Download ALTERNATING PROJECTIONS ON NON

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Shapley–Folkman lemma wikipedia , lookup

Jordan normal form wikipedia , lookup

Singular-value decomposition wikipedia , lookup

Perron–Frobenius theorem wikipedia , lookup

Brouwer fixed-point theorem wikipedia , lookup

Cayley–Hamilton theorem wikipedia , lookup

Transcript
ALTERNATING PROJECTIONS ON NON-TANGENTIAL
MANIFOLDS.
FREDRIK ANDERSSON AND MARCUS CARLSSON
We consider sequences (Bk )∞
k=0 of points obtained by projecting
a given point B = B0 back and forth between two manifolds M1 and M2 ,
and give conditions guaranteeing that the sequence converges to a limit B∞ ∈
M1 ∩ M2 . Our motivation is the study of algorithms based on nding the
limit of such sequences, which have proven useful in a number of areas. The
intersection is typically a set with desirable properties, but for which there is
no ecient method for nding the closest point Bopt in M1 ∩ M2 . Under
appropriate conditions, we prove not only that the sequence of alternating
projections converges, but that the limit point is fairly close to Bopt , in a
manner relative to the distance kB0 − Bopt k, thereby signicantly improving
earlier results in the eld.
Abstract.
1. Introduction
R and let M be a manifold.
M = M1 ∩ M2 along with corre1
sponding projection operators denoted by π , π1 and π2 . Given a point B ∈ K,
suppose that π(B) is sought for and that π1 and π2 can be eciently computed,
whereas π cannot. A classical result by von Neumann [52] says that if M1 and M2
Let
K
be a nite dimensional Hilbert space over
Given two manifolds
M1 , M2 ⊂ K
satisfying
are ane linear manifolds, then the sequence of alternating projections
B0 = B,
(1.1)
converges to
π(B).
the angle between
Bk+1 =
π1 (Bk ),
π2 (Bk ),
if
if
k is even,
k is odd,
Moreover, the convergence rate is linear, and determined by
M1
and
M2 .
This paper is concerned with extensions of this
result to non-linear manifolds. Suppose for the moment that the limit point of a
(Bk )∞
k=0 exists and denote it by B∞ . In contrast to the case
where each Mj (j = 1, 2) is an ane linear manifold, B∞ is usually dierent from
π(B). However, given that Mj behave nicely, we may expect B∞ ≈ π(B), which
particular sequence
is one of the central results of the paper.
We begin with a brief review of related works. For a brief history of the early
developments of von Neumann's algorithm we refer to [29].
Alternating projec-
tion schemes for non-linear subsets have been used in a number of applications; cf.
2010 Mathematics Subject Classication. 41A65, 49Q99, 53B25.
Key words and phrases. Alternating projections, convergence, Non-convexity, low-rank approximation, manifolds.
1
If the manifold M is not convex, then there exist points with multiple closest points on the
manifold. To dene the projection onto M, one thus needs to involve point to set maps. We will
not use this formalism, but rather write π(B) to denote an arbitrarily chosen closest point to B
on M. This is done to simplify the presentation and because our results are stated in a local
environment where the π 's are shown to be well dened functions. See Proposition 2.3.
1
2
FREDRIK ANDERSSON AND MARCUS CARLSSON
[13, 24, 33, 36, 37, 38, 39, 43].
matrices and the sets
Mj
For instance,
K
can be the set
Mm,n
of
m × n-
be subsets with a certain structure, e.g. matrices with a
certain rank, self-adjoint matrices, Hankel or Toeplitz matrices etc. For a detailed
overview of optimization methods on matrix manifolds, cf. [1]. Alternating projection schemes between several linear subsets was investigated in [3, 29].
Other
applications concern e.g. the EM algorithm, see [8]. Much emphasis has been put
towards the use of alternating projections for the case of convex sets
M1
and
M2 ,
see for instance [5, 7, 12, 26]. It is worth pointing out that except for linear subspaces, there is no overlap between this theory and the one considered here. Dyk-
stra's alternating projection algorithm [6, 19] utilizes an additional dual variable
and a slightly dierent projection scheme that is more ecient than the classical
von Neumann method in the convex case. For connections between Dykstra's alternating projection method and the alternating direction method of multipliers
(ADMM), see [10].
However, for non-convex sets the eld remained rather undeveloped until the
90's. For example, in [13], Zangwill's Global Convergence Theorem [55] is used to
motivate the convergence of an alternating projection scheme. Zangwill's theorem
(Bk )∞
k=1 is bounded and the distance to M1 ∩ M2 is
(Bk )∞
k=1 has a convergent subsequence to a point B∞ ∈
implies that if the sequence
strictly decreasing, then
M1 ∩ M2 .
This result is an easy consequence of the fact that any sequence in a
compact set has a convergent subsequence. Thus, the use of Zangwill's theorem in
this context does not provide any information about whether the limit point of the
entire sequence exists, or if so, whether it is close to the optimal point
π(B).
To our knowledge, the rst rigorous attempt at dealing with the method of
alternating projections for non-convex sets was made in [14]. However, the setting
is rather general and the results are mainly concerned with convergence of the
sequence. In particular, when this does happen, the results in [14] does not reveal
anything about the size of the error
kπ(B) − B∞ k.
Recently, A. Lewis and J. Malick presented stronger results [35], although valid
under somewhat more restrictive conditions on
M1
and
M2 .
Before discussing
their results in more detail, we give two simple examples that illustrate some of
the diculties that may arise when using alternating projections on non-linear
manifolds.
(1, 1)
(3, 0)
(1, 0)
Figure 1.
An example demonstrating that the algorithm can get
stuck in loops where not even a subsequence converges to a point
in the intersection (the black dots). However, starting closer to the
intersection point, we do get convergence (the white dots).
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
Example 1.1.
Let
M2 = R × {0}.
K = R2
M1 = {(t, (t + 1)(3 − t)/4) : t ∈ R}
π1 ((1, 0)) = (1, 1) and π2 ((1, 1)) = (1, 0),
and set
It is easily seen that
3
and
and
hence the sequence of alternating projections does not converge, cf. Figure 1. On
the other hand, if
(1 + , 0) ∈ M2 , > 0,
is used as a starting point, the sequence
of alternating projections will converge to
(3, 0) ∈ M1 ∩ M2 .
In general, it seems reasonable to assume that if the starting point is suciently
close to an intersection point, then the sequence does converge to a point in the
intersection. The next example shows that this is not always the case.
Figure 2. Alternating projections stuck in a loop.
Example 1.2.
Without going into the details of the construction, we note that one
can construct a
C ∞ -function f
with
f (0) = 0 and horizontal segments arbitrarily
M1 = R × {0} and M2 = {(t, f (t)) : t ∈
near 0. Figure 2 explains the idea. With
R},
the sequence of alternating projections can then get stuck in projecting back
and forth between the same two points, even if the starting point is arbitrarily close
to the intersection point
However, if we have
(0, 0).
f 0 (0) 6= 0
it is hard to imagine how to make a similar
construction work. This is indeed impossible, which follows both from the present
f 0 (0) 6= 0 implies
that M1 and M2 are transversal at (0, 0). In general, given C -manifolds M1 , M2
and a point A ∈ M1 ∩ M2 , we say that A is transversal if
paper and [35]. In the terminology of the latter, the condition
1
TM1 (A) + TM2 (A) = K,
(1.2)
where
TMj (A)
denotes the tangent-space of
[35] is roughly the following:
Mj
at
A, j = 1, 2.
The main result in
Theorem 1.3. Let M1 and M2 be C 3 -manifolds and let A ∈ M1 ∩M2 be transversal. If B is close enough to A, then the sequence of alternating projections (Bk )∞
k=0
given by (1.1) converges at a linear rate to a point B∞ in M1 ∩ M2 .
(1, 1)
(0, 0)
Figure 3.
Lack of transversality due to tangential curves.
4
FREDRIK ANDERSSON AND MARCUS CARLSSON
The error
kB∞ −π(B)k is not discussed in [35], but inspection of the proofs yield
kB∞ − π(B)k ≤ 2kA − Bk.
The improvement over the previously mentioned results is thus that the entire
sequence converges and that the limit
in relative terms, i.e. comparing with
B∞ is not too far away from π(B), although
dist(B, M1 ∩ M2 ) = kB − π(B)k, it need not
be particularly close either. Moreover, the assumption of transversality is rather
restrictive.
To demonstrate the essence of the transversality assumption we now
present some examples.
Example 1.4.
A = (0, 0),
M2 = {(t, t2 ) : t ∈
R}, we obtain another example of manifolds that are not transversal at (0, 0), see
Figure 3. In this case, the sequence of alternating projections do converge to (0, 0),
since
The manifolds in Example 1.2 are not transversal at
TM1 (A) = TM2 (A) = M1 .
With
M1 = R × {0}
and
independent of the starting point, albeit extremely slowly.
Example 1.5.
M1 = R × {0}2 and M2 = {(t, t, t2 ) : t ∈ R},
transversality is not satised at A = (0, 0), but again it seems plausible that the
sequence of alternating projections converges to (0, 0). See Figure 4.
With
K = R3
and
1
M2
0.8
0.6
0.4
0.2
0
1
M1
0.5
1
0.5
0
0
−0.5
−0.5
−1
−1
Figure 4. Lack of transversality due to low dimensions.
In Example 1.5, the two manifolds clearly sit at a positive angle, but this situation
is not covered by Theorem 1.3, since the manifolds are of too low dimension to
K has dimension n and Mj has
mj , j = 1, 2, the transversality (1.2) can never be satised if m1 +m2 < n,
the sum of the dimensions of the manifolds is less than that of K. This
satisfy the transversality assumption. In fact, if
dimension
i.e. when
is, however, the case in many applications of practical interest.
The papers [34]
and [2] both consider alternating projections in a more general setting, but when
dealing with manifolds, the assumptions also imply
m1 + m2 ≥ n
(see Theorem
5.16 and condition (5.17) in [34] and Section 4.5 in [2]).
We now present the main results of the present paper. We introduce the concept
of a
non-tangential
sions of
M1
and
intersection point, which impose no restriction on the dimen-
M2 .
Loosely speaking, a point
A
in
M1 ∩ M2
is non-tangential
if the two manifolds have a positive angle in directions perpendicular to
M1 ∩ M2 .
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
5
We will show that this happens if and only if
TM1 (A) ∩ TM2 (A) = TM1 ∩M2 (A),
which we can use as denition for now. In microlocal analysis this is called
intersection
[28].
clean
We remark that non-tangential points are always transversal,
since these do satisfy the above identity by Denition 3 in [35]. The point
(0, 3)
in
Example 1.1 and the origin in Example 1.5 are both non-tangential, whereas the
origins in Examples 1.2 and 1.4 are not. We now present a simplied version of the
main result (Theorem 5.1). Figure 5 illustrates the idea.
Theorem 1.6. Let
M1 , M2 and M1 ∩ M2 be C 2 -manifolds. Given a nontangential point A ∈ M1 ∩ M2 there exists an s > 0 such that the sequence of
alternating projections (1.1) converges to a point B∞ ∈ M1 ∩ M2 , given that
kB − Ak < s. Moreover, given any > 0 one can take s such that
kB∞ − π(B)k < kB − π(B)k.
M1 ∩ M2
s
B∞ π(B)
< ǫd
A
d
B
Picture illustrating Theorem 1.6. B∞ may dier from
π(B), but the error is small compared to the distance d between
B and the manifold M1 ∩ M2 , given that B is in the ball BK (A, s).
Figure 5.
The improvement over Theorem 1.3 mainly consists of two items.
Primarily,
the assumption that the surfaces are non-tangential is not at all restrictive, and in
particular there is no implication about the dimensions of
M1
and
M2 .
