. Yang, J. Wright, T. Huang, and Y. Ma, “Image super-resolution as sparse representation of raw image patches,” in IEEE Conference on
Computer Vision and Pattern Recognition, June 2008, pp. 1–8. 2, 730
[9] R. Zeyde, M. Elad, and M. Protter, “On single image scale-up using sparse-representations,” in Curves and Surfaces. Springer, 2012, pp.
711–730. 2, 7, 20, 22
[10] R. Timofte, V. De, and L. Van Gool, “Anchored neighborhood regression for fast example-based super-resolution,” in IEEE International
Conference on Computer Vision, Dec 2013, pp. 1920–1927. 2, 7, 20
[11] R. Timofte, V. De Smet, and L. Van Gool, “A+: Adjusted anchored neighborhood regression for fast super-resolution,” in ACCV, 2014,
pp. 111–126. 2, 7, 20, 22
[12] Y. Romano, M. Protter, and M. Elad, “Single image interpolation via adaptive nonlocal sparsity-based modeling,” IEEE Trans. on Image
Processing, vol. 23, no. 7, pp. 3085–3098, 2014. 2
[13] D. Zoran and Y. Weiss, “From learning models of natural image patches to whole image restoration,” in ICCV. IEEE, 2011, pp. 479–486.
2
[14] J. Yang, Z. Wang, Z. Lin, S. Cohen, and T. Huang, “Coupled dictionary training for image super-resolution,” IEEE Transactions on Image
Processing, vol. 21, no. 8, pp. 3467–3478, 2012. 2, 7, 29
[15] T. Peleg and M. Elad, “A statistical prediction model based on sparse representations for single image super-resolution,” IEEE Transactions
on Image Processing, vol. 23, no. 6, pp. 2569–2582, 2014. 2, 5, 7
[16] C. Dong, C. C. Loy, K. He, and X. Tang, “Learning a deep convolutional network for image super-resolution,” in ECCV, 2014, pp.
184–199. 2, 7, 20, 22
[17] D. Chao, C. C. Loy, H. Kaiming, and T. Xiaoou, “Image super-resolution using deep convolutional networks,” IEEE Transactions on
Pattern Analysis and Machine Intelligence, vol. 38, no. 2, pp. 295–307, Feb 2016. 2, 7, 20, 22
[18] D. Dai, R. Timofte, and L. Van Gool, “Jointly optimized regressors for image super-resolution,” in Computer Graphics Forum, vol. 34,
no. 2. Wiley Online Library, 2015, pp. 95–104. 2, 7
[19] Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, and L. D. Jackel, “Backpropagation applied to handwritten
zip code recognition,” Neural computation, vol. 1, no. 4, pp. 541–551, 1989. 2
[20] R. Zabih and J. Woodfill, “Non-parametric local transforms for computing visual correspondence,” in ECCV, 1994, pp. 151–158. 3, 14
[21] A. Polesel, G. Ramponi, V. J. Mathews et al., “Image enhancement via adaptive unsharp masking,” IEEE Transactions on Image Processing,
vol. 9, no. 3, pp. 505–510, 2000. 3
[22] D. Marr and E. Hildreth, “Theory of edge detection,” Proceedings of the Royal Society of London B: Biological Sciences, vol. 207, no.
1167, pp. 187–217, 1980. 3
[23] H. Winnemoller, J. E. Kyprianidis, and S. C. Olsen, “XDoG: An extended difference-of-Gaussians compendium including advanced image ¨
stylization,” Computers & Graphics, vol. 36, no. 6, pp. 740–753, 2012. 3, 16, 18
[24] B. Zhang and J. P. Allebach, “Adaptive bilateral filter for sharpness enhancement and noise removal,” IEEE Transactions on Image
Processing, vol. 17, no. 5, pp. 664–678, 2008. 3, 19
[25] K. He, J. Sun, and X. Tang, “Guided image filtering,” in ECCV, 2010, pp. 1–14. 3, 19
[26] H. Talebi and P. Milanfar, “Nonlocal image editing,” IEEE Transactions on Image Processing, vol. 23, no. 10, pp. 4460–4473, 2014. 3,
19
[27] A. Kheradmand and P. Milanfar, “Nonlinear structure-aware image sharpening with difference of smoothing operators,” Frontiers in ICT,
vol. 2, p. 22, 2015. 3, 19
[28] X. Liu, G. Cheung, and X. Wu, “Joint denoising and contrast enhancement of images using graph laplacian operator,” in IEEE International
Conference on Acoustics, Speech and Signal Processing. IEEE, 2015, pp. 2274–2278. 3, 19
[29] C. Tomasi and R. Manduchi, “Bilateral filtering for gray and color images,” ICCV, Bombay, India, pp. 836–846, January 1998. 3
[30] F. Durand and J. Dorsey, “Fast bilateral filtering for the display of high-dynamic-range images,” in ACM transactions on graphics (TOG),
vol. 21, no. 3. ACM, 2002, pp. 257–266. 3
[31] R. Barrett, M. W. Berry, T. F. Chan, J. Demmel, J. Donato, J. Dongarra, V. Eijkhout, R. Pozo, C. Romine, and H. Van der Vorst, Templates
for the solution of linear systems: building blocks for iterative methods. SIAM, 1994. 5
[32] S.-C. Jeong and B. C. Song, “Training-based super-resolution algorithm using k-means clustering and detail enhancement,” in European
