A Closed Form Solution to Natural Image Matting

如果无法正常显示,请先停止浏览器的去广告插件。
分享至:
1. A Closed Form Solution to Natural Image Matting Anat Levin Dani Lischinski Yair Weiss School of Computer Science and Engineering The Hebrew University of Jerusalem {alevin,danix,yweiss}@cs.huji.ac.il Abstract color image, at each pixel there are 3 equations and 7 un- knowns. Obviously, this is a severely under-constrained problem, and user interaction is required to extract a good matte. Most recent methods expect the user to provide a trimap as a starting point; an example is shown in Figure 1(e). The trimap is a rough (typically hand-drawn) segmentation of the image into three regions: foreground (shown in white), background (shown in black) and unknown (shown in gray). Given the trimap, these methods typically solve for F, B, and α simultaneously. This is typically done by iterative nonlinear optimization, alternating the estimation of F and B with that of α. In practice, this means that for good re- sults the unknown regions in the trimap must be as small as possible. As a consequence, trimap-based approaches typically experience difficulty handling images with a sig- nificant portion of mixed pixels or when the foreground ob- ject has many holes [15]. In such challenging cases a great deal of experience and user interaction may be necessary to construct a trimap that would yield a good matte. Another problem with the trimap interface is that the user cannot di- rectly influence the matte in the most important part of the image: the mixed pixels. In this paper we present a new closed-form solution for extracting the alpha matte from a natural image. We derive a cost function from local smoothness assumptions on fore- ground and background colors F and B, and show that in the resulting expression it is possible to analytically elimi- nate F and B, yielding a quadratic cost function in α. The alpha matte produced by our method is the global optimum of this cost function, which may be obtained by solving a sparse linear system. Since our approach computes α di- rectly and without requiring reliable estimates for F and B, a surprisingly small amount of user input (such as a sparse set of scribbles) is often sufficient for extracting a high quality matte. Furthermore, our closed-form formulation enables one to understand and predict the properties of the solution by examining the eigenvectors of a sparse matrix, closely related to matrices used in spectral image segmentation al- gorithms. In addition to providing a solid theoretical basis for our approach, such analysis can provide useful hints to the user regarding where in the image scribbles should be placed. Interactive digital matting, the process of extracting a foreground object from an image based on limited user in- put, is an important task in image and video editing. From a computer vision perspective, this task is extremely chal- lenging because it is massively ill-posed — at each pixel we must estimate the foreground and the background col- ors, as well as the foreground opacity (“alpha matte”) from a single color measurement. Current approaches either re- strict the estimation to a small part of the image, estimating foreground and background colors based on nearby pixels where they are known, or perform iterative nonlinear es- timation by alternating foreground and background color estimation with alpha estimation. In this paper we present a closed form solution to nat- ural image matting. We derive a cost function from lo- cal smoothness assumptions on foreground and background colors, and show that in the resulting expression it is possi- ble to analytically eliminate the foreground and background colors to obtain a quadratic cost function in alpha. This allows us to find the globally optimal alpha matte by solv- ing a sparse linear system of equations. Furthermore, the closed form formula allows us to predict the properties of the solution by analyzing the eigenvectors of a sparse ma- trix, closely related to matrices used in spectral image seg- mentation algorithms. We show that high quality mattes can be obtained on natural images from a surprisingly small amount of user input. 1. Introduction Natural image matting and compositing is of central im- portance in image and video editing. Formally, image mat- ting methods take as input an image I, which is assumed to be a composite of a foreground image F and a background image B. The color of the i-th pixel is assumed to be a lin- ear combination of the corresponding foreground and back- ground colors, I i = α i F i + (1 − α i )B i , (1) where α i is the pixel’s foreground opacity. In natural image matting, all quantities on the right hand side of the com- positing equation (1) are unknown. Thus, for a 3 channel 1