For the
applications we are aware of, the set of tangential points is very small, if it exists at
all. Secondly, as has been highlighted before, we are usually interested not just in
any point of
M1 ∩ M2 ,
but the closest point
π(B).
Here the theorem says that in
M1 ∩ M2 , the error is small
M1 ∩ M2 decreases.
When M1 and M2 are linear, we have
relative terms, i.e. after dividing with the distance to
and moreover improves if the distance from
We now discuss the rate of convergence.
(1.3)
where
c = cos α
and
α
B
to
kBk − B∞ k ≤ const · ck
is the angle between the subspaces
M1
and
M2 ,
which
was shown in [26]. This is usually referred to as linear convergence. In the present
setting, we show that (1.3) holds if
satises
c > cos(α).
B
is suciently near a point whose angle
α
However, the denition of the angle between two manifolds
needs some preparation, and is discussed in Section 3.
For the precise results
6
FREDRIK ANDERSSON AND MARCUS CARLSSON
see Denition 3.1 and Theorem 5.1. The corresponding statement for transversal
manifolds has been proven in [35], and similar results can also be found in [5, 16,
18, 22, 29, 54].
In many applications, one has real algebraic varieties as opposed to manifolds, (of
which the usual complex algebraic varieties are special cases). However, algebraic
varieties are manifolds except at the singular set (singular locus). Moreover, the
singular set is a variety of smaller dimension, and hence makes up a very small part
of the original variety. Since Theorem 1.6 is a local statement, it is plausible that
it applies at non-singular intersection points of two given varieties. In Section 6,
we prove that this is indeed the case under mild assumptions. We also provide a
number of concrete results for verifying these assumptions. Finally, the paper ends
with a simple application where the alternating projection method is used for the
problem of nding correlation matrices with prescribed rank from measurements
with missing data.
The paper is organized as follows. Section 2 presents rudimentary results concerning manifolds, and Section 3 gives denitions and basic theory for determining
angles between them. The technical part of the paper begins in Section 4, where
the various projection operators are studied. With this at hand, the proof of the
main theorem (given in Section 5) is fairly straightforward. Finally, Section 6 deals
with real algebraic varieties and Section 7 presents the example. There are three
appendices with proofs of the results in Sections 2, 6 and 7 respectively.
2. Manifolds
K be
A∈K
We provide a review of necessary concepts from dierential geometry. Let
a Euclidean space, i.e. a Hilbert space of nite dimension
n ∈ N.
Given
r > 0 we write B(A, r) or BK (A, r) for the open ball centered at A with radius
K is nite-dimensional it has a unique Euclidean topology. Any subset M
of K will be given the induced topology from K. Let p ≥ 1 and let M ⊂ K be an
m-dimensional C p -manifold. We recall that around each A ∈ M there exists an
2
p
injective C -immersion
φ on an open set U in Rm such that
and
r.
Since
M ∩ BK (A, s) = Im φ ∩ BK (A, s),
(2.1)
for some s > 0, where Im φ denotes the image of φ. See e.g. Theorem 2.1.2 [9]. If
A = φ(xA ), we dene the tangent space TM (A) by TM (A) = Ran dφ(xA ), where
dφ denotes the Jacobian matrix. It is a standard fact from dierential geometry
that this denition is independent of φ, (see e.g. Section 2.5 of [9]). Moreover, we
set
T̃M (A) = A + TM (A),
T̃M (A)
i.e.,
is the ane linear manifold which is tangent to
M
at
A.
We now list
some necessary properties of manifolds. Some statements are well-known, but for
the convenience of the reader, we outline all proofs in Appendix A.
Proposition 2.1. Let
α : K → M be any C p −map, where M is a C p -manifold
and p ≥ 1. With φ as above, the map φ−1 ◦ α is also C p (on its natural domain of
denition).
2
i.e. a p times continuously dierentiable function such that dφ(x) is injective for all x ∈
BRm (0, r)
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
7
A ⊥ B to denote that A, B ∈ K are orthogonal. If X, Y ⊂ K are subsets
hA, Bi = 0 for all A ∈ X and B ∈ Y , we also write X ⊥ Y . If Y is the
orthogonal complement of X , i.e. the maximal subspace with the above property,
we shall denote it by Y = KX . When X is a subspace, we use PX : K → K for the
orthogonal projection onto X , i.e. if A ∈ X and B ∈ K X we set PX (A + B) = A.
We write
such that
Proposition 2.2. Let M be a C 1 -manifold. Then PT
of A.
The next proposition shows that projection on
is a continuous function
M (A)
M
locally
is a
well dened op-
eration. Thus, as long as we start the alternating projections method suciently
near the manifolds in question, there is no need to deal with point to set maps.
Proposition 2.3. Let
M be a C 2 -manifold and let A ∈ M be given. Then there
exists an s > 0 such that for all B ∈ BK (A, s) there exists a unique closest point
in M. Denoting this point by π(B), the map π : BK (A, s) → M is C 2 . Moreover,
C ∈ M ∩ BK (A, s) equals π(B) if and only if B − C ⊥ TM (C).
When there is risk of confusion we will write
it is easy to show that
π
is
C 1.
πM
instead of
C2
The fact that it actually is
π.
further developed in [31]. It is however not true in general that
a
C
1
We remark that
is noted in [23] and
π
is
C1
if
M
is
-manifold, in fact, it does not even have to be single valued. The last line in
the above proposition is an extension of Fermat's principle, for similar statements
in a more general environment we refer to [44].
We end with a proposition that
basically says that the ane tangent-spaces are close to
M
locally.
Proposition 2.4. Let M be a C 2 -manifold and A ∈ M be given. For each > 0
there exists s > 0 such that for all C ∈ BK (A, s) ∩ M we have
(i)
(ii)
dist(D, T̃M (C)) < kD − Ck, ∀D ∈ B(A, s) ∩ M.
dist(D, M) < kD − Ck,
∀D ∈ B(A, s) ∩ T̃M (C).
3. Non-tangentiality
Suppose now that we are given
tersection
M1 ∩ M2
is itself a
C
1
C 1 -manifolds M1
the objects from Section 2 associated to
m1 , m2
and
m
M2 such that their inM1 ∩ M2 by M, and
and
-manifold. We will denote
M1 , M2
and
M,
e.g. the dimension, by
respectively. We thus omit subindex when dealing with
We now introduce the angle between
M1
and
M2
M1 ∩ M2 .
A ∈ M.
at an intersection point
This issue is a bit delicate. We rst discuss the case when
M1
and
M2
are linear
subspaces. A good reference for various denitions of angles in this setting is [17].
We will here use the so called Friedrich's angle, which is common in functional analysis [3, 42] and also the one appearing in (1.3). The Friedrich's angle
two subspaces is then given by
(3.1)
α = cos−1 (kPM1 PM2 − PM1 ∩M2 k) .
To better understand this denition, set
Fj = {Bj ∈ Mj (M1 ∩ M2 )} \ {0}
and note that
(3.2)
−1
α = cos
sup
Bj ∈Fj
hB1 , B2 i
kB1 kkB2 k
!
,
α
between
8
FREDRIK ANDERSSON AND MARCUS CARLSSON
(assuming F1 6= ∅ =
6 F2 , in which case it becomes pointless to talk of angles). Hence,
α is the minimal angle in directions perpendicular to the common intersection M1 ∩
M2 .
To introduce angles between manifolds one obviously needs to let the angle be
A ∈ M1 ∩ M2 . Based on
TM1 (A) and TM2 (A), i.e.
dependent on the point of intersection
to let the angle at
A
be the angle of
(3.1), one idea is
α(A) = cos−1 kPTM1 (A) PTM2 (A) − PTM1 (A)∩TM2 (A) k .
(3.3)
This is indeed the denition adapted in [35], see Section 3.2.
However, consider
again the two manifolds in Example 1.4 and Figure 3. The expression (3.3) then
leads to the counterintuitive conclusion that
α((0, 0)) = cos−1 (0) = π/2.
In fact,
it is easy to see from (3.2) that the angle between two subspaces never can be
0,
whereas we argue that the angle should be 0 for the manifolds in Figure 3. For this
reason, we adopt the following denition, generalizing (3.2).
Denition 3.1. Given A ∈ M1 ∩ M2 , set
Fjr = {Bj ∈ Mj \ A, kBj − Ak < r and Bj − A ⊥ TM1 ∩M2 (A)}.
If Fjr 6= ∅ for all r > 0 and j = 1, 2, we dene the angle α(A) of M1 and M2 at A
as
α(A) = cos−1 (σ(A))
where
σ(A) = lim sup
r→0 Bj ∈F r
j
hB1 − A, B2 − Ai
kB1 − AkkB2 − Ak
Otherwise, we let both σ(A) and α(A) be undened.
.
This means that that the angle, if it is dened, is minimized locally in directions
perpendicular to
TM1 ∩M2 (A),
(as opposed to
TM1 (A) ∩ TM2 (A)).
With Denition
3.1, angles can be 0, and it is readily veried that the angle is zero in Example
1.4. The next proposition shows that the angle is undened only at points where
it makes no sense to talk of angles.
Proposition 3.2. The angle of M1 and M2 at A is undened is and only if one
of the manifolds is a subset of the other in a neighborhood of A, that is, if there
exists an s > 0 such that M1 ∩ BK (A, s) ⊂ M2 ∩ BK (A, s) or M2 ∩ BK (A, s) ⊂
M1 ∩ BK (A, s).
Proof.
There is clearly no restriction to assume that
M
Recall that a
C 1-
1
C -function
(TM (0))⊥ , (this is a simple
improvement of Theorem 2.1.2 (iv) in [9]). Now, either TM1 ∩M2 (A) = TM1 (A)
or not. The fact just mentioned implies that in the rst case M1 ∩ M2 and M1
r
coincide near 0, and in the second case F1 6= ∅ for all r > 0. The same dichotomy
obviously holds with the roles of 1 and 2 switched, and the proposition follows.
manifold
containing
0
A = 0.
locally can be written as the graph of a
dened in a neighborhood of 0 in
TM (0)
with values in
Denition 3.3. Points A ∈ M1 ∩M2 where the angle is dened will be called nontrivial intersection points. For such points, we say that A is tangential if α(A) = 0
and non-tangential if α(A) > 0.
Proposition 3.4.
σ is continuous on supp σ and
σ(A) = kPTM1 (A) PTM2 (A) − PTM1 ∩M2 (A) k.
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
Proof.
(3.4)
9
A ∈ M1 ∩ M2 be non-trivial. It is easy to see that
hB1 − A, B2 − Ai
σ(A) = sup
: Bj ∈ TMj (A) TM1 ∩M2 (A) =
kB1 − AkkB2 − Ak
= kPTM1 (A)TM1 ∩M2 (A) PTM2 (A)TM1 ∩M2 (A) k =
Let
= kPTM1 (A) PTM2 (A) − PTM1 ∩M2 (A) k.
The continuity of
σ
now follows by Proposition 2.2.
In particular, we get
α(A) = cos−1 kPTM1 (A) PTM2 (A) − PTM1 ∩M2 (A) k
(3.5)
which should be compared with (3.3). Moreover, we infer that non-tangentiality is a
stable property, i.e. if
in a neighborhood of
A is non-tangential, then the same holds for all B ∈ M1 ∩M2
A. Our experience is that in practice, tangential and trivial
intersection points are exceptional.
For example, we will show in Section 6 that
M1 and M2 are also algebraic varieties, then non-tangentiality at
in M1 ∩ M2 implies non-tangentiality of the vast majority of points.
if
one point
The next
proposition gives a useful criterion for checking non-tangentiality.
