ArticlePDF Available

Abstract and Figures

We present a general framework of a special type of least squares (LS) es-timator, which we call "HyperLS," for parametiper estimation that frequently arises in computer vision applications. It minimizes the algebraic distance un-der a special scale normalization, which is derived by a detailed error analysis in such a way that statistical bias is removed up to second order noise terms. We discuss in detail many theoretical issues involved in its derivation. By nu-merical experiments, we show that HyperLS is far superior to the standard LS and comparable in accuracy to maximum likelihood (ML), which is known to produce highly accurate results but may fail to converge if poorly initialized. We conclude that HyperLS is a perfect candidate for ML initialization.
Content may be subject to copyright.
IPSJ Transactions on Computer Vision and Applications Vol. 3 80–94 (Oct. 2011)
Regular Paper
HyperLS for Parameter Estimation in Geometric Fitting
Kenichi Kanatani,1Prasanna Rangarajan,2
Yasuyuki Sugaya3and Hirotaka Niitsuma1
We present a general framework of a special type of least squares (LS) es-
timator, which we call “HyperLS,” for parametiper estimation that frequently
arises in computer vision applications. It minimizes the algebraic distance un-
der a special scale normalization, which is derived by a detailed error analysis
in such a way that statistical bias is removed up to second order noise terms.
We discuss in detail many theoretical issues involved in its derivation. By nu-
merical experiments, we show that HyperLS is far superior to the standard LS
and comparable in accuracy to maximum likelihood (ML), which is known to
produce highly accurate results but may fail to converge if poorly initialized.
We conclude that HyperLS is a perfect candidate for ML initialization.
1. Introduction
An important task in computer vision is the extraction of 2-D/3-D geometric
information from image data7),8), for which we often need to estimate parame-
ters from observations that should satisfy implicit polynomials in the absence of
noise. For such a problem, maximum likelihood (ML) is known to produce highly
accurate solutions, achieving the theoretical accuracy limit to a first approxima-
tion in the noise level3),8),10). However, ML requires iterative search, which does
not always converge unless started from a value sufficiently close to the solu-
tion. For this reason, various numerical schemes that can produce reasonably
accurate approximations have been extensively studied7) . The simplest of such
schemes is algebraic distance minimization, or simply least squares (LS), which
minimizes the sum of the squares of polynomials that should be zero in the ab-
sence of noise. However, the accuracy of LS is very much limited. Recently, a
new approach for increasing the accuracy of LS has been proposed in several
1 Okayama University
2 Southern Methodist University
3 Toyohashi University of Technology
applications1),12),22)–24). In this paper, we call it HyperLS and present a unified
formulation and clarify various theoretical issues that have not been fully studied
so far.
Section 2 defines the mathematical framework of the problem with illustrating
examples. Section 3 introduces a statistical model of observation. In Section 4,
we discuss various issues of ML. Section 5 describes a general framework of al-
gebraic fitting. In Sections 6 and 7, we do a detailed error analysis of algebraic
fitting in general and in Section 8 derive expressions of covariance and bias of the
solution. In Section 9, we define HyperLS by choosing the scale normalization
that eliminates the bias up to second order noise terms. In Section 10, we do
numerical experiments to show that HyperLS is far superior to the standard LS
and is comparable in accuracy to ML, which implies that HyperLS is a perfect
candidate for initializing the ML iterations. In Section 11, we conclude.
2. Geometric Fitting
The term “image data” in this paper refers to values extracted from images
by image processing operations such as edge filters and interest point detectors.
An example of image data includes the locations of points that have special
characteristics in the images or the lines that separate image regions having
different properties. We say that image data are “noisy” in the sense that image
processing operations for detecting them entail uncertainty to some extent. Let
x1,... xNbe noisy image data, which we regard as perturbations in their true
values ¯
x1,...,¯
xNthat satisfy implicit geometric constraints of the form
F(k)(x;θ)=0, k = 1,...,L. (1)
The unknown parameter θallows us to infer the 2-D/3-D shape and motion of
the objects observed in the images7),8) . We call this type of problem geometric
fitting8). In many vision applications, we can reparameterize the problem to
make the functions F(k)(x;θ) linear in θ(but generally nonlinear in x), allowing
us to write Eq. (1) as
(ξ(k)(x),θ) = 0, k = 1,...,L, (2)
where and hereafter (a,b) denotes the inner product of vectors aand b. The
80 c
°2011 Information Processing Society of Japan
(x , y )
α α
(x , y )
α α
(x ’, y ’)
α α
(x , y )
α α (x, y )
α α
(a) (b) (c)
Fig. 1 (a) Fitting an ellipse to a point sequence. (b) Computing the fundamental matrix from corresponding points between two
images. (c) Computing a homography between two images.
vector ξ(k)(x) represents a nonlinear mapping of x.
Example 1 (Ellipse fitting). Given a point sequence (xα, yα), α= 1, . . . N,
we wish to fit an ellipse of the form
Ax2+ 2Bxy +Cy2+ 2(Dx +Ey) + F= 0.(3)
(Fig. 1(a)). If we let
ξ= (x2,2xy, y2,2x, 2y , 1)>,θ= (A, B, C, D, E , F)>,(4)
the constraint in Eq. (3) has the form of Eq. (2) with L= 1.
Example 2 (Fundamental matrix computation). Corresponding points
(x, y) and (x0, y0) in two images of the same 3-D scene taken from different posi-
tions satisfy the epipolar equation7)
(x,F x0)=0,x(x, y, 1)>,x0(x0, y0,10)>,(5)
where Fis called the fundamental matrix , from which we can compute the camera
positions and the 3-D structure of the scene7),8) (Fig. 1(b)). If we let
ξ= (xx0, xy0, x, yx0, y y0, y, x0, y0,1)>,
θ= (F11, F12 , F13, F21 , F22 , F23, F31 , F32 , F33)>,(6)
the constraint in Eq. (5) has the form of Eq. (2) with L= 1.
Example 3 (Homography computation). Two images of a planar or in-
finitely far away scene are related by a homography of the form
x0'Hx,x(x, y, 1)>,x0(x0, y0,10)>,(7)
where His a nonsingular matrix, and 'denotes equality up to a nonzero mul-
tiplier7),8) (Fig. 1(c)). We can alternatively express Eq. (7) as the vector product
equality
x0×Hx =0.(8)
If we let
ξ(1) = (0,0,0,x, y, 1, xy0, y y0, y0)>,
ξ(2) = (x, y, 1,0,0,0,xx0,yx0,x0)>,
ξ(3) = (xy0,yy0,y0, xx0, yx0, x0,0,0,0)>,(9)
θ= (H11, H12 , H13, H21, H22 , H23, H31 , H32 , H33)>,(10)
the three components of Eq. (8) have the form of Eq. (2) with L= 3. Note
that ξ(1),ξ(2) , and ξ(3) in Eq. (9) are linearly dependent; only two of them are
independent.
3. Statistical Model of Observation
Before proceeding to the error analysis of the above problems, we need to intro-
duce a statistical model of observation. We regard each datum xαas perturbed
from its true value ¯
xαby ∆xα, which we assume to be independent Gaussian
noise of mean 0and covariance matrix V[xα]. We do not impose any restrictions
on the true values ¯
xαexcept that they should satisfy Eq. (1). This is known
as a functional model. We could alternatively introduce some statistical model
according to which the true values ¯
xαare sampled. Then, the model is called
structural. This distinction is crucial when we consider limiting processes in the
following sense10).
Conventional statistical analysis mainly focuses on the asymptotic behavior as
the number of observations increases to . This is based on the reasoning that
81
the mechanism underlying noisy observations would better reveal itself as the
number of observations increases (the law of large numbers) while the number of
available data is limited in practice. So, the estimation accuracy vs. the number
of data is a major concern. In this light, efforts have been made to obtain a con-
sistent estimator for fitting an ellipse to noisy data or computing the fundamental
matrix from noisy point correspondences such that the solution approaches its
true value in the limit N→ ∞ of the number Nof the data17),18) .
In image processing applications, in contrast, one cannot “repeat” observa-
tions. One makes an inference given a single set of images, and how many times
one applies image processing operations, the result is always the same, because
standard image processing algorithms are deterministic; no randomness is in-
volved. This is in a stark contrast to conventional statistical problems, where
we view observations as “samples” from potentially infinitely many possibilities
and could obtain, by repeating observations, different values originating from
unknown, uncontrollable, or unmodeled causes, which we call “noise” as a whole.