2. (a) (b) (c) (d) (e) (f) Figure 1. (a) An image with sparse constraints: white scribbles indicate foreground, black scribbles indicate background. Applying Bayesian matting to such sparse input produces a completely erroneous matte (b). Foreground extraction algorithms, such as [9, 11] produce a hard segmentation (c). An automatically generated trimap from a hard segmentation may miss fine features (d). An accurate hand-drawn trimap (e) is required in this case to produce a reasonable matte (f). (Images taken from [15]) 1.1. Previous work Most existing methods for natural image matting re- quire the input image to be accompanied by a trimap [1, 2, 4, 5, 12, 14], labeling each pixel as foreground, back- ground, or unknown. The goal of the method is to solve the compositing equation (1) for the unknown pixels. This is typically done by exploiting some local regularity assump- tions on F and B to predict their values for each pixel in the unknown region. In the Corel KnockOut algorithm [2], F and B are assumed to be smooth and the prediction is based on a weighted average of known foreground and background pixels (closer pixels receive higher weight). Some algorithms [5, 12] assume that the local foreground and background come from a relatively simple color distri- bution. Perhaps the most successful of these algorithms is the Bayesian matting algorithm [5], where a mixture of ori- ented Gaussians is used to learn the local distribution and then α, F, and B are estimated as the most probable ones given that distribution. Such methods work well when the color distributions of the foreground and the background do not overlap, and the unknown region in the trimap is small. As demonstrated in Figure 1(b) a sparse set of constraints could lead to a completely erroneous matte. In contrast, while our approach also makes certain smoothness assump- tions regarding F and B, it does not involve estimating the values of these functions until after the matte has been ex- tracted. The Poisson matting method [14], also expects a trimap as part of its input, and computes the alpha matte in the mixed region by solving a Poisson equation with the matte gradient field and Dirichlet boundary conditions. In the global Poisson matting method the matte gradient field is approximated as ∇I/(F − B) by taking the gradient of the compositing equation, and neglecting the gradients in F and B. The matte is then found by solving for a function whose gradients are as close as possible to the approximated matte gradient field. Whenever F or B are not sufficiently smooth inside the unknown region, the resulting matte might not be correct, and additional local manipulations may need to be applied interactively to the matte gradient field in order to obtain a satisfactory solution. This interactive refinement process is referred to as local Poisson matting. As we shall see, our method makes weaker assumptions on the behavior of F and B, which generally leads to more accurate mattes. Recently, several successful approaches for extracting a foreground object from its background have been proposed [3, 9, 11]. Both approaches translate simple user-specified constraints (such as scribbles, or a bounding rectangle) into a min-cut problem. Solving the min-cut problem yields a hard binary segmentation, rather than a fractional alpha matte (Figure 1(c)). The hard segmentation could be trans- formed into a trimap by erosion, but this could still miss some fine or fuzzy features (Figure 1(d)). Although Rother et al. [11] do perform border matting by fitting a paramet- ric alpha profile in a narrow strip around the hard boundary, this is more akin to feathering than to full alpha matting, since wide fuzzy regions cannot be handled in this manner. Our approach is closely related to the colorization method of Levin et al. [7], and the random walk alpha mat- ting method of Grady et al. [6]. Both of these methods propagate scribbled constraints to the entire image by min- imizing a quadratic cost function. Here we apply a similar strategy, but our assumptions and cost function are modified so as to better suit the matting problem. Another scribble-based interface for interactive matting was recently proposed by Wang and Cohen [15]. Starting from a few scribbles indicating a small number of back- ground and foreground pixels, they use belief propagation to iteratively estimate the unknowns at every pixel in the image. While this approach has produced some impressive results, it has the disadvantage of employing an expensive iterative non-linear optimization process, which might con- verge to different local minima. 2. Derivation For clarity of exposition we begin by deriving a closed- form solution for alpha matting of grayscale images. This solution will then be extended to the case of color images in Section 2.1. As mentioned earlier, the matting problem is severely under-constrained. Therefore, some assumptions on the na- ture of F, B and/or α are needed. To derive our solution for the grayscale case we make the assumption that both F and B are approximately constant over a small window around