Proposition 3.5. Let A ∈ M1 ∩ M2 be a non-trivial intersection point. Then A
is non-tangential if and only if
TM1 (A) ∩ TM2 (A) = TM1 ∩M2 (A).
(3.6)
Proof.
By (3.4) it is easy to see that
A
is non-tangential if and only if
TM1 (A) TM1 ∩M2 (A) ∩ TM2 (A) TM1 ∩M2 (A) = {0}
which in turn happens if and only if (3.6) holds.
In particular, for non-trivial intersection points a simple criterion which implies
non-tangentiality is
dim(TM1 (A) ∩ TM2 (A)) ≤ m,
m = dim(M1 ∩ M2 )), which is immediate by
TM1 (A) ∩ TM2 (A) ⊃ TM1 ∩M2 (A). (3.6) is usually referred
(recall that
(3.6) and the inclusion
to as
clean intersection
in microlocal analysis, (see Appendix C.3, [28]). We have chosen the terminology
non-tangential since it is more intuitive. Note that by (3.5) and (3.6), the angle
as dened in Denition 3.1 and that of (3.3) are equivalent whenever
A
is non-
tangential. Finally, Proposition 3.5 shows that non-tangentiality is a weaker concept
than transversality, dened in (1.2). For if two manifolds are transversal at
A, then
(3.6) holds by Denition 3 in [35].
4. Properties of the projection operators
Let
M1 , M2
and
M = M1 ∩ M2
be
C 2 -manifolds
and
A ∈ M1 ∩ M2
a
non-tangential intersection point. This will be xed throughout the section. We
M1 ∩ M2 without
πM1 , πM2 and πM by π1 , π2
continue to use the convention of denoting objects related to
subindex. As in the introduction we shall abbreviate
and
π.
The following result will be the main tool for proving convergence of the
alternating projections.
10
FREDRIK ANDERSSON AND MARCUS CARLSSON
Theorem 4.1. For each ε > 0 there exists an s > 0 such that for all B ∈ B(A, s)
we have
kπ(πj (B)) − π(B)k < εkB − π(B)k,
j = 1, 2.
The proof is rather technical and depends on a long series of lemmas.
number
M.
s
and the map
φ
appearing in (2.1) clearly dier between
M1 , M2
The
and
φ1 , φ2 and φ, and let s−1 be a number
such that (2.1) holds in all three cases with s = s−1 . We will make repeated use of
Propositions 2.3 and 2.4, applied to each of the manifolds M1 , M2 and M1 ∩ M2 .
The numbers s appearing in these also depend on which manifold that is considered,
and also on the auxiliary constant . We will choose a xed value for in the proof
of Theorem 4.1. In the meantime, it will be treated as a xed number. We let s0
We denote the respective maps by
be a constant small enough that Proposition 2.3 applies to each of the manifolds
M1 , M2 and M1 ∩ M2
B(A, s0 ) and have A as a
in
B(A, s0 ).
Since
π1 , π2
π
and
then are continuous in
xed point, we may also assume that
that their respective images are contained in
B(A, s−1 ).
s0
is small enough
s0 ≤ s0
B(A, s0 ).
Finally, we let
such that also Proposition 2.4 applies for each of the 3 manifolds in
be
s4
B
E
F
π(B)
π1 (B)
D
ρ1 (B)
M1
M1 ∩ M2
Figure 6.
and
F
Illustration of the dierence between
We introduce maps
ρj : B(A, s0 ) → K,
(
ρj (B) = πT̃M
j
Thus,
ρj
resemble
πj
π(B),
j=1
and
Mj
or
j = 2),
π1 . D, E
via
(π(B)) (B).
but is slightly dierent.
projects onto the tangent plane of
i.e.
ρ1
appear in the proof of Proposition 4.3.
πj
Mj whereas ρj
B in M1 ∩ M2 ,
ρj and πj is given in
projects onto
taken at the closest point to
(see Fig 6). An estimate of the dierence between
Proposition 4.3.
Lemma 4.2. The functions ρ1 and ρ2 are C 1 -maps in B(A, s0 ). Moreover, we can
select a number s1 < s0 such that the image of B(A, s1 ) under ρ1 , ρ2 , π, π1 , π2 , as
well as any composition of two of those maps, is contained in B(A, s0 ).
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
Proof.
11
The second part is an immediate consequence of the continuity of the maps,
π, π1 and π2 are continuous by Proposition 2.3. Let j denote 1 or 2, and
−1
set Mj (B) = dφj (φj (π(B))), which is well dened in B(A, s0 ) by the choice of
2
s−1 and s0 . By Proposition 2.1 and 2.3, we have that φ−1
j (π(B)) is a C -map in
1
2
B(A, s0 ), and hence Mj is C as φj is C . Given any point B0 ∈ Mj it is easy to
and
see that
πT̃M
j
(B0 ) (B)
= B0 + PTMj (B0 ) (B − B0 )
and hence, by the formula already used in the proof of Proposition 2.2, we have
= π(B) + PTMj (π(B)) (B − π(B)) =
−1 ∗
= π(B) + Mj (B) Mj∗ (B)Mj (B)
Mj (B) B − π(B)
ρj (B) = πT̃M
j
(π(B)) (B)
from which the result follows.
B
r2
r
EB
r1
x
h
DB
M1
Figure 7.
kDB k)
and
Proof illustration.
r2 = kEB k − h.
Proposition 4.3. Suppose that
j = 2, we have
Proof.
√1+
1−2
r = kEB k+kDB k, h = 2(kEB k+
< 2. Given any B ∈ B(A, s1 ) and j = 1 or
√
kπj (B) − ρj (B)k < 4 kB − π(B)k.
C = π(B). By
π(B) = 0, which we do from now on. Denote
D = TMj (0), E = Span {B − ρj (B)} and F = K (D + E), (see Fig. 6). Let DB
and EB be elements of D and E such that B = DB + EB , and note that
Lemma 4.2 implies that Proposition 2.4 applies to the point
a translation we can assume that
ρj (B) = DB .
We thus have to show that
(4.1)
√
kπj (B) − DB k < 4 kBk.
First note that by Proposition 2.4 (ii) (applied with
the set
Mj ∩ B(DB , kDB k) is not void.
C = π(B) = 0
and
D = DB )
This set is included in the ball with center
12
FREDRIK ANDERSSON AND MARCUS CARLSSON
B
and radius
to
B
Mj ,
on
kB − DB k + kDB k = kEB k + kDB k.
Since
πj (B) is the closest point
we conclude that
kB − πj (B)k ≤ kEB k + kDB k.
If we write
πj (B) = D + E + F
(with
D, E, F
in the respective subspace
this becomes
D, E, F )
kD − DB k2 + kE − EB k2 + kF k2 < (kEB k + kDB k)2 .
(4.2)
However, by Proposition 2.4 (i) (applied with
also get
C = π(B) = 0
and
D = πj (B))
we
kEk2 + kF k2 < 2 kDk2 + kEk2 + kF 2 k .
(4.3)
The left hand side of (4.1) is thus dominated by the supremum of the function
δ(D, E, F ) = kD + E + F − DB k =
p
kD − DB k2 + kEk2 + kF k2 ,
subject to the conditions in (4.2) and (4.3). Either by geometrical considerations
or the method of Lagrange multipliers, it is not hard to deduce that this supremum
is attained for
F =0
and
D, E
of the form
D = dDB
and
E = eEB
where
d, e ∈ R.
We now have a two-dimensional problem of circles and cones, and in the remainder
D,pDB , E and EB as elements of R2 . Summing up,
δ(D, E) = kD − DB k2 + kEk2 for D, E ∈ R2 satisfying
of the proof we treat
to maximize
we want
kD − DB k2 + kE − EB k2 < (kEB k + kDB k)2
(4.4)
and
kEk < √
(4.5)
The rst inequality gives
kDk ≤ kD − DB k + kDB k < kEB k + kDB k + kDB k which
inserted in the second yields
kEk ≤ √
1+
As √
2
1−
< 2,
kDk.
1 − 2
(1 + )kDB k + kEB k .
1 − 2
the following condition
kEk < 2 kDB k + kEB k
(4.6)
is less restrictive than (4.5) (when combined with (4.4), see Fig. 7). The function
δ
to be maximized is just the distance from
distance
d
D+E
to
DB ,
and clearly the maximal
(with constraints (4.4) and (4.6)) is obtained by the point
x
in the
picture. Using the notation from the picture, we have
q
q
d = kDB − xk = r12 + h2 = r2 − r22 + h2 =
p
= (kEB k + kDB k)2 − (kEB k − 2(kEB k + kDB k))2 + (2(kEB k + kDB k))2 =
p
= 2kEB kkDB k + 2 kDB k2 + 4(kEB k2 + kEB kkDB k).
clearly implies < 1 so 2 < p
. Moreover, kEB k, kDB k ≤ kBk
√
and hence the value obtained above is less than
11kBk2 ≤ 4 kBk. Since the
value in question is also larger than kπj (B) − ρj (B)k the proposition follows.
The assumption on
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
13
Lemma 4.4. Given B ∈ B(A, s1 ) we have
π(B) = π(ρ1 (B)) = π(ρ2 (ρ1 (B))).
Proof.
We begin with the rst equality. By Lemma 4.2 we have π(B), ρ1 (B) ∈
B(A, s0 ), since s0 < s0 , so Proposition 2.3 applies to give B−π(B) ⊥ TM1 ∩M2 (π(B)).
We obviously also have B − ρ1 (B) ⊥ TM1 (π(B)) so
ρ1 (B) − π(B) = (ρ1 (B) − B) + (B − π(B)) ⊥ TM1 ∩M2 (π(B)).
By Proposition 2.3 this implies that
π(B) = π(ρ1 (B)), as desired. By Lemma 4.2,
ρ2 (B) in place of B , which yields
the above argument also applies to the point
π(ρ2 (B)) = π(ρ1 (ρ2 (B))).
The desired second identity follows by reversing the
roles of 1 and 2.
We nally have all the necessary ingredients.
Proof of Theorem 4.1.
B(A, s0 ),
By the choice of
and hence we can pick
C>0
s0
and Proposition 2.3,
π
is
C2
on
such that
kπ(B) − π(B 0 )k ≤ CkB − B 0 k
√
1+
0
< 2, and set
for all B, B ∈ B(A, s0 ). Now x such that C4 < ε and √
1−2
s = s1 . By Lemma 4.4 we have π(B) = π(ρj (B)), and Lemma 4.2 guarantees that
πj (B), ρj (B) ∈ B(A, s0 ). By (4.7) and Proposition 4.3 we get
(4.7)
kπ(πj (B)) − π(B)k = kπ(πj (B)) − π(ρj (B))k ≤ Ckπj (B) − ρj (B)k ≤
√
C4 kB − π(B)k < εkB − π(B)k,
as desired.
M1 ∩ M2 is reduced in proportion to the
M1 onto M2 or vice versa. Recall the function
Next we show that the distance to
angle, each time we project from
σ(A)
from Denition 3.1.
Theorem 4.5. For each
M2 ∩ B(A, s) we have
c > σ(A) there exists an s > 0 such that for all B ∈
kπ1 (B) − π(B)k < ckB − π(B)k.
Moreover the same holds true with the roles of M1 and M2 reversed.
Lemma 4.6. Let c̃ > 1 and ˜ > 0 be given. If E, F
kE − F k < ˜, then
kEk < ˜
Proof.
If
kEk < ˜ we
∈ K satises kEk > c̃kF k and
c̃
.
c̃ − 1
are done, otherwise
c̃ <
kEk
kEk
<
kF k
kEk − ˜
which easily gives the desired estimate.
Proof of Theorem 4.5.