Signal Processing Conference. IEEE, 2010, pp. 1791–1795. 7
[33] G. Yu, G. Sapiro, and S. Mallat, “Solving inverse problems with piecewise linear estimators: From gaussian mixture models to structured
sparsity,” IEEE Transactions on Image Processing, vol. 21, no. 5, pp. 2481–2499, 2012. 731
[34] V. Papyan and M. Elad, “Multi-scale patch-based image restoration,” IEEE Transactions on Image Processing, vol. 25, no. 1, pp. 249–261,
2016. 7
[35] X. Feng and P. Milanfar, “Multiscale principal components analysis for image local orientation estimation,” Proceedings of the 36th
Asilomar Conference on Signals, Systems and Computers, Pacific Grove, CA, November 2002. 9, 11
[36] M. Bevilacqua, A. Roumy, C. Guillemot, and M. L. Alberi-Morel, “Low-complexity single-image super-resolution based on nonnegative
neighbor embedding,” 2012. 20
[37] H. Chang, D.-Y. Yeung, and Y. Xiong, “Super-resolution through neighbor embedding,” in CVPR. IEEE, 2004. 22
[38] D. Martin, C. Fowlkes, D. Tal, and J. Malik, “A database of human segmented natural images and its application to evaluating segmentation
algorithms and measuring ecological statistics,” in ICCV, vol. 2. IEEE, 2001, pp. 416–423. 25
[39] H. Talebi, X. Zhu, and P. Milanfar, “How to SAIF-ly boost denoising performance,” IEEE Transactions on Image Processing, vol. 22,
no. 4, pp. 1470–1485, 2013. 29
[40] P. Milanfar, “A tour of modern image filtering: New insights and methods, both practical and theoretical,” IEEE Signal Processing Magazine,
vol. 30, no. 1, pp. 106–128, 2013. 29
[41] Y. Romano and M. Elad, “Improving K-SVD denoising by post-processing its method-noise,” in IEEE International Conference on Image
Processing. IEEE, 2013, pp. 435–439. 29
[42] H. Talebi and P. Milanfar, “Global image denoising,” IEEE Transactions on Image Processing, vol. 23, no. 2, pp. 755–768, 2014. 29
[43] A. Kheradmand and P. Milanfar, “A general framework for regularized, similarity-based image restoration,” IEEE Transactions on Image
Processing, vol. 23, no. 12, pp. 5136–5151, 2014. 29
[44] Y. Romano and M. Elad, “Patch-disagreement as a way to improve K-SVD denoising,” in IEEE International Conference on Acoustics,
Speech and Signal Processing. IEEE, 2015, pp. 1280–1284. 29
[45] ——, “Boosting of image denoising algorithms,” SIAM Journal on Imaging Sciences, vol. 8, no. 2, pp. 1187–1219, 2015. 29
[46] G. Ghimpeteanu, T. Batard, M. Bertalmio, and S. Levine, “A decomposition framework for image denoising algorithms,” IEEE Transactions
on Image Processing, vol. 25, no. 1, pp. 388–399, 2016. 29
[47] S. C. Park, M. K. Park, and M. G. Kang, “Super-resolution image reconstruction: a technical overview,” IEEE Signal Processing Magazine,
vol. 20, no. 3, pp. 21–36, 2003. 29
[48] W. Dong, L. Zhang, G. Shi, and X. Wu, “Nonlocal back-projection for adaptive image enlargement,” in IEEE International Conference
on Image Processing. IEEE, 2009, pp. 349–352. 29
[49] H. Song, X. He, W. Chen, and Y. Sun, “An improved iterative back-projection algorithm for video super-resolution reconstruction,” in
Proceedings of the Symposium on Photonics and Optoelectronic. IEEE, 2010, pp. 1–4. 29
[50] J. Yang, J. Wright, T. S. Huang, and Y. Ma, “Image super-resolution via sparse representation,” IEEE Transactions on Image Processing,
vol. 19, no. 11, pp. 2861–2873, 2010. 29
[51] L. He, H. Qi, and R. Zaretzki, “Beta process joint dictionary learning for coupled feature spaces with application to single image superresolution,” in IEEE Conference on Computer Vision and Pattern Recognition, 2013, pp. 345–352. 29
[52] R. R. Makwana and N. D. Mehta, “Single image super-resolution via iterative back projection based Canny edge detection and a Gabor
filter prior,” International Journal of Soft Computing & Engineering, 2013. 29
[53] M. Irani and S. Peleg, “Improving resolution by image registration,” CVGIP: Graphical models and image processing, vol. 53, no. 3, pp.
231–239, 1991. 29
回帖提示: 反政府言论将被立即封锁ID 在按“提交”前,请自问一下:我这样表达会给举报吗,会给自己惹麻烦吗? 另外:尽量不要使用Mark、顶等没有意义的回复。不得大量使用大字体和彩色字。【本论坛不允许直接上传手机拍摄图片,浪费大家下载带宽和论坛服务器空间,请压缩后(图片小于1兆)才上传。压缩方法可以在微信里面发给自己(不要勾选“原图),然后下载,就能得到压缩后的图片】。另外,手机版只能上传图片,要上传附件需要切换到电脑版(不需要使用电脑,手机上切换到电脑版就行,页面底部)。
您需要登录后才可以回帖 登录 | 注册

本版积分规则

手机版|Archiver|amobbs.com 阿莫电子技术论坛 ( 粤ICP备2022115958号, 版权所有:东莞阿莫电子贸易商行 创办于2004年 (公安交互式论坛备案:44190002001997 ) )

GMT+8, 2024-4-19 14:44

© Since 2004 www.amobbs.com, 原www.ourdev.cn, 原www.ouravr.com

快速回复 返回顶部 返回列表