3. each pixel. Note that assuming F and B are locally smooth does not mean that the input image I is locally smooth, since discontinuities in α can account for the discontinuities in I. This assumption, which will be somewhat relaxed in Sec- tion 2.1, allows us to rewrite (1) expressing α as a linear function of the image I: α i ≈ aI i + b, ∀i ∈ w, (2) 1 B where a = F−B , b = − F−B and w is a small image window. This linear relation is similar to the prior used in [17]. Our goal in this paper will be to find α, a and b minimizing the cost function ! J (α, a, b) = ∑ j∈I ∑ (α i − a j I i − b j ) 2 + εa 2 j , (3) i∈w j where w j is a small window around pixel j. The cost function above includes a regularization term on a. One reason for adding this term is numerical stability. For example, if the image is constant in the j-th window, a j and b j cannot be uniquely determined without a prior. Also, minimizing the norm of a biases the solution towards smoother α mattes (since a j = 0 means that α is constant over the j-th window). In our implementation, we typically use windows of 3 × 3 pixels. Since we place a window around each pixel, the windows w j in (3) overlap. It is this property that en- ables the propagation of information between neighboring pixels. The cost function is quadratic in α, a and b, with 3N unknowns for an image with N pixels. Fortunately, as we show below, a and b may be eliminated from (3), leaving us with a quadratic cost in only N unknowns: the alpha values of the pixels. J (α) = min J (α, a, b) . a,b J(α) = α T L α, (4) where L is an N × N matrix, whose (i, j)-th entry is: !! 1 1 + ε (I − µ k )(I j − µ k ) ∑ 2 i k|(i, j)∈w k |w k | + σ k (5) Here δ i j is the Kronecker delta, µ k and σ 2 k are the mean and variance of the intensities in the window w k around k, and |w k | is the number of pixels in this window. 1 δ i j − |w k | Proof: Rewriting (3) using matrix notation we obtain J (α, a, b) = ∑ G k k = (G Tk G k ) −1 G Tk ᾱ k (8) Substituting this solution into (6) and denoting Ḡ k = I − G k (G Tk G k ) −1 G Tk we obtain J (α) = ∑ ᾱ Tk Ḡ Tk Ḡ k ᾱ k , k and some further algebraic manipulations show that the (i, j)-th element of Ḡ Tk Ḡ k may be expressed as: 1 δ i j − |w k | ! 1 1 + ε (I − µ k )(I j − µ k ) . 2 i |w | + σ k k Summing over k yields the expression in (5). 2.1. Color Images A simple way to apply the cost function to color images is to apply the gray level cost to each channel separately. Alternatively we can replace the linear model (2), with a 4D linear model: α i ≈ ∑ a c I i c + b, ∀i ∈ w (9) c Theorem 1 Define J (α) as Then where for every window w k , G k is defined as a (|w k |+1)×2 matrix. For each i ∈ w k , G k contains a row √ of the form [I i , 1], and the last row of G k is of the form [ ε, 0]. For a given matte α we define ᾱ k as a (|w k | + 1) × 1 vector, whose entries are α i for every i ∈ w k , and whose last entry is 0. The elements in ᾱ k and G k are ordered correspondingly. For a given matte α the optimal pair a ∗ k , b ∗ k inside each window w k is the solution to the least squares problem:   a k ∗ ∗ (a k , b k ) = argmin G k − ᾱ k (7) b k  a k b k  − ᾱ k 2 , (6) The advantage of this combined linear model is that it re- laxes our previous assumption that F and B are constant over each window. Instead, as we show below, it is enough to assume that in a small window each of F and B is a lin- ear mixture of two colors; in other words, the values F i in a small window lie on a single line in the RGB color space: F i = β i F 1 + (1 − β i )F 2 , and the same is true for the back- ground values B i . In what follows we refer to this assump- tion as the color line model. Such a model is useful since it captures, for example, the varying shading on a surface with a constant albedo. Another example is a situation where the window contains an edge between two uniformly colored regions both be- longing to the background or the foreground. Furthermore, Omer and Werman [10] demonstrated that in many natural images the pixel colors in RGB space tend to form a rela- tively small number of elongated clusters. Although these clusters are not straight lines, their skeletons are roughly linear locally.