Fix
c1
such that
σ(A) < c1 < c and pick an s1 < s0
that
sup{σ(C) : C ∈ M1 ∩ M2 ∩ B(A, s1 )} < c1 ,
such
14
FREDRIK ANDERSSON AND MARCUS CARLSSON
σ is continuous by Proposition 3.4. Let c2 > 1 be such
c2 c1 < c. By the choice of s0 and Lemma 4.2, ρ1 and π are C 1 -functions on
B(A, s0 ), and hence we can pick C > 0 such that
which we can do since
that
kρj (B) − ρj (B 0 )k ≤ CkB − B 0 k
(4.8)
for all
(4.9)
Then
kπ(B) − π(B 0 )k ≤ CkB − B 0 k
and
B, B 0 ∈ B(A, s0 ). We now x such that √1+
< 2 and
1−2
√
√
c2
< c and (1 + 4 )c2 c1 < c.
4 (1 + C)
c2 − 1
x s < s1 such that
and let
B ∈ M2 ∩ B(A, s).
π(B(A, s)) ⊂ B(A, s1 ),
There is no restriction to assume that
we do from now on. Note that
π(B) ∈ M1 ∩ M2 ∩ B(A, s1 )
π(B) = 0,
which
so
σ(0) = σ(π(B)) < c1 .
(4.10)
Setting
D = π1 (B),
the desired inequality takes the form
TM1 (0)
A
TM2 (0)
M2
M1
B′
M1 ∩ M2
D′
B
D
Figure 8. Proof illustration
(4.11)
0
0
0
kDk/kBk < c.
B = ρ2 (B) and D = ρ1 (B ), (see Figure 8). Note that B ∈ B(A, s1 ) whereas
D, B 0 , D0 ∈ B(A, s0 ) by Lemma 4.2. By Lemma 4.4 it then follows that 0 =
π(ρ2 (B)) = π(B 0 ) and 0 = π(ρ1 (ρ2 (B))) = π(D0 ) so
Put
D0 = PTM1 (0) (B 0 ) = PTM1 (0) PTM2 (0) (B 0 ),
since
B 0 ∈ TM2 (0).
B 0 ⊥ TM1 ∩M2 (0)
D0 = PTM1 (0) PTM2 (0) − PTM1 ∩M2 (0) B 0 .
By Proposition 2.3 we also have
By Proposition 3.4 and (4.10) we thus have
(4.12)
kD0 k
≤ kPTM1 (0) PTM2 (0) − PTM1 ∩M2 (0) k = σ(0) < c1 ,
kB 0 k
and hence
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
which should be compared with (4.11). We now show that
First, note that by Proposition 4.3
B0 ≈ B
and
15
D0 ≈ D.
√
kB − B 0 k = kπ2 (B) − ρ2 (B)k < 4 kBk.
(4.13)
Moreover, by (4.8), (4.13) and Proposition 4.3 we have that
kD0 − Dk = kρ1 (B 0 ) − π1 (B)k ≤ kρ1 (B 0 ) − ρ1 (B)k + kρ1 (B) − π1 (B)k ≤
√
√
≤ CkB 0 − Bk + 4 kBk < 4 (1 + C)kBk
√
0
Apply Lemma 4.6 with E = D , F = D , ˜ = 4 (1 + C)kBk and c̃ = c2 . We
see
that either
kDk
≤ c2
kD0 k
(4.14)
√
2
kBk, in which case we are done since
kDk < 4 (1 + C) c2c−1
than c by (4.9). We thus assume that (4.14) holds. Note that
√
kB 0 k
(4.15)
≤1+4 kBk
or
the constant is less
by (4.13). Combining (4.12), (4.14) and (4.15) we get
√
kDk
kDk kB 0 k kD0 k
=
< c2 (1 + 4 )c1 < c,
0
0
kBk
kD k kBk kB k
where the last inequality follows by (4.9).
5. Alternating projections
We are nally ready for the main theorem.
Theorem 5.1. Let M1 , M2 and M1 ∩ M2 be C 2 -manifolds and let A ∈ M1 ∩ M2
be a non-tangential intersection point of M1 and M2 . Given > 0 and 1 > c >
σ(A), there exists an r > 0 such that for any B ∈ B(A, r) the sequence of alternating
projections
B0 = π1 (B), B1 = π2 (B0 ), B2 = π1 (B1 ), B3 = π2 (B2 ), . . .
(i)
(ii)
(iii)
Proof.
converges to a point B∞ ∈ M1 ∩ M2
kB∞ − π(B)k ≤ kB − π(B)k
kB∞ − Bk k ≤ const · ck kB − π(B)k
We may clearly assume that
(5.1)
and Theorem 4.5 with
c
< 1. Invoking
1−c
,
ε=
3−c
Theorem 4.1 with
as above, gives us two possibly distinct radii.
denote the minimum of the two and pick
r<
(5.2)
such that
(5.3)
π(B(A, r)) ⊂ B(A, s/4).
s(1 − )
4(2 + )
The latter condition ensures that
kπ(B) − Ak < s/4.
We let
s
16
Let
FREDRIK ANDERSSON AND MARCUS CARLSSON
l = kB − π(B)k
and observe that
(5.4)
As
π(B) ∈ M1 ∩ M2
l ≤ kB − Ak + kA − π(B)k ≤ r + s/4.
we have
kB0 − Bk = kπ1 (B) − Bk ≤ kπ(B) − Bk = l
and
kB0 − π(B0 )k ≤ kB0 − π(B)k ≤ kB0 − Bk + kB − π(B)k ≤ 2l.
(5.5)
Applying Theorem 4.5 we get
kBk − π(Bk )k ≤ kBk − π(Bk−1 )k ≤ ckBk−1 − π(Bk−1 )k
as long as
{Bk }k−1
k=0 ⊂ B(A, s),
(5.6)
which will be established by induction. First note that
kB0 − Ak ≤ kB0 − Bk + kB − Ak ≤ l + r ≤ 2r + s/4 ≤
by (5.2), (5.4) and (5.5), and hence (5.6) holds for
for a xed
k ≥ 1.
k = 1.
s(1 − ) s
+ <s
2(2 + ) 4
Now assume that it holds
We then get
(5.7)
kBk − π(Bk )k ≤ ck kB0 − π(B0 )k ≤ 2lck
and by Theorem 4.1 we also have
kπ(Bk ) − π(Bk−1 )k ≤ εkBk−1 − π(Bk−1 )k ≤ ε(2lck−1 ).
(5.8)
By the triangle inequality and Theorem 4.1 we get
kπ(Bk ) − π(B)k ≤ kπ(B) − π(B0 )k +
(5.9)
εl +
k
X
ε(2lck−1 ) < εl +
j=1
k
X
j=1
kπ(Bj ) − π(Bj−1 )k ≤
3−c
2εl
=
εl = l,
1−c
1−c
where the last inequality follows by (5.1). Combining this with (5.3), (5.4), (5.7)
we also have
kA − Bk k ≤ kA − π(B)k + kπ(B) − π(Bk )k + kπ(Bk ) − Bk k < s/4 + l + 2l ≤
3+
3+
1−
≤ s/4 + (r + s/4) + 2(r + s/4) =
s + (2 + )r <
s+
s = s,
4
4
4
where the last inequality follows by (5.2). This shows that (5.6) holds for k := k + 1
and so (5.6) is true for all k ∈ N by induction. In particular all the above estimates
hold true for any k ∈ N.
∞
By (5.8) we see that the sequence (π(Bk ))k=1 is a Cauchy sequence, and hence
∞
it converges to some point B∞ . By (5.7) the sequence (Bk )k=1 must also converge,
and the limit point is again B∞ , which thus satises B∞ = π(B∞ ) since π is
continuous. Hence (i) is established, and (ii) follows by taking the limit in (5.9).
For (iii), note that by a similar calculation to (5.9) we have
∞
X
ε2lck
,
kπ(Bk ) − B∞ k ≤
kπ(Bj ) − π(Bj−1 )k ≤
1−c
j=k+1
which combined with (5.7) gives
kBk − B∞ k ≤ kBk − π(Bk )k + kπ(Bk ) − B∞ k ≤ 2lck +
ε2lck
2 − 2c + 2ε k
=
c l,
1−c
1−c
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
17
as desired.
6. Non-tangentiality on real algebraic varieties
In this section we suppose that
with
Rn .
K
has been given a basis and thus identify it
In many applications the sets
M1 , M2
and
M are actually real algebraic
varieties, that is, they can be dened by the vanishing of a set of polynomials on
Rn .
To make the distinction explicit we denote them
V1 , V2
and
V,
respectively.
This poses a few additional problems, but we shall see in the end of this section
that these can be overcome and moreover, that non-tangentiality at one single
intersection point implies non-tangentiality at the vast majority of points in
V,
in
a sense to be specied below. This result is interesting for applications since it says
that the (a priori not known) closest point(s) to a given
B ∈ Rn
is likely to be
non-tangential, so that Theorem 5.1 applies, given that we are close enough.
The rst obstacle is that algebraic varieties have singular points, and are therefore not manifolds. This is true even for complex algebraic varieties, which of course
can be regarded as real by separating the real and imaginary parts. Fortunately,
we have the following theorem by H. Whitney [53].
Theorem 6.1. Given a real algebraic variety V ⊂ Rn , we can write V = ∪m
j=0 Mj
where each Mj is either void or a C ∞ -manifold of dimension j . Moreover, each
Mj , j = 0, . . . , m contains at most a nite number of connected components.
For example, consider the Cartan umbrella
V
given by
z(x2 + y 2 ) − x3 = 0
in
R3 .
The main part of the variety is a manifold of dimension 2, but it also contains
the
z−axis
as a submanifold of complex dimension 1. Thus, in terms of Theorem
6.1, we have
V = M1 ∪ M2
M2 = {(x, y, z) : (x, y) 6= 0
(6.1)
and
where
M1 = {0} × {0} × R.
and
z = x3 /(x2 + y 2 )}
For more interesting examples in this direction we refer
to [4], Chapter 3.
Theorem 6.1 shows that the main part of
V
is a manifold, and we can apply
the theory from the previous sections since the statements are local.
that
∪m−1
j=0 Mj
V . We will
m−1
m = 1, ∪j=0
Mj
makes up a very small part of
statement, but note that in the case
nite collection of points to be compared with the curve
is obtained by taking
V = Rn .
We note
not try to quantify this
would correspond to a
M1 .
The information we have about
Another example
∪m−1
j=0 Mj
is then
m−1
much stronger than e.g. saying that ∪j=0 Mj has Lebesgue measure zero, which
is a common criterion for negligible sets, see for example Sard's theorem [45].
Our next aim is to nd conditions on two real algebraic varieties
V1
and
V2
which guarantee that Theorem 5.1 can be applied to the majority of points in
V = V1 ∩ V2 ,
in the sense that the exceptional set is a union of manifolds of
lower dimension with nitely many connected components. To do this, we need to
connect Theorem 6.1 with classical algebraic geometry in
Cn .
This is also done in
the seminal paper [53], which has given rise to a rich theory on decomposition of
various sets into manifolds, now known as (Whitney) stratication theory. For an
overview and historical account we refer to [50]. However, although Propositions
6.2-6.4 below are implicitly shown in Sections 8-11 in [53], we have not been able
to nd a good reference for all the results we need here. Hence, for completeness
18
FREDRIK ANDERSSON AND MARCUS CARLSSON
we provide proofs in Appendix B. The material in this section is self-contained and
it is not necessary to know algebraic geometry in order to apply the results below.
Usually, by an algebraic variety one refers to a subset of
Cn
dened as the
common zeroes of a set of (analytic) polynomials (with complex coecients). To
distinguish between these varieties and real algebraic varieties, we will call them
complex algebraic varieties or simply complex varieties.