In image-based applications, the accuracy of inference deteriorates as the un-
certainty of image processing operations increases. Thus, the inference accuracy
vs. the uncertainty of image operations, which we call “noise” for simplicity, is a
major concern. Usually, the noise is very small, often subpixel levels. In light of
this observation, it has been pointed out that in image domains the “consistency”
of estimators should more appropriately be defined by the behavior in the limit
σ0 of the noise level σ3),10).
In this paper, we are interested in image processing applications and focus on
the perturbation analysis around σ= 0 with the number Nof data fixed. Thus,
the functional model suits our purpose. If we want to analyze the error behavior
in the limit of N→ ∞, we need to assume some structural model that specifies
how the statistical characteristics of the data depend on N. The derivation of
consistent estimators for N→ ∞ is based on such an assumption17),18). However,
it is difficult to predict the noise characteristics for different N. Image processing
filters usually output a list of points or lines or their correspondences along with
their confidence values, from which we use only those with high confidence. If
we want to collect a lot of data, we necessarily need to include those with low
confidence, but their statistical properties are hard to estimate, since such data
are possibly misdetections. This is the most different aspect of image processing
from laboratory experiments, in which any number of data can be collected by
repeated trials.
4. Maximum Likelihood for Geometric Fitting
Under the Gaussian noise model, maximum likelihood (ML) of our problem
can be written as the minimization of the Mahalanobis distance
I=
N
X
α=1
(¯
xαxα, V [xα]1(¯
xαxα)),(11)
with respect to ¯
xαsubject to the constraint that
(ξ(k)(¯
xα),θ)=0, k = 1,...,L, (12)
for some θ. If the noise is homogeneous and isotropic, Eq. (11) is the sum of the
squares of the geometric distances between the observations xαand their true
values ¯
xα, often referred to as the reprojection error7) . That name originates
from the following intuition: We infer the 3-D structure of the scene from its
projected images, and when the inferred 3-D structure is “reprojected” onto the
images, Eq. (11) measures the discrepancy between the “reprojections” of our
solution and the actual observations.
In statisitcs, ML is criticized for its lack of consistency17),18). In fact, estimation
of the true values ¯
xα, called nuisance parameters when viewed as parameters,
is not consistent as N→ ∞ in the ML framework, as pointed out by Neyman
and Scott21) as early as in 1948. As discussed in the preceding section, however,
the lack of consistency has no realistic meaning in vision applications. On the
contrary, ML has very desirable properties in the limit σ0 of the noise level
σ: the solution is “consistent” in the sense that it converges to the true value
as σ0 and “efficient” in the sense that its covariance matrix approaches a
theoretical lower bound as σ03),8),10).
According to the experience of many vision researchers, ML is known to pro-
duce highly accurate solutions7), and no necessity is felt for further accuracy
improvement. Rather, a major concern is its computational burden, because ML
usually requires complicated nonlinear optimization.
82
The standard approach is to introduce some auxiliary parameters to express
each of ¯
xαexplicitly in terms of θand the auxiliary parameters. After they are
substituted back into Eq. (11), the Mahalanobis distance Ibecomes a function of
θand the auxiliary parameters. Then, this joint parameter space, which usually
has very high dimensions, is searched for the minimum. This approach is called
bundle adjustment 7),27) , a term originally used by photogrammetrists. This is
very time consuming, in particular if one seeks a globally optimal solution by
searching the entire parameter space exhaustively6) .
A popular alternative to bundle adjustment is minimization of a function of θ
alone, called the Sampson error7) , which approximates the minimum of Eq. (11)
for a given θ(the actual expression is shown in Section 6). Kanatani and Sug-
aya16) showed that the exact ML solution can be obtained by repeating Sampson
error minimization, each time modifying the Sampson error so that in the end
the modified Sampson error coincides with the Mahalanobis distance. It turns
out that in many practical applications the solution that minimizes the Sampson
error coincides with the exact ML solution up to several significant digits; usually,
two or three rounds of Sampson error modification are sufficient11),14),15).
However, minimizing the Sampson error is not straightforward. Many numer-
ical schemes have been proposed, including the FNS (Fundamental Numerical
Scheme) of Chojnacki et al.4), the HEIV (Heteroscedastic Errors-in-Variable) of
Leedan and Meer19) and Matei and Meer20), and the projective Gauss-Newton
iterations of Kanatani and Sugaya13). All these rely on local search, but the iter-
ations do not always converge if not started from a value sufficiently close to the
solution. Hence, accurate approximation schemes that do not require iterations
are very much desired, even though the solution may not be optimal, and various
algebraic methods have been studied in the past.
5. Algebraic Fitting
For the sake of brevity, we abbreviate ξ(k)(xα) as ξ(k)
α.Algebraic fitting refers
to minimizing the algebraic distance
J=1
N
N
X
α=1
L
X
k=1
(ξ(k)
α,θ)2=1
N
N
X
α=1
L
X
k=1
θ>ξ(k)
αξ(k)>
αθ= (θ,M θ),(13)
where we define
M=1
N
N
X
α=1
L
X
k=1
ξ(k)
αξ(k)>
α.(14)
Equation (13) is trivially minimized by θ=0unless some scale normalization
is imposed on θ. The most common normalization is kθk= 1, which we call
the standard LS. However, the solution depends on the normalization. So, we
naturally ask: What normalization will maximize the accuracy of the solution?
This question was raised first by Al-Sharadqah and Chernov1) and Rangarajan
and Kanatani23) for circle fitting, then by Kanatani and Rangarajan12) for ellipse
fitting and by Niitsuma et al.22) for homography computation. In this paper, we
generalize these results to an arbitrary number of constraints. Following these
authors1),12),22),23), we consider the class of normalizations
(θ,Nθ) = c, (15)
with some symmetric matrix Nfor a nonzero constant c. In Eq. (15), θis the
optimization parameter and Nis an unknown matrix to be determined, while the
constant cis fixed for the problem. We need not specify the value of c, because N
is unknown. Since Eq. (15) can be written as (θ,(N/c)θ) = 1, we may determine
N0=N/c instead of N, but the form of Eq. (15) with cunspecified is more
convenient in our analysis.
Traditionally, the matrix Nis positive definite or semidefinite, but in the fol-
lowing, we allow Nto be nondefinite (i.e., neither positive nor negative definite),
so the constant cin Eq. (15) is not necessarily positive. The standard treatment
of algebraic fitting goes as follows. Given the matrix N, the solution θthat
minimizes Eq. (13) subject to Eq. (15), if it exists, is given by the solution of the
generalized eigenvalue problem
Mθ =λN θ.(16)
Note that Mis always positive semidefinite from its definition in Eq. (14). If
there is no noise in the data, we have (θ,ξ(k)
α) = 0 for all kand α. Hence, Eq. (14)
implies Mθ =0, so λ= 0. In the presence of noise, Mis positive definite, so
λis positive whether Nis positive definite or semidefinite. The corresponding
solution is obtained as the eigenvector θfor the smallest λ. For the standard LS,
for which N=I, Eq. (16) becomes an ordinary eigenvalue problem
83
Mθ =λθ,(17)
and the solution is the unit eigenvector θof Mfor the smallest eigenvalue λ.
This is the traditional treatment of algebraic fitting, but the situation is slightly
different here: Nis not yet given and can be nondefinite, and the eigenvalues of
Eq. (16) may not be all positive. So, we face the problem of which eigenvalues
and eigenvectors of Eq. (16) to choose as a solution. In the following, we do
perturbation analysis10) of Eq. (16) by assuming that λ0 and choose the solu-
tion to be the eigenvector θfor the λwith the smallest absolute value, although
in theory there remains a possibility that another choice happens to produce a
better result in some cases. We also regard Eq. (16) as the definition of “alge-
braic fitting,” rather than Eq. (13) and Eq. (15). This is because, while Eq. (16)
always has a solution, Eq. (13) may not be minimized subject to Eq. (15) by a
finite θ. This can occur, for example, when the contour of (θ,Mθ), which is a
hyperellipsoid in the space of θ, happens to be elongated in a direction in the null
space of N. Then, the minimum of (θ,Mθ) could be reached in the limit of kθk
→ ∞. Note that Eq. (15) is unable to normalize the norm kθkinto a finite value
if Nhas a null space. Theoretically, such an anomaly can always occur because
Mis a random variable defined by noisy data, and if the probability of such
an occurrence is nearly 0, it may still lead to E[kˆ
θk] = 2). Once the problem
is converted to Eq. (16), for which eigenvectors θhave scale indeterminacy, we
can adopt normalization kθk= 1 rather than Eq. (13). Then, the solution θis
always a unit vector.