4. Theorem 2 If the foreground and background colors in a window satisfy the color line model we can express α i = ∑ a c I i c + b, ∀i ∈ w. c Proof: Substituting into (1) the linear combinations F i = β Fi F 1 + (1 − β Fi )F 2 and B i = β Bi B 1 + (1 − β Bi )B 2 , where F 1 , F 2 , B 1 , B 2 are constant over a small window, we obtain: I i c = α i (β Fi F 1 c +(1−β Fi )F 2 c )+(1−α i )(β Bi B c 1 +(1−β Bi )B c 2 ). Let H be a 3 × 3 matrix whose c-th row is [F 2 c + B c 2 , F 1 c − F 2 c , B c 1 − B c 2 ]. Then the above may be rewritten as:   α i  = I i − B 2 , α i β Fi H  (1 − α i )β Bi where I i and B 2 are 3 × 1 vectors representing 3 color chan- nels. We denote by a 1 , a 2 , a 3 the elements in the first row of H −1 , and by b the scalar product of first row of H −1 with the vector B 2 . We then obtain α i = ∑ c a c I i + b. Using the 4D linear model (9) we define the following cost function for matting of RGB images: !   J (α, a, b) = ∑ j∈I ∑ i∈w j α i − ∑ a cj I i c − b j c 2 + ε ∑ a c j 2 c (10) Similarly to the grayscale case, a c and b can be eliminated from the cost function, yielding a quadratic cost in the α unknowns alone: J (α) = α T L α. (11) Here L is an N × N matrix, whose (i, j)-th element is:    ε 1 −1 ) (I − µ ) I 1 + (I − µ )(Σ + δ − 3 j i ij k k k ∑ |w k | |w k | k|(i, j)∈w k (12) where Σ k is a 3 × 3 covariance matrix, µ k is a 3 × 1 mean vector of the colors in a window w k , and I 3 is the 3 × 3 identity matrix. We refer to the matrix L in equations (5) and (12) as the matting Laplacian. Note that the elements in each row of L sum to zero, and therefore the nullspace of L includes the constant vector. If ε = 0 is used, the nullspace of L also includes every color channel of I. (b) (c) In our system the user-supplied constraints on the matte are provided via a scribble-based GUI. The user uses a background brush (black scribbles in our examples) to in- dicate background pixels (α = 0) and a foreground brush (white scribbles) to indicate foreground pixels (α = 1). To extract an alpha matte matching the user’s scribbles we solve for (13) (d) where S is the group of scribbled pixels and s i is the value indicated by the scribble. Theorem 3 Let I be an image formed from F and B accord- ing to (1), and let α ∗ denote the true alpha matte. If F and B satisfy the color line model in every local window w k , and if the user-specified constraints S are consistent with α ∗ , then α ∗ is an optimal solution for the system (13), where L is constructed with ε = 0. Proof: Since ε = 0, if the color line model is satisfied in every window w k , it follows from the definition (10) that J (α ∗ , a, b) = 0, and therefore J (α ∗ ) = α ∗ T L α ∗ = 0. We demonstrate this in figure 2. The first image (fig- ure 2(a)) is a synthetic example that was created by com- positing computer-simulated (monochromatic) smoke over a simple background with several color bands, which sat- isfies the color line model. The black and white scribbles show the input constraints. The matte extracted by our method (figure 2(b)) is indeed identical to the ground truth matte. The second example (figure 2(c)) is a real image, with fairly uniform foreground and background colors. By scribbling only two black and white points, a high quality matte was extracted (figure 2(d)). 4. Spectral Analysis The matting Laplacian matrix L is a symmetric positive definite matrix, as evident from theorem 1 and its proof. This matrix may also be written as L = D −W , where D is a diagonal matrix D(i, i) = ∑ j W (i, j) and W is a symmet- ric matrix, whose off-diagonal entries are defined by (12). Thus, the matrix L is the graph Laplacian used in spectral methods for segmentation, but with a novel affinity function given by (12). For comparison, the typical way to define the affinity function (e.g., in normalized cuts image segmenta- tion algorithms [13]) is to set W G (i, j) = e −kI i −I j k 3. Constraints and User Interface α = argmin α T L α, s.t. α i = s i , ∀i ∈ S (a) Figure 2. Matting examples. (a,c) Input images with scribbles. (b,d) Extracted mattes. 2 /σ 2 , (14) where σ is a global constant (typically chosen by hand). This affinity is large for nearby pixels with similar col- ors and approaches zero when the color difference is much greater than σ. The random walk matting algorithm [6] has used a similar affinity function for the matting problem, but the color distance between two pixels was taken after apply- ing a linear transformation to their colors. The transforma- tion is image-dependent and is estimated using a manifold learning technique.