Note that all complex
varieties are real varieties, but not conversely. However, if we identify
subset of
C
n
in the usual way, then each real variety
VZar , dened
vanish on V . Given
V
Cn
as the subset in
all polynomials that
a real algebraic manifold
V.
with a
has a related complex variety
given by its Zariski closure
denote the set of real polynomials that vanish on
Rn
of common zeroes to
V
IR (V)
we let
It is easy to see that
VZar = {z ∈ Cn : p(z) = 0, ∀p ∈ IR (V)}.
(6.2)
There are several equivalent ways of dening the dimension of a complex algebraic
variety, see e.g. [15, 47]. The below proposition basically states that these coincide
with the maximal non-trivial dimension in any Whitney composition.
Proposition 6.2. Let
V be a real algebraic variety and let V = ∪m
j=0 Mj be a
decomposition as in Theorem 6.1, with Mm 6= ∅. Then m equals the algebraic
dimension of VZar .
Hence, although Whitney decompositions are not unique (consider e.g.
(R \ {0}) ∪ {0}),
the maximal dimension of its components is. The number
V.
henceforth be called the dimension of
R =
m will
For example, it can be veried that the
Zariski closure of the Cartan umbrella equals the variety in
C3
dened by the same
equation, and that this variety has algebraic dimension 2, as predicted by the above
proposition and (6.1).
V is irreducible if there does not exist any non-trivial decompositions
V = V1 ∪V2 , where V1 and V2 are algebraic varieties. A simple condition
We say that
of the form
which guarantees this (and involves no algebraic geometry) is given in Proposition
6.8. We say that a point
A∈V
is non-singular if it is non-singular in the sense
of algebraic geometry as an element of
VZar .
The theory becomes much simpler if
we restrict attention to irreducible varieties, which we do from now on. Let
denote the gradient operator and set
this is a linear subspace of
Rn .
NV (A) = {∇p(A) : p ∈ IR (V)}.
∇
to
Note that
Proposition 6.3. Let V ⊂ Rn be an irreducible real algebraic variety of dimension
m. Then dim NV (A) ≤ n − m for all A ∈ V and A is non-singular if and only if
The subset of
V
dim NV (A) = n − m.
of non-singular points will be denoted
Proposition 6.4. Let
V ns .
V be an irreducible real algebraic variety of dimension m.
Then the decomposition V = ∪m
j=0 Mj in Theorem 6.1 can be chosen such that
V ns = Mm . Moreover, given A ∈ V ns we have
(6.3)
TV ns (A) = (NV (A))⊥ .
Note that the counterpart of (6.3) in the complex setting also holds, (this will
follow in Appendix B). We remark that in this case, the right hand side is the
denition of the tangent space of
V
in algebraic geometry, so (6.3) implies that the
algebraic tangent space coincides with the dierential geometry tangent space
at non-singular points.
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
19
We now return to the issue of alternating projections. If we have irreducible varieties
V1 , V2 , V
of dimensions
m1 , m2 , m and restrict attention to V1ns , V2ns , V ns ,
it makes sense to talk about non-tangential intersection points and Theorem 5.1
can be applied. We suppose that one of the varieties is not a subset of the other.
Then all intersection-points are non-trivial, i.e.
V1ns
the angle between
exists.
V2ns
and
Proposition 6.5. Suppose that
V1 and V2 are irreducible real algebraic varieties
and that V = V1 ∩ V2 is irreducible and strictly smaller than both V1 and V2 . Then
each point in V1ns ∩ V2ns ∩ V ns is a non-trivial intersection point.
The next theorem shows that in this setting, the majority of points are nontangential if one is. It also gives a simple criterion under which this is the case.
Denote by
V ns,nt ⊂ V
the set of all points in
ns
with respect to the manifolds V1 and
V2ns .
V ns ∩V1ns ∩V2ns
that are non-tangential
Theorem 6.6. Suppose that V1 and V2 are irreducible real algebraic varieties and
that V = V1 ∩ V2 is irreducible and strictly smaller than both V1 and V2 . Let the
dimension of V be m. If V ns,nt 6= ∅, then V \ V ns,nt is a real algebraic variety of
dimension strictly less than m. A sucient condition for this to happen is that
there exists a point A ∈ V1ns ∩ V2ns such that
dim (TV1ns (A) ∩ TV2ns (A)) ≤ m.
(6.4)
Combined with Theorem 6.1 and Proposition 6.2, the theorem implies that
V1ns,nt
C ∞ -manifolds
is a union of nitely many disjoint connected
strictly less than
of points in
V.
m.
V\
of dimension
In other words, the good points make up the vast majority
Note that (6.4) is equivalent with
dim (TV1ns (A) + TV2ns (A)) ≥ m1 + m2 − m
by basic linear algebra.
Theorem 6.6 thus provides a concrete condition under which Theorem 5.1 applies to a majority of points in
preconditions of Theorem 6.6.
V.
We now provide concrete means to verify the
Denition 6.7. Suppose we have given a
j ∈ N and an index set I such that
for each i ∈ I , there exists an open connected Ωi ⊂ Rj and a real analytic map
φi : Ωi → V . We will say that V can be covered with analytic patches if, for each
A ∈ V , there exists an i ∈ I and a radius rA such that
V ∩ BRn (A, rA ) = Im φi ∩ BRn (A, rA ).
Note that we put no restriction on the rank of
dφi
at any point, so that
Rj
may
contain redundant variables and degenerate points.
Proposition 6.8. Let V be a real algebraic variety. If V is connected and can be
covered with analytic patches, then V is irreducible. The same conclusion holds if
a dense subset of V can be given as the image of one real analytic function φ.
To apply Theorem 6.6, one needs to know the various dimensions involved. This
can sometimes be tricky. For example, the curve
y 2 = x2 (x−1) in R2
has dimension
1 but an isolated point at the origin. Thus a local analysis of the surface may lead to
false conclusions concerning the dimension. A more complex example is provided by
the Cartan umbrella discussed earlier. We end this section with a simple method
20
FREDRIK ANDERSSON AND MARCUS CARLSSON
of determining the dimension. An alternative is to work with quotient manifolds,
as explained in [1], but we will not pursue this here.
Proposition 6.9. Under either of the assumptions of Proposition 6.8, suppose
in addition that an open subset of V is the image of a bijective real analytic map
dened on a subset of Rd . Then V has dimension d.
A nal remark. Suppose that
case one may of course identify
K
C
to begin with is a vector space over
with
R2
C.
In this
and apply the results of this section.
However, all subspaces considered above then become closed over
C
and all results
have a natural complex counterpart where e.g. dimension refers to the dimension
over
C.
We leave these reformulations as well as their proofs to the reader.
7. An example
We provide a simple example in which Theorem 5.1 applies; Given an
symmetric matrix
lies closest to
choose
K
A
B
and an integer
k,
we seek the rank
matrices with 1's on the diagonal, and
k,
n×n
correlation matrix that
with respect to the Hilbert-Schmidt norm. For this problem, we
n × n-matrices, V1
to be the set of real symmetric
or equal to
k
where
k < n.
V2
the ane subspace of
the subset of matrices with rank less than
This example is also considered in [11] as well as in
[2], where the reformulation suggested in [25] is used so that it ts the framework
of [2].
For the alternating projections method to be useful, we need fast implementations of the projections
π1
and
π2 .
Since
V1 is
π2 ,
to compute and well-dened everywhere. For
an ane subspace,
π1
is trivial
a theorem essentially due to E.
Schmidt [46], sometimes attributed to Eckart-Young [20], states that the closest
rank
k
B is obtained by computing the singular
n − k smallest singular values by 0. This
Example 5. Note that when the k th sin-
(or less) approximation for a given
value decomposition and then replace the
is well known and explained e.g. in [35],
gular value has multiplicity higher than 1, the statement still applies but the best
approximation is no longer unique. This is the reason why
π2
needs to be a point
to set map when dened globally. However recall Proposition 2.3 and Proposition
6.4 which imply that
π2
V2 .
V = V1 ∩ V2 .
is a well dened function near non-singular points of
We show that Theorem 5.1 applies to the majority of points in
First we need to obtain basic results concerning the structure of these sets.
Proposition 7.1.
K has dimension
Proposition 7.2.
V2ns equals all matrices with rank equal to k and
n2 +n
2
. V1 is an ane subspace of dimension
2
+k
. V2 is an irreducible real algebraic variety of dimension 2nk−k
. V is an
2
2nk−k2 +k
irreducible real algebraic variety of dimension
−
n
.
2
n2 −n
2
V ns,nt 6= ∅.
Combining the above results with Proposition 6.5 and Theorem 6.6 we immediately obtain.
Corollary 7.3. The set
2nk−k2 +k
2
V ns,nt of points on V where Theorem 5.1 applies is an
− n dimensional manifold. Its complement V \ V ns,nt is a nite set of
connected manifolds of lower dimension.
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
21
−1.2
−1.4
−1.6
kB∞ −π(B)k
kkB−π(B)k
−1.8
−2
log10
−2.2
−2.4
−2.6
−2.8
−3
−3.2
−1.5
−1
−0.5
0
0.5
1
log10 (kB − π(B)k)
Figure 9.
Simulations using alternating projections for nding
the closest correlation matrix given a data set with missing data
elements. The relation between the initial distance to the closest
correlation matrix and the distance between the result obtained
by alternating projections and the closest correlation matrix is depicted. Each dot shows the result for a simulation with a certain
number of missing data points.
The proofs are given in Appendix C. We now present a concrete application.
Consider random variables
X1 , . . . , Xn
Xl =
(7.1)
that arise as
k
X
cl,j Yj ,
j=1
where
{Yj }kj=1
Pk
and where
are independent random variables with a unit normal distribution,
2
j=1 cl,j
= 1.
Then the correlation matrix
C(l, l0 ) = var(Xl , Xl0 ) =
k
X
C
is given by
cl,j cl0 ,j ,
j=1
n × n-matrix in K with rank k and ones on the diagonal, i.e.
N ∈ N be given and X ∈ MN,n a matrix where each row is a
n
T
realization of (Xl )l=1 . If N is large we should have X X/N ≈ C . However, in real
situations, for instance in nancial applications [27], some of the values X(j, l) are
which clearly is an
C ∈ V1 ∩ V2 .
Let
22
FREDRIK ANDERSSON AND MARCUS CARLSSON
not available. For the computation of the correlation matrix corresponding to (7.1),
it is then common to replace the missing elements by zeros. Given an index set
that indicates the missing data points, let
X̃I
I
denote the matrix corresponding to
|I| denote the
X̃IT X̃I /N will no longer be in
V1 ∩ V2 . The distance between X̃IT X̃I /N and V1 ∩ V2 will tend to increase as the
T
number of missing data points increases. π(X̃I X̃I /N ) is known to be computable
X
where the missing elements has been replaced by zeros, and let
number of missing elements. Generically, the matrix
by using semi denite programming [49, Section 4.4].
We have used the SDPT3
implementation [48, 51].
Figure 9 shows results for 1000 simulations with varying number of missing
B = X̃IT X̃I /N . The plot displays logarithmic
kB∞ −π(B)k
values of the relative error
kB−π(B)k as a function of logarithmic values of the
5
distance kB − π(B)k to V1 ∩ V2 . The matrices X are of size 10 × 50 and have
3
6
rank k = 10. The number of missing data points varied linearly from 10 to 10 .
elements to generate initial matrices
An increasing linear trend between the errors can clearly be seen.
This should
be compared with Theorem 5.1, which predicts that we should see an increasing
function.
8. Appendix A
A be a point in K. We will
φ(0) = A and that the domain of
denition of φ is BRm (0, r) for some r > 0. We also let s be such that (2.1) holds.
Proposition 2.1 Let α : K → M be any C p −map, where M is a C p -manifold
and p ≥ 1. Then the map φ−1 ◦ α is also C p (on its natural domain of denition).