6. Error Analysis
We can expand each ξ(k)
αin the form
ξ(k)
α=¯
ξ(k)
α+ ∆1ξ(k)
α+ ∆2ξ(k)
α+···,(18)
where ¯
ξ(k)
αis the noiseless value, and ∆iξ(k)
αis the ith order term in ∆xα. The
first order term is written as
1ξ(k)
α=T(k)
αxα,T(k)
αξ(k)(x)
x¯¯¯¯¯x=¯
xα
.(19)
We define the covariance matrices of ξ(k)
α,k= 1, ...,L, by
V(kl)[ξα]E[∆1ξ(k)
α1ξ(l)>
α] = T(k)
αE[∆xαx>
α]T(l)>
α=T(k)
αV[xα]T(l)>
α,
(20)
where E[·] denotes expectation.
The Sampson error that we mentioned in Section 4, which approximates the
minimum of the Mahalanobis distance in Eq. (11) subject to the constraints in
Eq. (12), has the following form7),8):
K(θ) = 1
N
N
X
α=1
L
X
k,l,=1
W(kl)
α(ξ(k)
α,θ)(ξ(l)
α,θ).(21)
Here, W(kl)
αis the (kl) element of (Vα)
r, and Vαis the matrix whose (kl)
element is
Vα=³(θ, V (kl)[ξα]θ)´,(22)
where the true data values ¯
xαin the definition of V(kl)[ξα] are replaced by their
observations xα. The operation ( ·)
rdenotes the pseudoinverse of truncated
rank r(i.e., with all eigenvalues except the largest rreplaced by 0 in the spectral
decomposition), and ris the rank (the number of independent equations) of
the constraint in Eq. (12). The name “Sampson error” stems from the classical
ellipse fitting scheme of Sampson25). For given x(k)
α, Eq. (21) can be minimized
by various means including the FNS4) , HEIV19),20), and the pro jective Gauss-
Newton iteration13).
Example 4 (Ellipse fitting). For the ellipse fitting in Example 1, the first
order error ∆1ξis written as
1ξα= 2 ïxα¯yα0 1 0 0
0 ¯xα¯yα0 1 0 !>Ãxα
yα!.(23)
The second order error ∆2ξαhas the following form:
2ξα= (∆x2
α,2∆xαyα,y2
α,0,0,0)>.(24)
Example 5 (Fundamental matrix computation). For the fundamental ma-
trix computation in Example 2, the first order error ∆1ξis written as
84
1ξα=
¯x0
α¯y0
α1 0 0 0 0 0 0
0 0 0 ¯x0
α¯y0
α1 0 0 0
¯xα0 0 ¯yα0 0100
0 ¯xα0 0 ¯yα0 0 1 0
>
xα
yα
x0
α
y0
α
.(25)
The second order error ∆2ξαhas the following form:
2ξα= (∆xαx0
α,xαy0
α,0,yαx0
α,yαy0
α,0,0,0,0)>.(26)
Example 6 (Homography computation). For the fundamental matrix com-
putation in Example 2, the first order error ∆1ξis written as
1ξ(1)
α=
0001 0 0 ¯y0
α0 0
0 0 0 0 1 0 0 ¯y0
α0
0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 ¯xα¯yα1
>
xα
yα
x0
α
y0
α
,
1ξ(2)
α=
100000¯x0
α0 0
010000 0 ¯x0
α0
000000¯xα¯yα1
000000 0 0 0
>
xα
yα
x0
α
y0
α
,
1ξ(3)
α=
¯y0
α0 0 ¯x0
α0 0000
0¯y0
α0 0 ¯x0
α0000
0 0 0 ¯xα¯yα1000
¯xα¯yα1 0 0 0 0 0 0
>
xα
yα
x0
α
y0
α
.(27)
The second order error ∆2ξ(k)
αhas the following form:
2ξ(1)
α= (0,0,0,0,0,0,xαy0
α,yαy0
α,0)>,
2ξ(2)
α= (0,0,0,0,0,0,x0
αxα,x0
αyα,0)>,
2ξ(3)
α= (y0
αxα,y0
αyα,0,x0
αxα,x0
αyα,0,0,0,0)>.(28)
7. Perturbation Analysis
Substituting Eq. (18) into Eq. (14), we obtain
M=¯
M+ ∆1M+ ∆2M+···,(29)
where
¯
M=1
N
N
X
α=1
L
X
k=1
¯
ξ(k)
α¯
ξ(k)>
α,(30)
1M=1
N
N
X
α=1
L
X
k=1
(¯
ξ(k)
α1ξ(k)>
α+ ∆1ξ(k)
α¯
ξ(k)>
α),(31)
2M=1
N
N
X
α=1
L
X
k=1
(¯
ξ(k)
α2ξ(k)>
α+ ∆1ξ(k)
α1ξ(k)>
α+ ∆2ξ(k)
α¯
ξ(k)>
α).(32)
We also expand the solution θand λof Eq. (16) in the form
θ=¯
θ+ ∆1θ+ ∆2θ+···, λ =¯
λ+ ∆1λ+ ∆2λ+···.(33)
Substituting Eq. (29) and Eq. (33) into Eq. (16), we have
(¯
M+ ∆1M+ ∆2M+···)(¯
θ+ ∆1θ+ ∆2θ+···)
= (¯
λ+ ∆1λ+ ∆2λ+···)N(¯
θ+ ∆1θ+ ∆2θ+···).(34)
Note that Nis a variable to be determined, not a given function of observations,
so it is not expanded. Since we consider perturbations near the true values, the
resulting matrix Nmay be a function of the true data values. In that event, we
replace the true data values by their observations and do an a posteriori analysis
to see how this affects the accuracy. For the moment, we regard Nas an unknown
variable. From a strictly mathematical point of view, the two sides of Eq. (34)
may not define an absolutely convergent series expansion. Here, we do not go into
such a theoretical question; we simply test the usefulness of the final results by
experiments a posteriori, as commonly done in physics and engineering. At any
rate, we are concerned with only up to the second order terms in the subsequent
analysis.
Equating terms of the same order in Eq. (34), we obtain
¯
M¯
θ=¯
λN¯
θ,(35)
¯
M1θ+ ∆1M¯
θ=¯
λN1θ+ ∆1λN¯
θ,(36)
¯
M2θ+ ∆1M1θ+ ∆2M¯
θ=¯
λN2θ+ ∆1λN1θ+ ∆2λN¯
θ.(37)
We have ¯
M¯
θ=0for the true values, so ¯
λ= 0. From Eq. (31), we have
(¯
θ,1¯
M¯
θ) = 0. Computing the inner product of Eq. (36) and ¯
θon both sides,
we see that ∆1λ= 0. Multiplying Eq. (36) by the pseudoinverse ¯
Mof ¯
Mfrom
left, we obtain
85
1θ=¯
M1M¯
θ.(38)
Note that since ¯
M¯
θ=0, the matrix ¯
M¯
M(P¯
θ) is the projection operator
in the direction orthogonal to ¯
θ. Also, equating the first order terms in the
expansion k¯
θ+ ∆1θ+ ∆2θ+···k2= 1 shows (¯
θ,1θ) = 010), hence P¯
θ1θ=
1θ. Substituting Eq. (38) into Eq. (37) and computing its inner product with
¯
θon both sides, we obtain
2λ=(¯
θ,2M¯
θ)(¯
θ,1M¯
M1M¯
θ)
(¯
θ,N¯
θ)=(¯
θ,T¯
θ)
(¯
θ,N¯
θ),(39)
where we put
T= ∆2M1M¯
M1M.(40)
Next, we consider the second order error ∆2θ. Since θis normalized to unit
norm, we are interested in the error component orthogonal to ¯
θ. So, we consider
2θP¯
θ2θ(= ¯
M¯
M2θ).(41)
Multiplying Eq. (37) by ¯
Mfrom left and substituting Eq. (38), we obtain
2θ= ∆2λ¯
MN¯
θ+¯
M1M¯
M1M¯
θ¯
M2M¯
θ
=(¯
θ,T¯
θ)
(¯
θ,N¯
θ)¯
MN¯
θ¯
MT¯
θ.(42)
8. Covariance and Bias
8.1 Covariance Analysis
From Eq. (38), the covariance matrix V[θ] of the solution θhas the leading
term
V[θ] = E[∆1θ1θ>] = 1
N2¯
ME[(∆1Mθ)(∆1M θ)>]¯
M
=1
N2¯
MEhN
X
α=1
L
X
k=1
(∆ξ(k)
α,θ)¯
ξ(k)
α
N
X
β=1
L
X
l=1
(∆ξ(l)
β,θ)¯
ξ(l)>
βi¯
M
=1
N2¯
M
N
X
α,β=1
L
X
k,l=1
(θ, E[∆ξ(k)
αξ(l)>
β]θ)¯
ξ(k)
α¯
ξ(l)>
β¯
M
=1
N2¯
M³N
X
α=1
L
X
k,l=1
(θ, V (kl)[ξα]θ)¯
ξ(k)
α¯
ξ(l)>
α´¯
M
=1
N¯
M¯
M0¯
M,(43)
where we define
¯
M0=1
N
N
X
α=1
L
X
k,l=1
(θ, V (kl)[ξα]θ)¯
ξ(k)
α¯
ξ(l)>
α(44)
In the above derivation, we have noted that from our noise assumption we have
E[∆1ξ(k)
α1ξ(l)>
β] = δαβ V(kl)[ξα], where δαβ is the Kronecker delta.