5. (a) (b) (c) (d) Figure 4. Smallest eigenvectors (a-b) are used for guiding scribble placement (c). The resulting matte is shown in (d). Input image Global σ eigenvectors Matting eigenvectors Figure 3. Smallest eigenvectors of the different Laplacians. In contrast, by rewriting the matting Laplacian as L = D −W we obtain the following affinity function, which we refer to as “the matting affinity”:  ε 1 1+(I i −µ k )(Σ k + I 3 ) −1 (I j −µ k ) |w k | |w k | k|(i, j)∈w k (15) To gain intuition regarding the matting affinity, consider an image patch containing exactly two colors (e.g., an ideal edge). In this case, it can be shown (see [8] for a proof) that the affinity between two pixels of the same color decreases with distance, while the affinity between pixels of different colors is zero. In general, we obtain a similar situation to that of standard affinity functions: nearby pixels with simi- lar colors have high affinity, while nearby pixels with differ- ent colors have small affinity. However, note that the mat- ting affinity does not have a global scaling parameter σ and rather uses local estimates of means and variances. As we show subsequently, this adaptive property leads to a signif- icant improvement in performance. A similar observation was also made in [16], which suggests that local adaptation of the scaling parameter improves image segmentation re- sults. To compare the two affinity functions we examine the smallest eigenvectors of the corresponding Laplacians, since these eigenvectors are used by spectral segmentation algorithms for partitioning images. Figure 3 shows the second smallest eigenvector (the first smallest eigenvector is the constant image in both cases) for both Laplacian matrices on two example images. The first example is a simple image with concentric circles of dif- ferent color. In this case the boundaries between regions are very simple, and both Laplacians capture the transitions correctly. The second example is an image of a peacock. The global σ eigenvector (used by the normalized-cuts al- gorithm) fails to capture the complex fuzzy boundaries between the peacock’s tail feathers and the background. In contrast, the matting Laplacian’s eigenvector separates the peacock from the background very well. The matting Laplacian in this case was computed with ε = 0.0001. W M (i, j) = ∑ 4.1. The eigenvectors as guides While the matting problem is ill-posed without some user input, the matting Laplacian matrix contains a lot of information on the image even before any scribbles have been provided, as demonstrated in the previous section. This suggests that looking at the smallest eigenvectors of the matting Laplacian can guide the user where to place scribbles. For example, the extracted matte tends to be piecewise constant in the same regions where the small- est eigenvectors are piecewise constant. If the values inside a segment in the eigenvector image are coherent, a single scribble within such a segment should suffice to propagate the desired value to the entire segment. On the other hand, areas where the eigenvector’s values are less coherent cor- respond to more “difficult” regions in the image, suggesting that more scribbling efforts might be required there. Stated somewhat more precisely, the alpha matte may be predicted by examining the smaller eigenvectors of the mat- ting Laplacian, since an optimal solution to (13) will be to a large degree spanned by the smaller eigenvectors. In fact (see [8]), it is possible to bound the weight that the optimal solution will assign to larger eigenvectors, as a function of the ratios of the corresponding eigenvalues. Figure 4 illustrates how a scribbling process may be guided by the eigenvectors. By examining the two smallest eigenvectors (fig 4(a-b)) we placed a scribble inside each re- gion exhibiting coherent eigenvector values (fig 4(c)). The resulting matte is shown in fig 4(d). Note that the scribbles in fig 4(c) were our first, and single attempt to place scrib- bles on this image. 