Recall the map
φ
introduced in Section 2 and let
throughout without restriction assume that
Proof.
Given
and dene
x0 ∈ BRm (0, r)
f1 , . . . , fn−m ∈ K with the
⊥
TM (φ(x0 )) = Span {f1 , . . . , fn−m }
pick
ωx0 : BRm (0, r) × Rn−m → K
property that
via
ωx0 (x, y) = φ(x) +
n−m
X
yi fi .
i=1
By the inverse function theorem ([9] 0.2.22)
of
(x0 , 0). The proposition now
φ−1 ◦ α = ωx−1
◦ α.
0
ωx0
follows by noting
we have
Proposition 2.2 Let
function of A.
Proof.
Given
A ∈ Im φ,
set
C p -inverse in a neighborhood
that for values of α near φ(x0 )
has a
M be a C 1 -manifold. Then PTM (A) is a continuous
M = dφ(φ−1 (A)).
It is easy to see that
PTM (A) = M (M M )−1 M ∗ .
The conclusion now follows as
∗
dφ
and
φ−1
are continuous.
Propositions 2.3 and 2.4 are a bit harder. We begin with a lemma.
Lemma 8.1. If B ∈ K is given and A is the closest point in M, then B − A ⊥
TM (A). Moreover, kφ(x) − φ(y)k/kx − yk is uniformly bounded above and below
for x, y in any B(0, r0 ), r0 < r.
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
Proof.
23
Note that
φ(x) = A + dφ(0)x + o(x)
where
o stands for a function with the property that o(x)/kxk extends by continuity
to 0 and takes the value 0 there. Thus we have
kφ(x) − Bk2 = kA + dφ(0)(x) + o(x) − Bk2 =
= kA − Bk2 + 2hA − B, dφ(0)xi + o(kxk)
x's.
and hence the scalar product needs to be zero for all
w = (φ(y) − φ(x))/kφ(y) − φ(x)k and apply
hφ(x + (y − x)t) − φ(x), wi to conclude that
set
For the second claim,
the mean value theorem to
γ(t) =
kφ(y) − φ(x)k = γ(1) − γ(0) = hdφ(z)(y − x), wi
z
for some
on the line between
the singular values of
dφ(z),
x
and
y.
Letting
σ1 (dφ(z)), . . . , σm (dφ(z))
{σm (dφ(z))}ky − xk ≤ kφ(y) − φ(x)k ≤
inf
z∈BRm (0,r 0 )
Now,
dφ(z)
depends continuously on
z
on the matrix entries [21, p191]. Since
for all
sup
x ∈ B(0, r),
denote
we thus have
sup
{σ1 (dφ(z))}ky − xk.
z∈BRm (0,r 0 )
and its singular values depend continuously
φ
so by compactness of
is an immersion we have
cl(B(0, r0 ))
σm (dφ(x)) 6= 0
inf and
we get that both the
amount to nite positive numbers, as desired.
Since
s
s0 in the below formulation.
M be a C 2 -manifold. Given any xed A ∈ M, there
has been specied above, we work with
Proposition 2.3. Let
exists s0 > 0 and a C 2 -map
π : BK (A, s0 ) → M
such that for all B ∈ BK (A, s0 ) there exists a unique closest point in M which is
given by π(B). Moreover, C ∈ M ∩ BK (A, s0 ) equals π(B) if and only if B − C ⊥
TM (C).
Proof.
e.g.
By standard dierential geometry there exists an
C 1 -functions f1 , . . . , fn−m : BRm (0, r0 ) → K with the property that
⊥
TM (φ(x)) = Span {f1 (x), . . . , fn−m (x)}
for all
M,
r0 < r
We repeat the standard construction of a tubular neighborhood of
[9] or [40]).
(see
and
x ∈ BRm (0, r0 ). Moreover, applying the Gram-Schmidt process we may
{f1 (x), . . . , fn−m (x)} is an orthonormal set for all x ∈ BRm (0, r0 )
τ : BRm (0, r0 ) × Rn−m → K via
assume that
Dene
τ (x, y) = φ(x) +
n−m
X
yi fi (x).
i=1
τ
is
C1
by construction, so the inverse function theorem implies that there exists
such that
s0
τ
such that
(8.1)
is a dieomorphism from
s0 < r1 /4, 2s0 < s
and
BRn (0, r1 )
onto a neighborhood of
BK (A, 2s0 ) ⊂ τ (B(0, r1 /2)).
A.
r1
Choose
24
FREDRIK ANDERSSON AND MARCUS CARLSSON
B ∈ BK (A, s0 ) there thus exists a
τ ((xB , yB )) and k(xB , yB )k ≤ r1 /2. We dene
Given any
unique
(xB , yB )
such that
B =
π(B) = φ(xB ).
To see that
π
note that on
1
C -map, let θ : Rm × Rn−m → Rm
B(A, s0 ) we have
is a
be given by
θ((x, y)) = x
and
π = φ ◦ θ ◦ (τ |BRn (0,r1 ) )−1 .
For the fact that
π
is actually
C 2,
(which is not needed in this paper), we refer
π(B) have the desired properties.
C ∈ M is a closest point to B . Since kA − Bk < s0 we clearly must have
kC − Ak < 2s0 so by (2.1) and (8.1) there exists a xC ∈ BRm (0, r1 /2) with φ(xC ) =
C . Moreover, by Lemma 8.1 B − C ⊥ TM (C) so (by the orthonormality of the f 's
n−m
0
at xC ) there exists a y ∈ R
with τ (xC , y) = B and kyk = kB − Ck < s < r1 /4.
Since k(xC , y)k ≤ r1 /2 + r1 /4 and τ is a bijection in B(0, r1 ), we conclude that
(xC , y) = (xB , yB ). Hence π(B) = φ(xB ) = φ(xC ) = C . This establishes the rst
to [31] or Sec. 14.6 in [23]. We now show that
Suppose
part of the proposition.
π(B) − B ∈ Span {f1 (xB ), . . . , fn−m (xB )} is orthogonal
TM (φ(xB )) = TM (π(B)). Conversely, let C be as in the second part of the
proposition. As above we have C = φ(xC ) with kxC k < r1 /2 and there exists a
y with B = τ ((xC , y)) and kyk = kB − Ck < 2s0 < r1 /2. As earlier, this implies
C = π(B), as desired.
By the construction,
to
Proposition 2.4 Let
M be a locally C 2 -manifold at A. For each > 0 there
exists s > 0 such that for all C ∈ BK (A, s ) ∩ M we have
(i)
(ii)
Proof.
dist(D, T̃M (C)) < kD − Ck, D ∈ B(A, s ) ∩ M.
dist(D, M) < kD − Ck,
D ∈ B(A, s ) ∩ T̃M (C).
We rst prove
dφ(xC )(xD − xC ))
(i).
s < s. Set w = D − (C +
dist(D, T̃M (C)) ≤ kwk. Apply the mean value
We may clearly assume that
and note that
theorem to the function
D w E
γ(t) = φ (xD − xC )t + xC − φ(xC ) − dφ(xC )(xD − xC )t,
kwk
y on the line [xC , xD ] between xC and xD such
D
w E
kwk = γ(1) − γ(0) = dφ(y + xC ) − dφ(xC ) (xD − xC ),
.
kwk
to conclude that there exists a
that
By the Cauchy-Schwartz's inequality we get
dist(D, T̃M (C)) ≤ kdφ(y + xC ) − dφ(xC )kkxD − xC k
for some
y ∈ [xC , xD ].
By Lemma 8.1 and the equicontinuity of continuous functions
r < r such that dist(D, T̃M (C)) < kD − Ck for
xD , xC ∈ B(0, r ). As in the proof of Proposition 2.3 we can now dene τ and
use it to pick an s such that xC , xD ∈ B(0, r ) whenever C, D ∈ B(A, s ). Hence
(i) holds.
The proof of (ii) is similar. We assume without restriction that C ∈ Im φ and let
y be such that D = C +dφ(xC )y . We may chose s small enough that kxC +yk < r,
on compact sets, we can choose an
all
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
and then clearly
dist(D, M) ≤ kφ(xC + y) − Dk.
Setting
applying the mean value theorem to
γ(t) =
we easily obtain
φ(xC + yt) − C − dφ(xC )yt,
25
w = φ(xC + y) − D
w
kwk
and
dist(D, M) ≤ kwk ≤ kdφ(xC + t0 y) − dφ(xC )kkyk
for some
t0 ∈ [0, 1].
We omit the remaining details.
9. Appendix B
When dealing with concepts such as dimension or the ideal generated by a variety,
we will use
R
and
C
as subscripts when there can be confusion about which eld is
involved. To begin, a short argument shows that
IC (VZar ) = IR (V) + iIR (V),
(9.1)
from which (6.2) easily follows.
(∂z1 , . . . , ∂zn )
where
∂zj
In case we work over
C,
we dene
∇
as
∇ =
refers to the formal partial derivatives, or, which is the
same, the standard derivatives for analytic functions (see e.g. Denition 3.3, Ch.
II of [30]). We need the following classical result from algebraic geometry. See e.g.
Theorem 3 and the comments following it, Ch. II, Sec. 1. [47].
Theorem 9.1. Let V be an irreducible complex algebraic variety with algebraic
dimension d. Then dim C Span C {∇p(A) : p ∈ IC (V)} < n − d for all singular
points A ∈ V and dim C Span C {∇p(A) : p ∈ IC (V)} = n − d for all non-singular
points A ∈ V . Moreover the set of singular points form a complex variety of lower
dimension.
P
Rn
If
(in
is a set of polynomials, we will write
or
Cn
V(P ) for the variety of common zeroes
S1 and S2 and a common
depending on the context). Given two sets
loc
A, we will write that S1 = S2 near A, if there exists an open set U containing
A such that S1 ∩ U = S2 ∩ U . Theorem 9.1 shows that for an irreducible variety, the
point
non-singular points coincide with those of maximal rank in the work of H. Whitney
[53].
This observation combined with Sections 6 and 7 of that paper yields the
following key result.
Theorem 9.2. Let
V be an irreducible complex algebraic manifold and A a nonsingular point. Given any p1 , . . . , pn−d ∈ I(V) such that ∇p1 (A), . . . , ∇pn−d (A)
loc
are linearly independent, we have V = V({p1 , . . . , pn−d }) near A.
The remark after Proposition 6.4 now easily follows from the above results and
the complex version of the implicit function theorem (see e.g. Theorem 3.5, Ch. II,
[30]). In order to transfer these results to the real setting, we rst need a lemma.
Lemma 9.3. A real algebraic variety V is irreducible in Rn if and only if VZar is
irreducible in Cn .
Proof.
V
Rn
into smaller varieties as U1 ∪ U2 ,
(U1 )Zar ∪ (U2 )Zar is a nontrivial decomposition of VZar . Conversely,
n
let VZar = U1 ∪ U2 be a nontrivial decomposition of VZar in C . Let IC (Ui ) denote
the respective complex ideals, i = 1, 2, and set Ii = (Re IC (Ui )) ∪ (Im IC (Ui )). The
n
real variety corresponding to each Ii clearly coincides with Ui ∩ R , which thus is
If
has a nontrivial decomposition in
then clearly
26
FREDRIK ANDERSSON AND MARCUS CARLSSON
a variety in
Rn .
Since VZar is
U1 or U2 , and
of V .
the smallest complex variety including
subset of either
decomposition
hence
V = (U1 ∩ Rn ) ∪ (U2 ∩ Rn )
V, V
is not a
is a nontrivial real
Lemma 9.4. Let V be an irreducible real algebraic variety such that VZar has
algebraic dimension d. Then dim R NV (A) < n − d for all singular points A ∈ V
and dim R NV (A) = n − d for all non-singular points A ∈ V .