8.2 Bias Analysis
The important observation is that the covariance matrix V[θ] does not contain
N. Thus, all algebraic methods have the same covariance matrix in the leading
order, as pointed out by Al-Sharadqah and Chernov1) for circle fitting. This
observation leads us to focus on the bias. We now seek an Nthat reduces the
bias as much as possible. It would be desirable if we could find such an Nthat
minimizes the total mean square error E[k1θ+2θ+···k2], but at the moment
this seems to be an intractable problem; minimizing the bias alone is a practical
compromise, whose effectiveness is tested by experiments a posteriori.
From Eq. (38), we see that the first order bias E[∆1θ] is 0, hence the leading
bias is E[∆
2θ]. From Eq. (42), we have
E[∆
2θ] = (¯
θ, E[T]¯
θ)
(¯
θ,N¯
θ)¯
MN¯
θ¯
ME[T]¯
θ.(45)
We now evaluate the expectation E[T] of Tin Eq. (40). From Eq. (32), we see
that E[∆2M] is given by
E[∆2M] = 1
N
N
X
α=1
L
X
k=1³¯
ξ(k)
αE[∆2ξ(k)
α]>
+E[∆1ξ(k)
α1ξ(k)>
α]+E[∆2ξ(k)
α]¯
ξ(k)>
α´
=1
N
N
X
α=1
L
X
k=1³V(kk)[ξα]+2S[¯
ξ(k)
αe(k)>
α]´,(46)
86
where we have used Eq. (20) and defined
e(k)
αE[∆2ξ(k)
α].(47)
The operator S[·] denotes symmetrization (S[A]=(A+A>)/2). The expecta-
tion E[∆1M¯
M1M] has the following form (see Appendix):
E[∆1M¯
M1M] = 1
N2
N
X
α=1
L
X
k,l=1³tr[ ¯
MV(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
α
+(¯
ξ(k)
α,¯
M¯
ξ(l)
α)V(kl)[ξα] + 2S[V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
α]´.(48)
From Eq. (46) and Eq. (48), the expectation of Tis
E[T] = NT1
N2
N
X
α=1
L
X
k,l=1³tr[ ¯
MV(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
α
+(¯
ξ(k)
α,¯
M¯
ξ(l)
α)V(kl)[ξα]+2S[V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
α]´,(49)
where we put
NT=1
N
N
X
α=1
L
X
k=1³V(kk)[ξα]+2S[¯
ξ(k)
αe(k)>
α]´.(50)
9. HyperLS
Now, let us choose Nto be the expression E[T] in Eq. (49) itself, namely,
N=NT1
N2
N
X
α=1
L
X
k,l=1³tr[ ¯
MV(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
α+(¯
ξ(k)
α,¯
M¯
ξ(l)
α)V(kl)[ξα]
+2S[V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
α]´,(51)
Letting N=E[T] in Eq. (45), we see that
E[∆
2θ] = ¯
M³(¯
θ,N¯
θ)
(¯
θ,N¯
θ)NN´¯
θ=0.(52)
Since the right-hand side of Eq. (49) contains the true values ¯
ξαand ¯
M, we
replace ¯
xαin their definition by the observation xα. This does not affect the
result, since the odd order noise terms have expectation 0 and hence the resulting
error in E[∆
2θ] is of the fourth order. Thus, the second order bias is exactly 0.
In fitting a circle to a point sequence, Al-Sharadqah and Chernov1) proposed
to choose N=NTand showed that the second order bias E[∆
2θ] is zero up to
O(1/N2). They called their method Hyper. What we have shown here is that
the second order bias is completely removed by including the second term on the
right-hand side of Eq. (51). We call our scheme HyperLS.
Note that Nhas scale indeterminacy: If Nis multiplied by c(6= 0), Eq. (16)
has the same solution θ; only λis divided by c. Thus, the noise characteristics
V(kl)[ξα] in Eq. (20) and hence V[xα] need to be known only up to scale; we need
not know the absolute magnitude of the noise.
For numerical computation, standard linear algebra routines for solving the
generalized eigenvalue problem of Eq. (16) assume that Nis positive definite,
but here Nis nondefinite. This causes no problem, because Eq. (16) can be
written as
Nθ =1
λMθ.(53)
As mentioned earlier, the matrix Min Eq. (14) is positive definite for noisy
data, so we can solve Eq. (53) instead, using a standard routine. If the smallest
eigenvalue of Mhappens to be 0, it indicates that the data are all exact, so
any method, e.g., the standard LS, gives an exact solution. For noisy data, the
solution θis given by the eigenvector of Eq. (53) for the eigenvalue 1with the
largest absolute value.
Example 7 (Ellipse fitting). If the noise in (xα, yα) is independent and Gaus-
sian with mean 0 and standard deviation σ, the vector eα(= e(1)) in Eq. (47) is
given by
eα=σ2(1,0,1,0,0,0)>.(54)
Hence, the matrix NTin Eq. (50) is given by
87
NT=σ2
N
N
X
α=1
6x2
α6xαyαx2
α+y2
α6xα2yα1
6xαyα4(x2
α+y2
α) 6xαyα4yα4xα0
x2
α+y2
α6xαyα6y2
α2xα6yα1
6xα4yα2xα4 0 0
2yα4xα6yα0 4 0
1 0 1 0 0 0
.(55)
The Taubin method26) is to use as N
NTaubin =4σ2
N
N
X
α=1
x2
αxαyα0xα0 0
xαyαx2
α+y2
αxαyαyαxα0
0xαyαy2
α0yα0
xαyα0 1 0 0
0xαyα0 1 0
0 0 0 0 0 0
,(56)
which we see is obtained by letting eα=0in Eq. (50). As pointed out earlier,
the value of σin Eq. (55) and Eq. (56) need not be known. Hence, we can simply
let σ= 1 in Eq. (16) and Eq. (53) in actual computation.
Example 8 (Fundamental matrix computation). If the noise in (xα, yα)
and (x0
α, y0
α) is independent and Gaussian with mean 0 and standard deviation
σ, the vector eα(= e(1)) in Eq. (47) is 0, so the NTin Eq. (50) becomes
NT=σ2
N
N
X
α=1
x2
α+x02
αx0
αy0
αx0
αxαyα0 0 xα0 0
x0
αy0
αx2
α+y02
αy0
α0xαyα0 0 xα0
x0
αy0
α1 0 0 0 0 0 0
xαyα0 0 y2
α+x02
αx0
αy0
αx0
αyα0 0
0xαyα0x0
αy0
αy2
α+y02
αy0
α0yα0
0 0 0 x0
αy0
α1 0 0 0
xα0 0 yα0 0 1 0 0
0xα0 0 yα0 0 1 0
0 0 0 0 0 0 0 0 0
.
(57)
It turns out that the use of this matrix NTcoincides with the well known Taubin
method26) . As in ellipse fitting, we can let σ= 1 in Eq. (57) in actual computa-
tion.