5. Results We show here results of our closed form solution for extracting alpha mattes by minimizing the matting Lapla- cian subject to the scribbled constraints. Since the mat- ting Laplacian is quadratic in alpha, the minimum can be found exactly by solving a sparse set of linear equations. We usually define the matting Laplacian using 3 × 3 win- dows. When the foreground and background color distribu- tions are not very complex using wider windows is helpful. However, using wider windows increases computation time since the resulting system is less sparse. To overcome this, we consider the linear coefficients (eq. 9) that relate an al- pha matte to an image. The coefficients obtained using wide windows on a fine resolution image are similar to those ob- tained with smaller windows on a coarser image. Therefore we can solve for the alpha matte using 3 × 3 windows on a coarse image and compute the linear coefficients which relate it to the coarse image. We then interpolate the lin-
6. ear coefficients and apply them on a finer resolution image. The alpha matte obtained using this approach is similar to the one that would have obtained by solving the matting sys- tem directly on the fine image with wider windows. More details are provided in [8]. For the results shown here we solve the matting sys- tem using Matlab’s direct solver (the “backslash” operator) which takes 20 seconds for a 200 by 300 image on a 2.8GHz CPU. Processing big images using the Matlab solver is im- possible due to memory limitations. To overcome this we use a coarse-to-fine scheme. We downsample the image and the scribbles and solve in a lower resolution. The recon- structed alpha is then interpolated to the finer resolution, alpha values are thresholded and pixels with alpha close to 0 or 1 are considered constraints in the finer resolution. Constraint pixels can be eliminated from the system, reduc- ing the system size. We have also implemented a multigrid solver for matte extraction. The multigrid solver runs in a couple of seconds even for very large images, but with a small degradation in matte quality. We show here only the extracted alpha mattes. Note that for compositing on a novel background, we also need to solve for F. After the matte has been found, it is possible to solve for the a c and b coefficients directly from equa- tion (10) and extract the foreground and background from them. However, we have found that better estimations of foreground and background are obtained by solving a new set of linear equations in F and B, derived by introducing some explicit smoothness priors on F and B into equation (1). More information on the foreground reconstruction as well as some compositing results can be found in [8]. Figure 5 shows the mattes extracted using our technique on two challenging images used in [15] and compares our results to several other recent algorithms. It can be seen that our results on these examples are comparable in quality to those of [15], even though we use a far simpler algorithm. Global Poisson matting cannot extract a good matte from sparse “scribbles” although its performance with a trimap is quite good. The random walk matting algorithm [6] also minimizes a Laplacian but uses an affinity function with a global scaling parameter and hence has a particular diffi- culty with the peacock image. To obtain a more quantitative comparison between the al- gorithms, we performed an experiment with synthetic com- posites for which we have the ground truth alpha matte. We randomly extracted 2000 subimages from the image shown in figure 6(h). We used each subimage as a background and composited over it a uniform foreground image using two different alpha mattes: the first matte is a the computer simulated smoke most of which is partially transparent; the other matte is a part of a circle, mostly opaque with a feath- ered boundary. The mattes are shown in figure 6(c). Conse- quently, we obtained 4000 composite images, two of which are shown in figure 6(a).) On this set of images we com- pared the performance of four matting algorithms — Wang and Cohen, global Poisson matting, random walk matting, and our own (using 3 × 3 windows with no pyramid). All algorithms were provided a trimap as input. Examples of the trimaps and the results produced by the different meth- ods are shown in figure 6(a,d–g)). For each algorithm, we measured the summed absolute error between the extracted matte and the ground truth. Figures 6(i,j) plot