Proof.
VZar is irreducible, by Lemma 9.3, and
VZar . Given any real vector space V , it follows by
dimR V = dimC (V + iV ). By (9.1) we thus get
First note that
applies to
hence Theorem 9.1
linear algebra that
dim R Span R {∇p(A) : p ∈ IR (V)} = dim C Span {∇p(A) : p ∈ IC (V)}
for all
A ∈ Rn ,
and hence the lemma follows by Theorem 9.1.
Lemma 9.5. Let
V be an irreducible real algebraic variety such that VZar has
algebraic dimension d. Then V ns is a C ∞ -manifold of dimension d and V \ V ns
is a real algebraic manifold such that (V \ V ns )Zar has algebraic dimension strictly
less than d.
Proof.
By Lemma 9.3,
singular points of
d.
Hence
V \ V ns
VZar
VZar
is irreducible.
By Theorem 9.1, we have that the
form a proper subvariety of dimension strictly less than
is included in a complex subvariety of lower dimension than
d.
Since the algebraic dimension decreases when taking subvarieties, we conclude that
dim(V \ V ns )Zar < d. If V ns would be empty, (V \ V ns )Zar would include all of
V , contradicting the denition of VZar as the smallest complex variety including
V . Finally, given A ∈ V ns , Theorem 9.2 and Lemma 9.4 imply that we can nd
p1 , . . . , pn−d ∈ IR (V) with linearly independent derivatives at A such that
loc
(9.2)
near
A.
The fact that
from Theorem 2.1.2
(ii)
V ns
V ns = V({p1 , . . . , pn−d })
is a
d-dimensional C ∞ -manifold
now follows directly
in [9].
We are now ready to prove the results in Section 6.
Proof of Proposition 6.2 and 6.4.
have algebraic dimension
d.
We begin with Proposition 6.2.
d, and
(V \ V ns )Zar can
is a non-empty manifold of dimension
strictly less than
d.
Thus
V is irreducible.
(V \ V ns )Zar has
First assume that
Let
VZar
V ns
By Lemma 9.5,
algebraic dimension
be decomposed into nitely many
irreducible components of dimension strictly less than
d,
(see Section 4.6 and 9.4
in [15]). Lemma 9.5 can then be applied to the real part of each such component.
Continuing like this, the dimension drops at each step and hence the process must
terminate. This process will give us a decomposition
has dimension
j.
V = ∪dj=0 Mj
where each
Mj
Now, such decompositions are not unique, but basic dierential
geometry implies that the number
d
is an invariant of
V.
V = ∪m
j=0 M̃j
m > d. Let φ be
Indeed, let
M̃m 6= 0, and suppose that
M̃m , as in (2.1). φ is then dened on an open subset
φ−1 (Mj ) are manifolds of dimension strictly less than
be another such decomposition with
a chart covering a patch of
U of Rm , and the subsets
m. Hence each one has Lebesgue measure zero, which his not compatible with that
their union should equal U. Reversing the roles of m and d, Proposition 6.2 follows
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
27
in the irreducible case. Incidentally, we have also shown the rst part of Proposition
6.4.
If
V
is not irreducible, we can apply the above argument to each of its irreducible
components. Since the dimension of
VZar
is the maximum of the dimension of its
components (Proposition 8, Sec. 9.6, [15]), Proposition 6.2 follows as above.
A ∈ V ns , the identity (6.3) follows by (9.2) and the implicit function
theorem. This establishes the second part of Proposition 6.4 and we are done.
Proof of Proposition 6.3. This is now immediate by Lemma 9.4.
Proof of Proposition 6.5. VZar is a strict submanifold of both (V1 )Zar and
(V2 )Zar , and all three are irreducible by assumption and Lemma 9.3. Hence VZar
Finally, with
has strictly lower dimension than the other two, by basic algebraic geometry, see e.g.
Theorem 1, Chap I, Sec. 6, in [47]. By Lemma 9.5 the manifold
ns
and
lower dimension than both V1
intersection at any intersection point
3.2 means that the angle at
Proof of Theorem 6.6.
A
V2ns .
V ns
has strictly
This means that these have a proper
A ∈ V ns ∩ V1ns ∩ V2ns ,
which by Proposition
is dened, as desired.
By Proposition 6.5, all points
non-trivial, (i.e. the angle between
V1ns and V2ns
A ∈ V ns ∩ V1ns ∩ V2ns
are
exists). We rst prove the latter
statement. By Proposition 6.4, (6.4) implies that
dim (NV1 (A) + NV2 (A)) ≥ n − m.
(9.3)
Since obviously
NV (A) ⊃ NV1 (A) + NV2 (A),
(9.4)
dim(NV (A)) ≥ n − m,
we have
which combined with Lemma 9.4 and Proposition
6.2 shows that in fact we must have
A ∈ V ns
dim(NV (A)) = n − m. Moreover,
NV (A) = NV1 (A) + NV2 (A) which,
and
combined with (9.3) and (9.4) this implies that
upon taking the complement and recalling (6.3), yields
TV1ns (A) ∩ TV2ns (A) = TV ns (A).
A is non-tangential, so A ∈ V ns,nt . For the
ns
ns
ns
rst statement, we note that V \ (V1 ∩ V2 ∩ V
) = (V \ V1ns ) ∪ (V \ V2ns ) ∪ (V \ V ns )
ns
ns
and V \ Vi
= V ∩ (Vi \ Vi ) for i = 1, 2. Hence V \ (V1ns ∩ V2ns ∩ V ns ) is a real
algebraic variety by Lemma 9.5 and the trivial fact that unions and intersections
ns
ns
ns
ns,nt
of varieties yield new varieties. Now, suppose A ∈ V1 ∩ V2 ∩ V
is not in V
.
Then it is tangential, which by the earlier arguments happens if and only if
By Proposition 3.5 we conclude that
dim(NV1 (A) + NV2 (A)) < n − m.
(9.5)
By Hilbert's basis theorem we can pick nite sets
is generated by these sets for
of each
{pj,l }j,l
l = 1, 2.
Let
M
{pj,1 } and {pj,2 } such that IR (Vl )
be the matrix with the gradients
as columns. The condition (9.5) can then be reformulated as the
vanishing of the determinants of all
(n − m) × (n − m)
each such determinant is a polynomial, we conclude that
submatrices of
V \ V ns,nt
M.
Since
is dened by the
vanishing of a nite number of polynomials, so it is a real algebraic variety. Finally,
if
V ns,nt
is not void, then
(V \ (V ns,nt ))Zar
latter is irreducible by Lemma 9.3, and hence
than
[47]).
VZar
IR (V)
(V \ (V ns,nt ))Zar
by standard algebraic geometry, (see e.g.
Proof of Proposition 6.8.
if
is a proper subvariety of
VZar .
The
has lower dimension
Theorem 1, Ch.
I, Sec.
6,
It is well-known that
is prime, i.e. if and only if
f g ∈ IR (V)
V
is irreducible if and only
implies that either
f ∈ IR (V)
or
28
FREDRIK ANDERSSON AND MARCUS CARLSSON
g ∈ IR (V)
(see e.g.
Proposition 3, Chap.
we have such a product
xed
A ∈ V,
let
i0 ∈ I
fg
and that
V
4, Sec.
5 of [15]).
be as in Denition 6.7. Then
by analyticity implies that one of the functions, say
Note that
V
Suppose now that
is covered by analytic patches. Given any
(f ◦ φi0 )(g ◦ φi0 ) ≡ 0, which
f ◦ φi0 , vanishes identically.
is path connected. To see this, use the analytic patches to show that
any path-connected component is both open and closed, which gives the desired
conclusion since
V
is connected (see e.g. Sec. 23-25 in [41]). Now, let
other point and let
γ
be a continuous path connecting
A
B.
with
B∈V
be any
The image
Im γ
{BRn (C, rC )}C∈Im
γ an open covering, where rC is as in Denition
L
6.7. We pick a nite subcovering at the points {Cl }l=0 with C0 = A and CL = B .
Clearly these can be ordered such that V ∩ B(Cl , rCl ) ∩ B(Cl+1 , rCl+1 ) 6= ∅. Also let
is compact and
φil of B(Cl , rCl ) in accordance with Denition 6.7,
i0 already has been chosen. Let D ∈ V ∩ B(C0 , rC0 ) ∩ B(C1 , rC1 ) be given
and let R be a radius such that B(D, R) ⊂ B(C0 , rC0 ) ∩ B(C1 , rC1 ). By assumption,
f vanishes on all points of V ∩ B(C0 , rC0 ), and hence f ◦ φi1 vanishes on the open
−1
set φi (B(D, R)), which by analyticity means that it vanishes identically on Ωi1
1
(since it is assumed to be connected). By induction it follows that f (B) = 0, and
il
be the index of the covering
where
the rst part is proved. The second is a simple consequence of continuity. We omit
the details.
Proof of Proposition 6.9.
that
V ns
is dense in
V.
By Proposition 6.8,
Consider the case when
V
V
is irreducible. We rst show
is covered by analytic patches,
(the proof in the second case is easier and will be omitted).
nontrivial subvariety by Lemma 9.5, there exists a polynomial
V \ V ns
Since
f
V \ V ns
is a
which vanishes on
V \ V ns contains an open set, then the argument
in Proposition 6.8 shows that f ≡ 0 on V , a contradiction.
d
Now, let θ : U → V be the bijection in question, where U ⊂ R and V ⊂ V are
ns
open. Let m be the dimension of V and pick any A ∈ V
∩ V . By Proposition
m
6.4 and (2.1), there exists open sets Ũ ⊂ R
and Ṽ ⊂ V containing A and a
C ∞ -bijection φ : Ũ → Ṽ . Moreover, by Proposition 2.1, φ−1 ◦ θ is bijective and
−1
dierentiable between the open sets θ
(Ṽ ) ⊂ Rd and Ũ ⊂ Rd . That m = d is now
a well-known consequence of the implicit function theorem.
but not on
V.
However, if
10. Appendix C
If we were to write out all the details, this section would get rather long. Considering that it is just an illustration, we will be a bit brief.
Proof of Proposition 7.1.
Given a matrix
A ∈ K,
we can consider all elements
above and on the diagonal as variables, and the remaining to be determined by
AT = A.
K is a linear space of dimension n(n+1)/2. The statements
V2 is a real algebraic variety, note
that B has rank greater than k if and only if one can nd a non-zero (k + 1) × (k +
1) invertible minor (that is, a matrix obtained by deleting n − (k + 1) rows and
It follows that
concerning
V1
are now immediate. To see that
columns). The determinant of each such minor is a polynomial, (more precisely,
the determinant composed with the map that identies
K
with
R(n
2
+n)/2
), and
is clearly the variety obtained from the collection of such polynomials. Thus
a real algebraic variety. The same is true for
adding the algebraic equations
V2 .
We denote by
Mi,j
the set
V = V1 ∩ V2
V2
V2
is
since it is obtained by
{Bj,j = 1}nj=1 to those dening V2 . We now study
of i × j matrices with real entries. By the spectral
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
theorem, each
B∈K
with
Rank B ≤ k
U ∈ Mn,k
can be written as
B = U ΣU T
(10.1)
where
29
and
Σ ∈ Mk,k
is a diagonal matrix, and conversely any ma-
trix given by such a product has rank less than or equal to
k.
We see that
can be covered with one real polynomial, which by Proposition 6.8 shows that
is irreducible and Proposition 6.9 applies.