Example 9 (Homography computation). If the noise in (xα, yα) and (x0
α, y0
α)
is independent and Gaussian with mean 0 and standard deviation σ, the vectors
e(k)
αin Eq. (47) are all 0, so the NTin Eq. (50) becomes
NT=σ2
N
N
X
α=1
x2
α+y02
α+ 1 xαyαxαx0
αy0
α0
xαyαy2
α+y02
α+ 1 yα0x0
αy0
α
xαyα1 0 0
x0
αy0
α0 0 x2
α+x02
α+ 1 xαyα
0x0
αy0
α0xαyαy2
α+x02
α+ 1
0 0 0 xαyα
x0
α0 0 y0
α0
0x0
α0 0 x0
α
0 0 0 0 0
0x0
α0 0
0 0 x0
α0
0 0 0 0
xαy0
α0 0
yα0y0
α0
1 0 0 0
0x2
α+x02
α+y02
α2xαyα2xα
0 2xαyαy2
α+x02
α+y02
α2yα
0 2xα2yα2
,(58)
For homography computation, the constraint is a vector equation in Eq. (8).
Hence, the Taubin method26), which is defined for a single constraint equation,
cannot be applied. However, the use of the above NTas Nplays the same role
of the Taubin method26) for ellipse fitting and fundamental matrix computation,
as first pointed out by Rangarajan and Papamichalis24). As before, we can let σ
= 1 in the matrix NTin actual computation.
In the following, we call the use of NTas Nthe Taubin approximation. For
88
fundamental matrix computation, it coincides with the Taubin method26) , but
for homography computation the Taubin method was not defined. For ellipse
fitting, the Taubin method and the Taubin approximation are slightly different;
the Taubin method is equivalent to use only the first term on the right hand
side of Eq. (50). For circle fitting, the Taubin approximation is the same as the
“Hyper” of Al-Sharadqah and Chernov1).
10. Numerical Experiments
We did the following three experiments:
Ellipse fitting: We fit an ellipse to the point sequence shown in Fig. 2(a). We
took 31 equidistant points on the first quadrant of an ellipse with major and
minor axes 100 and 50 pixels, respectively.
Fundamental matrix computation: We compute the fundamental matrix
between the two images shown in Fig. 2(b), which view a cylindrical grid
surface from two directions. The image size is assumed to be 600 ×600
(pixels) with focal lengths 600 pixels for both. The 91 grid points are used
as corresponding points.
Homography computation: We compute the homography relating the two
images shown in Fig. 2(c), which view a planar grid surface from two direc-
tions. The image size is assumed to be 800 ×800 (pixels) with focal lengths
600 pixels for both. The 45 grid points are used as corresponding points.
In all experiments, we divided the data coordinate values by 600 (pixels) (i.e., we
used 600 pixels as the unit of length) to make all the data values within the range
of about ±1. This is for stabilizing numerical computation with finite precision
(a) (b) (c)
Fig. 2 (a) 31 points on an ellipse. (b) Two views of a curved grid. (c) Two views of a planar grid.
θ
θ
θ
O
Fig. 3 The true value ¯
θ, the computed value θ, and its
orthogonal component ∆θto ¯
θ.
length; without this data scale normalization, serious accuracy loss is incurred,
as pointed out by Hartley5) for fundamental matrix computation.
For each example, we compared the standard LS, HyperLS, its Taubin approx-
imation, and ML, for which we used the FNS of Chojnacki et al.4) for ellipse
fitting and fundamental matrix computation and the multiconstraint FNS of Ni-
itsuma et al.22) for homography computation. As mentioned in Section 4, FNS
minimizes not directly Eq. (11) but the Sampson error in Eq. (21), and the exact
ML solution can be obtained by repeated Sampson error minimization16) . The
the solution that minimizes the Sampson error usually agrees with the ML solu-
tion up to several significant digits11),14),15). Hence, FNS can safely be regarded
as minimizing Eq. (11).
Let ¯
θbe the true value of the parameter θ, and ˆ
θits computed value. We
consider the following error:
θ=P¯
θˆ
θ,P¯
θI¯
θ¯
θ>.(59)
The matrix P¯
θrepresents the orthogonal projection onto the space orthogonal
to ¯
θ. Since the computed value ˆ
θis normalized to a unit vector, it distributes
around ¯
θon the unit sphere. Hence, the meaningful deviation is its component
orthogonal to ¯
θ, so we measure the error component in the tangent space to the
unit sphere at ¯
θ(Fig. 3).
We added independent Gaussian noise of mean 0 and standard deviation σto
the xand ycoordinates of data each point and repeated the fitting Mtimes for
each σ, using different noise. We let M= 10000 for ellipse fitting and fundamen-
tal matrix computation and M= 1000 for homography computation. Then, we
89
0.1
0.2
0.3
0.4
0.5
0.6
0 1 2 3
123
4
σ
(a) (b) (c)
Fig. 4 RMS error vs. the standard deviation σof the noise added to each point: 1. standard LS, 2. Taubin approximation, 3.
HyperLS, 4. ML. The dotted lines indicate the KCR lower bound. (a) Ellipse fitting. (b) Fundamental matrix computation.
(c) Homography computation.
evaluated the root-mean-square (RMS) error
E=v
u
u
t
1
M
M
X
a=1
kθ(a)k2,(60)
where ∆θ(a)is the value of ∆θin the ath trial. The theoretical accuracy limit,
called the KCR lower bound3),8),10) , is given by
E[∆θθ>]Âσ2
N³1
N
N
X
α=1
L
X
k,l=1
¯
W(kl)
α¯
ξ(k)
α¯
ξ(l)>
α´
VKCR[θ],(61)
where ¯
W(kl)
αis the value of W(kl)
αin Eq. (21) evaluated by assuming σ= 1 and
using the true values ¯
θand ¯
ξ(kl)
α. The relation Âmeans that the left-hand side
minus the right-hand side is a positive semidefinite symmetric matrix, and the
operation ( ·)denotes pseudoinverse. We compared the RMS error in Eq. (60)
with the trace Eq. (61):
qE[kθk2]ptrVKCR[θ].(62)
Figure 4 plots for σthe RMS error of Eq. (60) for each method and the KCR
lower bound of Eq. (62).
We also compared the reprojection error for different methods. According to
statistics, the reprojection error Iin Eq. (11) for ML is subject to a χ2distribu-
tion with rN ddegrees of freedom, where ris the codimension of the constraint
and dis the dimension of the parameters8). Hence, if ML is computed by assum-
ing σ= 1, the square root of the average reprojection error per datum is expected
to be σprd/N.Figure 5 plots the square root of the average, per datum, of
the computed reprojection error, which was approximated by the Sampson error
K(θ) in Eq. (21), along with the theoretical expectation.
We observe the following:
Ellipse fitting: The standard LS performs poorly, while ML exhibits the high-
est accuracy, almost reaching the KCR lower bound. However, ML compu-
tation fails to converge above a certain noise level. In contrast, HyperLS
produces, without iterations, an accurate solution close to ML. The accuracy
of its Taubin approximation is practically the same as the traditional Taubin
method and is slightly lower than HyperLS. Since r= 1, d= 5, the square
root of the average reprojection error per datum has theoretical expectation
σp15/N to a first approximation. We see that the computed value almost
coincides with the expected value expect for the standard LS.
Fundamental matrix computation: Again, the standard LS is poor, while
ML has the highest accuracy, almost reaching the KCR lower bound. The
accuracy of HyperLS is very close to ML. Its Taubin approximation (= the
traditional Taubin method) has practically the same accuracy as HyperLS.
90
0.2
0.4
0.6
0.8
0 0.2 0.4 0.6 0.8
σ
1 4
2 3
1
2
3
0 1 2 3
σ
1
2 3 4
(a) (b) (c)
Fig. 5 Root square average reprojection error per datum vs. the standard deviation σof the noise added to each point: 1. standard
LS, 2. Taubin approximation, 3. HyperLS, 4. ML. The dotted lines indicate theoretical expectation. (a) Ellipse fitting.
(b) Fundamental matrix computation. (c) Homography computation.
The fundamental matrix has the constraint that its rank be 2. The compari-
son here is done before the rank constraint is imposed. Hence, r= 1, d= 8,
and the square root of the average reprojection error per datum is expected
to be σp18/N. We see that the computed value almost coincides with
the expected value expect for the standard LS.
Homography computation: In this case, too, the standard LS is poor, while
ML has the highest accuracy, almost reaching the KCR lower bound. How-
ever, ML computation fails to converge above a certain noise level. The
accuracy of HyperLS is very close to ML. Its Taubin approximation has prac-
tically the same accuracy as HyperLS. Since r= 2, d= 4, the square root of
the average reprojection error per datum is expected to be σp2(1 4/N).
We see that the computed value almost coincides with the expected value
expect for the standard LS.