the average error of the four algorithms as a function of the smooth- ness of the background (specifically we measured the aver- age gradient strength, binned into 10 bins). The errors in the smoke matte are plotted in figure 6(i), while errors in the circle matte are plotted in figure 6(j). When the back- ground is smooth, all algorithms perform well with both mattes. When the background contains strong gradients, global Poisson matting performs poorly (recall that it as- sumes that background and foreground gradients are negli- gible). Of the remaining algorithms, our algorithm consis- tently produced the most accurate results. Figure 7 shows an example (from [15]), where Wang and Cohen’s method fails to extract a good matte from scrib- bles due to color ambiguity between the foreground and the background. The same method, however, is able to pro- duce an acceptable matte when supplied with a trimap. Our method produces a cleaner, but also not perfect matte from the same set of scribbles, but adding a small number of ad- ditional scribbles results in a better matte. Figure 8 shows another example (a closeup of the Koala image from [14]), where there’s an ambiguity between fore- ground and background colors. In this case the matte pro- duced by our method is clearly better than the one produced by the Wang-Cohen method. To better understand why this is the case, we show an RGB histogram of representative pixels from the F and B scribbles. Some pixels in the back- ground fit the foreground color model much better then the background one (one such pixel is marked red in 8(b) and indicated by an arrow in 8(d)). As a result such pixels are classified as foreground with a high degree of certainty in the first stage. Once this error has been made it only re- inforces further erroneous decisions in the vicinity of that pixel, resulting in a white clump in the alpha matte. Since our method does not make use of global color models for F and B it can handle ambiguous situations such as that in Figure 8. However, there are also cases where our method fails to produce an accurate matte for the very same reason. Figure 9 shows an actress in front of a background with two colors. Even though the black B scribbles cover both colors the generated matte includes parts of the back- ground (between the hair and the shoulder on the left). In such cases, the user would have to add another B scribble in that area. 6. Discussion Matting and compositing are tasks of central importance in image and video editing and pose a significant challenge for computer vision. While this process by definition re-
7. (a) Peacock scribbles (b) Poisson from scribbles (c) Wang-Cohen (d) Our result (e) Peacock trimap (f) Poisson from trimap (g) Bayesian (h) Random walk (i) Fire scribbles (j) Poisson from scribbles (k) Wang-Cohen (l) Our result (m) Fire trimap (n) Poisson from trimap (o) Bayesian (p) Random walk Figure 5. A comparison of alpha mattes extracted by different algorithms. Images (a,c,e,g,i,k,m,o) are taken from [15]. The remaining images were generated by our own implementation of the respective methods. (a) Composite (b) Trimap (c) Ground truth (d) Wang-Cohen (e) Poisson 200 (f) Random walk 180 160 160 140 120 120 100 100 80 80 60 60 40 40 20 Wang&Cohen Poisson Random Walk Ours 180 140 0 (g) Our result 200 Wang&Cohen Poisson Random Walk Ours 20 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10 (h) Background source image (i) Errors (smoke matte) (j) Errors (circle matte) Figure 6. A quantitative comparison using two ground truth mattes. The errors in (i) and (j) are plotted as a function of average gradient strength of the background, binned into 10 bins. To produce these results we used our own implementation of the respective methods, using the parameter values specified in the original papers. quires user interaction, the performance of most existing al- gorithms deteriorates rapidly as the amount of user input decreases. In this paper, we have introduced a cost function based on the assumption that foreground and background colors vary smoothly and showed how to analytically elim- inate the foreground and background colors to obtain a quadratic cost function in alpha. The resulting cost func- tion is similar to cost functions obtained in spectral meth-