In all examples, the standard LS performs poorly, while ML provides the highest
accuracy. Note that the differences among different methods are more marked
when measured in the RMS error than in the reprojection error. This is because
the RMS error measures the error of the parameters of the equation, while the
reprojection error measures the closeness of the fit to the data. For ellipse fit-
ting, for example, the RMS error compares the fitted ellipse equation and the
true ellipse equation, while the reprojection error measures how close the fitted
ellipse is to the data. As a result, even if two ellipse have nearly the same dis-
tances to the data, their shapes can be very different. This difference becomes
more conspicuous as the data cover a shorter segment of the ellipse. The same
observation can be done for fundamental matrix computation and homography
computation.
We also see from our experiments that ML computation may fail in the presence
of large noise. The convergence of ML critically depends on the accuracy of the
initialization. In the above experiments, we used the standard LS to start the
FNS iterations. We confirmed that the use of HyperLS to start the iterations
significantly extends the noise range of convergence, though the computation fails
sooner or later. On the other hand, HyperLS is algebraic and hence immune to
the convergence problem, producing a solution close in accuracy to ML in any
noise level.
The Taubin approximation is clearly inferior to HyperLS for ellipse fitting but is
almost equivalent to HyperLS for fundamental matrices and homographies. This
reflects the fact that while ξis quadratic in xand yfor ellipses (see Eq. (4)),
the corresponding ξand ξ(k)are bilinear in x,y,x0, and y0for fundamental
matrices (see Eq. (6)) and homographies (see Eq. (9)), so e(k)
αin Eq. (47) is 0.
In structure-from-motion applications, we frequently do inference from multiple
images based on “multilinear” constraints involving homographies, fundamental
matrices, trifocal tensors, and other geometric quantities7). For such problems,
the constraint itself is nonlinear but is linear in observations of each image.
91
Then, e(k)
α=0, because noise in different images are assumed to be independent.
In such a problem, the accuracy of HyperLS is nearly the same as its Taubin
approximation. However, HyperLS is clearly superior in a situation where the
constraint involves nonlinear terms in observations of the same image, e.g., ellipse
fitting.
11. Concluding Remarks
We have presented a general formulation for a special type of least squares
(LS) estimator, which we call “HyperLS,” for geometric problems that frequently
appear in vision applications. We described the problem in the most general
terms and discussed various theoretical issues that have not been fully studied
so far. In particular, we pointed out that the characteristics of image-based
inference is very different to the conventional statistical domains and discussed
in detail various issues related to ML and algebraic fitting. Then, we derived
HyperLS by introducing a normalization that eliminates statistical bias of LS up
to second order noise terms.
It would be ideal if we could minimize the total mean squares error by taking
all higher order terms into account. Due to technical difficulties, we limited our
attention to the bias up to the second order. Also, we introduced in our deriva-
tion several assumptions about the choice of the eigenvalues and the convergence
of series expansion. However, the purpose of this paper is not to establish math-
ematical theorems with formal proofs. Our aim is to derive techniques that are
useful in practical problems; the usefulness is to be tested by experiments.
Our numerical experiments for computing ellipses, fundamental matrices, and
homographies showed that HyperLS yields a solution far superior to the stan-
dard LS and comparable in accuracy to ML, which is known to produce highly
accurate solutions but may fail to converge if poorly initialized. Thus, HyperLS
is a perfect candidate for ML initialization. We compared the performance of
HyperLS and its Taubin approximation and attributed the performance differ-
ences to the structure of the problem. In this paper, we did not show real image
demos, concentrating on the general mathematical framework, because particular
applications have been shown elsewhere1),12),22),23).
Acknowledgments The authors thank Ali Al-Sharadqah and Nikolai Cher-
nov of the University of Alabama at Birmingham, U.S.A, Wolfgang F¨orstner
of the University of Bonn, Germany, and Alexander Kukush of National Taras
Shevchenko University of Kyiv, Ukraine, for helpful discussions. This work was
supported in part by the Ministry of Education, Culture, Sports, Science, and
Technology, Japan, under a Grant in Aid for Scientific Research (C 21500172).
References
1) Al-Sharadqah, A. and Chernov, N.: Error analysis for circle fitting algorithms.
Elec. J. Stat., Vol.3, pp.886–911 (2009).
2) Cheng, C.-L. and Kukush, A.: Non-existence of the first moment of the adjusted
least squares estimator in multivariate errors-in-variables model, Metrika, Vol.64,
No.1, pp.41–46 (2006).
3) Chernov, N. and Lesort, C.: Statistical efficiency of curve fitting algorithms, Com-
put. Stat. Data Anal., Vol.47, Vol.4, pp.713–728 (2004).
4) Cho jnacki, W., Brooks, M.J., van den Hengel, A. and Gawley, D.: On the fitting
of surfaces to data with covariances, IEEE Trans. Patt. Anal. Mach. Intell., Vol.22,
No.11, pp.1294–1303 (2000).
5) Hartley, R.I.: In defense of the eight-point algorithm, IEEE Trans. Patt. Anal.
Mach. Intell., Vol.19, No.6, pp.580–593 (1997).
6) Hartley, R. and Kahl, F.: Optimal algorithms in multiview geometry, Proc. 8th
Asian Conf. Computer Vision, Tokyo, Japan, November 2007, Vol.1, pp.13–34
(2007).
7) Hartley, R. and Zisserman, A.: Multiple View Geometry in Computer Vision, 2nd
ed., Cambridge University Press, Cambridge, U.K. (2004).
8) Kanatani, K.: Statistical Optimization for Geometric Computation: Theory and
Practice, Elsevier Science, Amsterdam, The Netherlands (1996); reprinted, Dover,
New York, U.S.A. (2005).
9) Kanatani, K.: Ellipse fitting with hyperaccuracy, IEICE Trans. Inf. & Syst.,
Vol.E89-D, No.10, pp.2653–2660 (2006).
10) Kanatani, K.: Statistical optimization for geometric fitting: Theoretical accuracy
analysis and high order error analysis, Int. J. Comput. Vis., Vol.80, No.2, pp.167–
188 (2006).
11) Kanatani, K. and Niitsuma, H.: Optimal two-view planar scene triangulation, IPSJ
Trans. Comput. Vis. Appl., Vol.4 (2011), to appear.
12) Kanatani, K. and Rangara jan, P.: Hyper least squares fitting of circles and ellipses,
Comput. Stat. Data Anal., Vol.55, Vol.6, pp.2197–2208 (2011).
13) Kanatani, K. and Sugaya, Y.: Performance evaluation of iterative geometric fitting
algorithms, Comp. Stat. Data Anal., Vol.52, No.2, pp.1208–1222 (2007).
92
14) Kanatani, K. and Sugaya, Y.: Compact algorithm for strictly ML ellipse fitting,
Proc. 19th Int. Conf. Pattern Recog., Tampa, FL, U.S.A. (2008).
15) Kanatani, K. and Sugaya, Y.: Compact fundamental matrix computation, IPSJ
Trans. Comput. Vis. Appl., Vol.2, pp.59–70 (2010).
16) Kanatani, K. and Sugaya, Y.: Unified computation of strict maximum likelihood
for geometric fitting, J. Math. Imaging Vis., Vol.38, pp.1–13 (2010).
17) Kukush, K., Markovski, I. and Van Huffel, S.: Consistent fundamental matrix
estimation in a quadratic measurement error model arising in motion analysis,
Comp. Stat. Data Anal., Vol.41, No.1, pp.3–18 (2002).
18) Kukush, K., Markovski, I. and Van Huffel, S.: Consistent estimation in an implicit
quadratic measurement error model, Comp. Stat. Data Anal., Vol.47, No.1, pp.123–
147 (2004).
19) Leedan, Y. and Meer, P.: Heteroscedastic regression in computer vision: Problems
with bilinear constraint, Int. J. Comput. Vis., Vol.37, No.2, pp.127–150 (2000).
20) Matei, B.C. and Meer, P.: Estimation of nonlinear errors-in-variables models for
computer vision applications, IEEE Trans. Patt. Anal. Mach. Intell., Vol.28, No.10,
pp.1537–1552 (2006).
21) Neyman, J. and Scott, E.L.: Consistent estimates based on partially consistent
observations, Econometrica, Vol.16, No.1, pp.1–32 (1948).
22) Niitsuma, H., Rangara jan, P. and Kanatani, K.: High accuracy homography com-
putation without iterations, Proc. 16th Symp. Sensing Imaging Inf., Yokohama,
Japan (2010).
23) Rangara jan, P. and Kanatani, K.: Improved algebraic methods for circle fitting,
Elec. J. Stat., Vol.3, pp.1075–1082 (2009).
24) Rangara jan, P. and Papamichalis, P.: Estimating homographies without normal-
ization, Proc. Int. Conf. Image Process., Cairo, Egypt, pp.3517–3520 (2009).
25) Sampson, P.D.: Fitting conic sections to “very scattered” data: An iterative refine-
ment of the Bookstein algorithm, Comput. Graphics Image Process., Vol.18, No.1,
pp.97–108 (1982).
26) Taubin, G.: Estimation of planar curves, surfaces, and non-planar space curves de-
fined by implicit equations with applications to edge and range image segmentation,
IEEE Trans. Patt. Anal. Mach. Intell., Vol.13, No.11, pp.1115–1138 (1991).
27) Triggs, B., McLauchlan, P.F., Hartley, R.I. and Fitzgibbon, A.: Bundle
adjustment—A modern synthesis, Vision Algorithms: Theory and Practice, Triggs,
B, Zisserman, A. and Szeliski, R. (Eds.), pp.298–375, Springer (2000).
Appendix
The term E[∆1M¯
M1M] is computed as follows:
E[∆1M¯
M1M]
=Eh1
N
N
X
α=1
3
X
k=1³¯
ξ(k)
α1ξ(k)>
α+ ∆1ξ(k)
α¯
ξ(k)>
α´¯
M1
N
N
X
β=1
3
X
l=1 ³¯
ξ(l)
β1ξ(l)>
β
+∆1ξ(l)
β¯
ξ(l)>
β´i
=1
N2
N
X
α,β=1
3
X
k,l=1
E[(¯
ξ(k)
α1ξ(k)>
α+∆1ξ(k)
α¯
ξ(k)>
α)¯
M(¯
ξ(l)
β1ξ(l)>
β+∆1ξ(l)
β¯
ξ(l)>
β)]
=1
N2
N
X
α,β=1
3
X
k,l=1
E[¯
ξ(k)
α1ξ(k)>
α¯
M¯
ξ(l)
β1ξ(l)>
β+¯
ξ(k)
α1ξ(k)>
α¯
M1ξ(l)
β¯
ξ(l)>
β
+∆1ξ(k)
α¯
ξ(k)>
α¯
M¯
ξ(l)
β1ξ(l)>
β+ ∆1ξ(k)
α¯
ξ(k)>
α¯
M1ξ(l)
β¯
ξ(l)>
β]
=1
N2
N
X
α,β=1
3
X
k,l=1
E[¯
ξ(k)
α(∆1ξ(k)
α,¯
M¯
ξ(l)
β)∆1ξ(l)>
β+¯
ξ(k)
α(∆1ξ(k)
α,¯
M1ξ(l)
β)¯
ξ(l)>
β
+∆1ξ(k)
α(¯
ξ(k)
α,¯
M¯
ξ(l)
β)∆1ξ(l)>
β+ ∆1ξ(k)
α(¯
ξ(k)
α,¯
M1ξ(l)
β)¯
ξ(l)>
β]
=1
N2
N
X
α,β=1
3
X
k,l=1
E[(∆1ξ(k)
α,¯
M¯
ξ(l)
β)¯
ξ(k)
α1ξ(l)>
β+(∆1ξ(k)
α,¯
M1ξ(l)
β)¯
ξ(k)
α¯
ξ(l)>
β
+(¯
ξ(k)
α,¯
M¯
ξ(l)
β)∆1ξ(k)
α1ξ(l)>
β+ ∆1ξ(k)
α(¯
M1ξ(l)
β,¯
ξ(k)
α)¯
ξ(l)>
β]
=1
N2
N
X
α,β=1
3
X
k,l=1
E[¯
ξ(k)
α(( ¯
M¯
ξ(l)
β)>1ξ(k)
α)∆1ξ(l)>
β
+tr[ ¯
M1ξ(l)
β1ξ(k)>
α]¯
ξ(k)
α¯
ξ(l)>
β+ (¯
ξ(k)
α,¯
M¯
ξ(l)
β)∆1ξ(k)
α1ξ(l)>
β
+∆1ξ(k)
α(∆1ξ(l)>
β¯
M¯
ξ(k)
α)¯
ξ(l)>
β]
=1
N2
N
X
α,β=1
3
X
k,l=1³¯
ξ(k)
α¯
ξ(l)>
β¯
ME[∆1ξ(k)
α1ξ(l)>
β]
+tr[ ¯
ME[∆1ξ(l)
β1ξ(k)>
α]]¯
ξ(k)
α¯
ξ(l)>
β+ (¯
ξ(k)
α,¯
M¯
ξ(l)
β)E[∆1ξ(k)
α1ξ(l)>
β]
+E[∆1ξ(k)
α1ξ(l)>
β]¯
M¯
ξ(k)
α¯
ξ(l)>
β´
=1
N2
N
X
α,β=1
3
X
k,l=1³¯
ξ(k)
α¯
ξ(l)>
β¯
Mδαβ V(kl)[ξα] + tr[ ¯
Mδαβ V(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
β
93
+(¯
ξ(k)
α,¯
M¯
ξ(l)
β)δαβ V(kl)[ξα] + δαβ V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
β´
=1
N2
N
X
α=1
3
X
k,l=1³¯
ξ(k)
α¯
ξ(l)>
α¯
MV(kl)[ξα] + tr[ ¯
MV(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
α
+(¯
ξ(k)
α,¯
M¯
ξ(l)
α)V(kl)[ξα] + V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
α´
=1
N2
N
X
α=1
3
X
k,l=1³tr[ ¯
MV(kl)[ξα]]¯
ξ(k)
α¯
ξ(l)>
α+ (¯
ξ(k)
α,¯
M¯
ξ(l)
α)V(kl)[ξα]
+2S[V(kl)[ξα]¯
M¯
ξ(k)
α¯
ξ(l)>
α]´.(63)
Thus, Eq. (48) is obtained.
(Received January 1, 2011)
(Revised May 16, 2011)
(Accepted August 1, 2011)
(Released October 17, 2011)
(Communicated by Peter Sturm)
Kenichi Kanatani received his B.E., M.S., and Ph.D. in applied
mathematics from the University of Tokyo in 1972, 1974 and 1979,
respectively. After serving as Professor of computer science at
Gunma University, Gunma, Japan, he is currently Professor of
computer science at Okayama University, Okayama, Japan. He is
the author of many books on computer vision and received many
awards including the best paper awards from IPSJ (1987) and
IEICE (2005). He is an IEEE Fellow.
Prasanna Rangarajan received his B.E. in electronics and com-
munication engineering from Bangalore University, Bangalore, In-
dia, in 2000 and his M.S. in electrical engineering from Columbia
University, New York, NY, U.S.A., in 2003. He is currently a
Ph.D. candidate in electrical engineering at Southern Methodist
University, Dallas, TX, U.S.A. His research interests include im-
age processing, structured illumination and parameter estimation
for computer vision.
Yasuyuki Sugaya received his B.E., M.S., and Ph.D. in com-
puter science from the University of Tsukuba, Ibaraki, Japan, in
1996, 1998, and 2001, respectively. From 2001 to 2006, he was
Assistant Professor of computer science at Okayama University,
Okayama, Japan. Currently, he is Associate Professor of informa-
tion and computer sciences at Toyohashi University of Technology,
Toyohashi, Aichi, Japan. His research interests include image pro-
cessing and computer vision. He received the IEICE best paper award in 2005.
Hirotaka Niitsuma received his B.E. and M.S. in applied physics
from Osaka University, Japan, in 1993 and 1995, respectively, and
his Ph.D. in informatic science from NAIST, Japan, in 1999. He
was a researcher at TOSHIBA, at JST Corporation, at Denso IT
Laboratory, Inc., at Kwansei Gakuin University, Japan, at Kyung-
pook National University, Korea, and at AIST, Japan. From April
2007, he is Assistant Professor of computer science at Okayama
University, Japan. His research interests include computer vision, machine learn-
ing, and neural networks.
94
... A recent series of papers by Kanatani and coworkers [8][9][10] have used a perturbative treatment to predict the systematic errors introduced into the fitting process by noisy data, and hence to identify analytically (rather than merely numerically) the choices of normalisation and weighting that minimise bias and statistical error. These are forward statistical calculations, that is, the noisy data is described by a sampling distribution around the true values; the final result is an unbiased point estimate of the generic conic coefficients. ...