8. (a) (b) (c) (d) Figure 7. An example (from [15]) with color ambiguity between foreground and background. (a) scribbles and matte by [15]; (b) [15] results using a trimap; (c) our result with scribbles similar to those in (a); (d) our results with a few additional scribbles. (a) Scribbles (b) Wang-Cohen (c) Our result 1 0.9 0.8 0.7 0.6 0.5 1 0.5 0.6 0.95 0.7 0.9 0.85 0.8 0.8 0.75 0.7 0.9 0.65 1 (d) RGB histogram of F (red) and B (blue) pixels. Figure 8. An example with ambiguity between F and B. Figure 9. Failure due to lack of a color model. ods to image segmentation but with a novel affinity function that is derived from the formulation of the matting problem. The global minimum of our cost can be found efficiently by solving a sparse set of linear equations. Our experiments on real and synthetic images show that our algorithm clearly outperforms other algorithms that use quadratic cost func- tions which are not derived from the matting equations. Our experiments also demonstrate that our results are competi- tive with those obtained by much more complicated, non- linear, cost functions. However, compared to previous non- linear approaches, we can obtain solutions in a few seconds, and we can analytically prove properties of our solution and provide guidance to the user by analyzing the eigenvectors of our operator. While our approach assumes smoothness in foreground and background colors, it does not assume a global color distribution for each segment. Our experiments have demonstrated that our local smoothness assumption often holds for natural images. Nevertheless, it would be interest- ing to extend our formulation to include additional assump- tions on the two segments (e.g., global models, local texture models, etc.). The goal is to incorporate more sophisticated models of foreground and background but still obtain high quality results using simple numerical linear algebra. References [1] N. E. Apostoloff and A. W. Fitzgibbon. Bayesian video mat- ting using learnt image priors. In Proc. CVPR, 2004. [2] A. Berman, P. Vlahos, and A. Dadourian. Comprehensive method for removing from an image the background sur- rounding a selected object. US Patent no. 6,135,345, 2000. [3] Y. Boykov and M. P. Jolly. Interactive graph cuts for optimal boundary & region segmentation of objects in n-d images. In Proc. ICCV, 2001. [4] Y. Chuang, A. Agarwala, B. Curless, D. Salesin, and R. Szeliski. Video matting of complex scenes. ACM Trans. Graph., 21(3):243–248, 2002. [5] Y. Chuang, B. Curless, D. Salesin, and R. Szeliski. A Bayesian approach to digital matting. In Proc. CVPR, 2001. [6] L. Grady, T. Schiwietz, S. Aharon, and R. Westermann. Ran- dom walks for interactive alpha-matting. In Proc. VIIP05. [7] A. Levin, D. Lischinski, and Y. Weiss. Colorization using optimization. ACM Transactions on Graphics, 2004. [8] A. Levin, D. Lischinski, and Y. Weiss. A closed form solu- tion to natural image matting. Hebrew University Technical Report, 2006. [9] Y. Li, J. Sun, C.-K. Tang, and H.-Y. Shum. Lazy snapping. ACM Trans. Graph., 23(3):303–308, 2004. [10] I. Omer and M. Werman. Color lines: Image specific color representation. In Proc. CVPR 2004, June 2004. [11] C. Rother, V. Kolmogorov, and A. Blake. ”grabcut”: inter- active foreground extraction using iterated graph cuts. ACM Trans. Graph., 23(3):309–314, 2004. [12] M. Ruzon and C. Tomasi. Alpha estimation in natural im- ages. In Proc. CVPR, 2000. [13] J. Shi and J. Malik. Normalized cuts and image segmenta- tion. In Proc. CVPR, pages 731–737, 1997. [14] J. Sun, J. Jia, C.-K. Tang, and H.-Y. Shum. Poisson matting. ACM Trans. Graph., 23(3):315–321, 2004. [15] J. Wang and M. Cohen. An iterative optimization approach for unified image segmentation and matting. In Proc. IEEE Intl. Conf. on Computer Vision, 2005. [16] L. Zelnik-Manor and P. Perona. Self-tuning spectral cluster- ing. In Advances in Neural Information Processing Systems 17. 2005. [17] A. Zomet and S. Peleg. Multi-sensor super resolution. In Proceedings of the IEEE Workshop on Applications of Com- puter Vision, 2002.

trang chủ - Wiki
Copyright © 2011-2024 iteam. Current version is 2.133.0. UTC+08:00, 2024-09-24 03:29
浙ICP备14020137号-1 $bản đồ khách truy cập$