Article
Full-text available
A perturbative approach is used to quantify the effect of noise in data points on fitted parameters in a general homogeneous linear model, and the results applied to the case of conic sections. There is an optimal choice of normalisation that minimises bias, and iteration with the correct reweighting significantly improves statistical reliability. By conditioning on an appropriate prior, an unbiased type-specific fit can be obtained. Error estimates for the conic coefficients may also be used to obtain both bias corrections and confidence intervals for other curve parameters.
Article
This paper presents a method of significantly improving the previously proposed simple, flexible, and accurate phase retrieval algorithm for the random phase-shifting interferometry named HEFS [ K. Yatabe , J. Opt. Soc. Am. A34, 87 (2017)]. Although its effectiveness and performance were confirmed by numerical experiments in the original paper, it is found that the algorithm may not work properly if observed fringe images contains untrusted (extremely noisy) pixels. In this paper, a method of avoiding such untrusted pixels within the estimation processes of HEFS is proposed for the practical use of the algorithm. In addition to the proposal, an experiment of measuring a sound field in air was conducted to show the performance for real data, where the proposed improvement is critical for that situation. MATLAB codes (which can be downloaded from http://goo.gl/upcsFe) are provided within the paper to aid understanding the main concept of the proposed methods.
Article
This paper presents a non-iterative phase retrieval method from randomly phase-shifted fringe images. By combining the hyperaccurate least squares ellipse fitting method with the subspace method (usually called the principal component analysis), a fast and accurate phase retrieval algorithm is realized. The proposed method is simple, flexible, and accurate. It can be easily coded without iteration, initial guess, or tuning parameter. Its flexibility comes from the fact that totally random phase-shifting steps and any number of fringe images greater than two are acceptable without any specific treatment. Finally, it is accurate because the hyperaccurate least squares method and the modified subspace method enable phase retrieval with a small error as shown by the simulations. A MATLAB code, which is used in the experimental section, is provided within the paper to demonstrate its simplicity and easiness.
Chapter
Extracting elliptic edges from images and fitting ellipse equations to them is one of the most fundamental tasks of computer vision. This is because circular objects, which are very common in daily scenes, are projected as ellipses in images. We can even compute their 3D positions from the fitted ellipse equations, which along with other applications are discussed in Chap. 9. In this chapter, we present computational procedures for accurately fitting an ellipse to extracted edge points by considering the statistical properties of image noise. The approach is classified into algebraic and geometric. The algebraic approach includes least squares, iterative reweight, the Taubin method, renormalization, HyperLS, and hyper-renormalization; the geometric approach includes FNS, geometric distance minimization, and hyper-accurate correction. We then describe the ellipse-specific method of Fitzgibbon et al. and the random sampling technique for avoiding hyperbolas, which may occur when the input information is insufficient. The RANSAC procedure is also discussed for removing nonelliptic arcs from the extracted edge point sequence.
Chapter
If we take two images of a planar surface from two different places, the two images are related by a mapping called a homography. Computing a homography from point correspondences over two images is one of the most fundamental processes of computer vision. This is because, among other things, the 3D positions of the planar surface we are viewing and the two cameras that took the images can be computed from the computed homography. Such applications are discussed in Chaps. 7 and 8. This chapter describes the principles and typical computational procedures for accurately computing the homography by considering the statistical properties of the noise involved in correspondence detection. As in ellipse fitting and fundamental matrix computation, the methods are classified into algebraic (least squares, iterative reweight, the Taubin method, renormalization, HyperLS, and hyper-renormalization) and geometric (FNS, geometric distance minimization, and hyperaccurate correction). We also describe the RANSAC procedure for removing wrong correspondences (outliers).
Article
The technique of "renormalization" for geometric estimation attracted much attention when it appeared in early 1990s for having higher accuracy than any other then known methods. The key fact is that it directly specifies equations to solve, rather than minimizing some cost function. This paper expounds this "non-minimization approach" in detail and exploits this principle to modify renormalization so that it outperforms the standard reprojection error minimization. Doing a precise error analysis in the most general situation, we derive a formula that maximizes the accuracy of the solution; we call it hyper-renormalization. Applying it to ellipse fitting, fundamental matrix computation, and homography computation, we confirm its accuracy and efficiency for sufficiently small noise. Our emphasis is on the general principle, rather than on individual methods for particular problems.
Article
Full-text available
Extracting edge points from an image and fitting ellipses to them is a fundamental technique for computer vision applications. However, since the extracted edge points sometimes contain non-elliptic arcs such as line segments, it is a very difficult to extract only elliptic arcs from them. In this paper, we propose a new method for extracting elliptic arcs from a spatially connected point sequence. We first fit an ellipse to an input point sequence and segment the sequence into partial arcs at the intersection points of the fitted ellipse. Next, we compute residuals of the fitted ellipse for all input points and select elliptic arcs among the segmented arcs by checking the curvatures of the residual graph. Then, we fit an ellipse to the selected arcs and repeat the above process until the selected arcs do not change. By using simulated data and real images, we compare the performance of our method with existing methods and show the efficiency of our proposed method.
Article
Full-text available
Inspired by the circle fitting algorithm "Hyper" of Al-Sharadqah and Chernov (1), which eliminates the second order "essen-tial bias" that excludes terms of O(σ 2 /N 2), we extend their analysis and show that by a small modification the second order bias can be eliminated completely. By numerical experiments, we show that this results in better performance when the number N of points is small and the noise is large.
Article
Full-text available
An examination of the geometric interpretation of the error-of-fit measure of the Bookstein algorithm for fitting conic sections shows why it may not be entirely satisfoctory when the data are “very scattere” in the sense that the data points are distributed rather widely about an underlying smooth curve. A simple iterative refinement of the Bookstein al;gorithm, similar in spirit to iterative weighted least-squares methods in regression analysis, results in a fitted conic section that approximates the conic that would minimize the sum of squared orthogonal distances of data points from the fitted conic. The usefulness and limitations of the refined algorithm are demonstrated on two different types of “very scattered” data.
Article
A very small algorithm is presented for computing the fundamental matrix from point correspondences over two images. The computation is based on the strict maximum likelihood (ML) principle, minimizing the reprojection error. The rank constraint is incorporated by the EFNS procedure. Although our algorithm produces the same solution as all existing ML-based methods, it is probably the smallest of all. By numerical experiments, we confirm that our algorithm behaves as expected.
Conference Paper
A very compact algorithm is presented for fitting an ellipse to points in images by maximum likelihood (ML) in the strict sense. Although our algorithm produces the same solution as existing ML-based methods, it is probably the simplest and the smallest of all. By numerical experiments, we show that the strict ML solution practically coincides with the Sampson solution.
Conference Paper
Existing linear methods for estimating homographies, rely on coordinate normalization, to reduce the error in the estimated homography. Unfortunately, the estimated homography depends on the choice of the normalization. The proposed extension to the (linear) Taubin estimator is the perfect substitute for such methods, as it does not rely on coordinate normalization, and produces homographies whose error is consistent with existing methods. Also, unlike existing linear methods, the proposed Taubin estimator is theoretically unbiased, and unaffected by similarity transformations of the correspondences in the two views. In addition, it can be adapted to estimate other quantities such as trifocal tensors.
Article
This work extends the circle fitting method of Rangarajan and Kanatani (2009) to accommodate ellipse fitting. Our method, which we call HyperLS, relies on algebraic distance minimization with a carefully chosen scale normalization. The normalization is derived using a rigorous error analysis of least squares (LS) estimators so that statistical bias is eliminated up to second order noise terms. Numerical evidence suggests that the proposed HyperLS estimator is far superior to the standard LS and is slightly better than the Taubin estimator. Although suboptimal in comparison to maximum likelihood (ML), our HyperLS does not require iterations. Hence, it does not suffer from convergence issues due to poor initialization, which is inherent in ML estimators. In this sense, the proposed HyperLS is a perfect candidate for initializing the ML iterations.
Article
The convergence performance of typical numerical schemes for geometric fitting for computer vision applications is compared. First, the problem and the associated KCR lower bound are stated. Then, three well-known fitting algorithms are described: FNS, HEIV, and renormalization. To these, we add a special variant of Gauss–Newton iterations. For initialization of iterations, random choice, least squares, and Taubin's method are tested. Simulation is conducted for fundamental matrix computation and ellipse fitting, which reveals different characteristics of each method.