V2
V2
In order to determine the dimension,
consider the open subset of matrices with positive eigenvalues. This has the easier
U U T , where again U ∈ Mn,k is arbitrary. However, neither this is
T
bijective. In fact, if B = U U , then any other such parametrization of B is given
T
T
by B = (U W )(U W ) where W ∈ Mk,k is unitary. Pick any B0 = U0 U0 such that
the upper k × k submatrix of U0 is invertible, and denote this by V0 . By the theory
of QR-factorizations, there is a unique unitary matrix W0 for which V0 W0 is lower
parametrization
triangular and has positive values on the diagonal [32, Theorem 1, p. 262]. The
(k−1)k
independent
2
(k−1)k
nk− 2
variables, so we can identify the set of such matrices with R
. Denote the
(k−1)k
(k−1)k
nk− 2
inverse of this identication by ι : R
→ Mn,k , and let Ω ⊂ Rnk− 2 be
n×k
set of
matrices that are lower triangular contain
nk −
the open set corresponding to the matrices with strictly positive diagonal elements.
Dene
φ : Ω → V2
by
φ(y) = ι(y)(ι(y))T .
(10.2)
φ
It is easy to see that
is in bijective correspondence with an open set including
Thus Proposition 6.9 implies that V2 has
2nk−k2 +k
, as desired.
2
We turn our attention to V and rst prove that it can be covered with analytic
B0 ,
and moreover
φ
is a polynomial.
dimension
σ : {1, . . . , n} → {1, . . . , k} and τ : {1, . . . , n} → {−1, 1} be given and
U ∈ Mn,k where Uj,σj = xj is an undetermined variable whereas all
other values are xed. Denote the j th row of U by Uj , and let Σ ∈ Mk,k be a xed
T
diagonal matrix. Then Uj ΣUj = 1 is a quadratic equation with xj as unknown,
patches. Let
consider all
which may have 0, 1, 2 or innitely many real solutions. Suppose the remaining
values of
U
are such that this has two solutions for all
be the solution whose sign coincides with
by
Ũ
τ (j).
1 ≤ j ≤ n,
and x
xj
to
Denote the corresponding matrix
and note that it is a real analytic function if we now consider the remaining
Ũ as variables. These variables and the values on the diagonal of Σ are
n(k − 1) + k in number, and so can be identied with points y in an open subset of
Rnk−n+k . Let Ω be a particular connected component of this open set. Consider
Ũ and Σ as functions of y on Ω, in the obvious way, and set
values of
ψσ,τ,Ω (y) = Ũ (y)Σ(y)(Ũ (y))T ,
(10.3)
Let
I
be the set of all possible triples
sees that each
B ∈V
σ, τ, Ω.
{ψi }i∈I
V
1
is irreducible, but rst we need to
is connected. The proof gets very lengthy and technical, so we will
only outline the details. The idea is that any
matrix
ψi , i ∈ I . It
V . We
is a covering with analytic patches of
wish to use Proposition 6.8 to conclude that
V
By the spectral theorem one easily
is in the image of at least one of these maps
is now not hard to see that
show that
y ∈ Ω.
B ∈V
is path connected with the
with all elements equal to 1. To see this, rst note that the subset of
that can be represented as (10.1) with all elements of
U
Hence it suces to nd a path from such an element to
non-zero, is dense in
1.
Let
B
V
V.
be xed. Now,
30
FREDRIK ANDERSSON AND MARCUS CARLSSON
all values in
Σ
are not negative, for then the diagonal values of
We can assume that the diagonal elements of
Σ1,1 = 1.
Pick
σ
such that
σ(j) = k
for all
Σ
j,
B
would be as well.
are ordered decreasingly and that
and choose
τ
and
Ω
such that the
representation (10.1) can be written in the form (10.3). Now, if the second diagonal
value in
Ω.
Σ is negative, we may continuously change it until it is not, without leaving
y corresponding to the rst and second column of Ũ can
Then the values of
be continuously moved until all elements of the rst column are positive. At this
point, we can reduce all values of
Ũ
except the rst column to zero, increasing the
rst value of each row whenever necessary to stay in
Ω.
Then we can move
y so that
the values in the rst column become the same. Finally, we can let these values
increase simultaneously until they reach 1. We have now obtained the matrix
and conclude that
V
1,
is connected and hence also irreducible, as desired.
Finally, we shall determine the dimension of
V.
Consider again the map
ι
intro-
duced earlier, with the dierence that this time the last lower diagonal element in
each row is not a variable, but instead determined by the other variables in that
row and the constraint that it be strictly positive and that the norm of the row be
1. The number of free variables is thus
2nk−k2 +k
2
− n,
and the above construction
thus naturally denes a real analytic map on an open subset
Denote this map by
θ
and dene
ψ:Ξ→V
Ξ
of
R
2nk−k2 +k
−n
2
.
by
ψ(y) = θ(y)(θ(y))T .
It is not hard to see that
6.9 the dimension of
V
ψ is a bijection with an open subset of V , so by Proposition
2nk−k2 +k
− n, as desired.
2
is
Proof of Proposition 7.2.
By Proposition 6.3 and 7.1 we need to show that
dim NV2 (A) = (n2 + n)/2 − (2nk − k 2 + k)/2 = (n − k + 1)(n − k)/2
Rank (A) = k. By the same propositions we already have that
dim NV2 (A) ≤ (n − k + 1)(n − k)/2, so it suces to show that this inequality is
strict when Rank (A) < k and that the reverse inequality holds when Rank (A) = k .
Based on the representation (10.1), it is easily seen that each A ∈ V2 can be written
if and only if
(10.4)
A = U ΣU T
Σ and U lie in Mn,n , U is unitary and Σ is diagonal with only
0's after the k th element. In this proof the particular identication of K with
2
2
R(n +n)/2 is important. We dene ω : R(n +n)/2 → K by letting the rst n entries
where now both
be placed on the diagonal and the remaining ones be distributed over the upper
triangular part, but multiplied by the factor
√
1/ 2.
Finally, the lower triangular
AT = A. This identication will be implicit. For example, if p is
a polynomial on K then we will write ∇p instead of the correct ω(∇(p ◦ ω)). Note
however that ∇p depends on ω . Now, given a polynomial p ∈ I(V2 ) and C ∈ Mn,n ,
qC (·) = p(C T · C) is clearly also in I(V2 ). Due to the particular choice of ω , we have
T
T
that ∇qC (B) = C ∇p(C BC)C as can be veried by direct computation. Letting
A be xed of rank j ≤ k , it is easy to use (10.4) to produce an invertible matrix C
T
such that C AC = Ij , where Ij ∈ K is the diagonal matrix whose rst j diagonal
part is dened by
values are 1 and 0 elsewhere. In particular
∇qC (A) = C T ∇p(Ij )C,
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
which implies that
of
K
31
dim NV2 (A) = dim NV2 (Ij ). Now, all Mk+1,k+1 subdeterminants
I(V2 ) and their derivatives at Ik are easily computed by
form polynomials in
hand. In this way one easily gets
dim NV2 (Ik ) ≥ (n − k + 1)(n − k)/2,
k element of V2 is non-singular. Conversely, if j < k ,
u ∈ Rn as a row-vector and dene the map θu : R → V2 via
θu (x) = Ij + xuT u. Letting {el }nl=1 be the standard basis of Rn and considering
θel as well as all dierences θel +el0 − θel − θel0 , one easily sees that
d
n
θu (0) : u ∈ R
=K
Span
dx
which proves that any rank
consider a xed
and hence
dim NV2 (Ij ) = 0.
This shows that
Ij
is singular, by the remarks in the
beginning of the proof.
Finally, we shall establish that
V ns,nt
is not void. By Proposition 6.5, Theorem
6.6, Proposition 7.1 and the rst part of this proposition, it suces to show that
2nk − k 2 + k
−n
2
T
for some point A which has rank k . We choose the point A = U U where U ∈ Mn×k
has a 1 on the last lower diagonal element (i.e. with index (j, j) for j < k and (j, k)
for j ≤ k ) of each row, and zeroes elsewhere. Recall the map φ in (10.2) and let
dim(TV1 (A) ∩ TV2 (A)) ≤
(10.5)
y∈R
2nk−k2 +k
2
U = φ(y). Given (i, j) let E(i,j) be the matrix with a 1
(j, i) and zeroes elsewhere. The partial derivatives of φ at y
contain multiples of all E(i,j) with i ≤ j < k , as well as n − k + 1 derivatives related
to the last row of U , which we denote by F1 , . . . , Fn−k+1 . These are not hard to
compute, but we only need the fact each Fl has precisely one non-zero diagonal
value on one of the n − k + 1 last elements on the diagonal, (with distinct positions
for distinct l's). Since TV1 (A) = Span {Ei,j : i 6= j}, it is easily seen that
on positions
be such that
(i, j)
and
Span (TV1 (A) ∩ TV2 (A)) = Span {Ei,j : i < j < k}.
n(n−1)
These are
2
−
(n−k+1)(n−k)
in number, and (10.5) follows.
2
11. Acknowledgements
This work was supported by the Swedish Research Council and the Swedish
Foundation for International Cooperation in Research and Higher Education, as
well as Dirección de Investigación Cientíca y Technológica del Universidad de
Santiago de Chile, Chile. Part of the work was conducted while Marcus Carlsson
was employed at Universidad de Santiago de Chile.
We thank Arne Meurman
for fruitful discussions. We also would like to thank one of the reviewers for the
constructive criticism and useful suggestions that have helped to improve the quality
of the paper.
References
Optimization algorithms on matrix mani-
[1]
P. A. Absil, R. Mahony, and R. Sepulchre,
[2]
H. Attouch, J. Bolte, P. Redont, and A. Soubeyran,
folds, Princeton Univ Pr, 2008.
Proximal alternating minimization and projection methods for nonconvex problems: An approach based on the KurdykaLojasiewicz inequality, Mathematics of Operations Research, 35 (2010), pp. 438457.
32
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
FREDRIK ANDERSSON AND MARCUS CARLSSON
A generalization of the Friedrichs angle and
the method of alternatingprojections, Comptes Rendus Mathematique, 348 (2010), pp. 53 56.
S. Basu, R. Pollack, and M.-F. Roy, Algorithms in real algebraic geometry, vol. 10 of
Algorithms and Computation in Mathematics, Springer-Verlag, Berlin, 2006.
H. H. Bauschke and J. M. Borwein, On the convergence of von Neumann's alternating
projection algorithm for two sets, Set-Valued Analysis, 1 (1993), pp. 185212.
, Dykstra's alternating projection algorithm for two sets, Journal of Approximation
Theory, 79 (1994), pp. 418443.
H. H. Bauschke and J. M. Borwein, On projection algorithms for solving convex feasibility
problems, SIAM Rev., 38 (1996), pp. 367426.
H. H. Bauschke, D. Noll, A. Celler, and J. M. Borwein, An EM algorithm for
dynamic SPECT, IEEE TRANS. MED. IMAG, 18 (1999).
M. Berger and B. Gostiaux, Dierential Geometry: Manifolds, Curves and Surfaces.,
Springer Verlag, 1988.
S. Boyd, N. Parikh, E. Chu, B. Peleato, and J. Eckstein, Distributed optimization
and statistical learning via the alternating direction method of multipliers, Now Publishers,
2011.
S. Boyd and L. Xiao, Least squares covariance matrix adjustment., SIAM J. Matrix Anal.
Appl., 27 (2005), pp. 532546.
L. M. Brègman, Finding the common point of convex sets by the method of successive
projection, Dokl. Akad. Nauk SSSR, 162 (1965), pp. 487490.
J. A. Cadzow, Signalenhancement-a composite property mapping algorithm, Acoustics,
Speech and Signal Processing, IEEE Transactions on, 36 (1988), pp. 49 62.
P. L. Combettes and H. J. Trussell, Method of successive projections for nding a
common point of sets in metric spaces, Journal of optimization theory and applications, 67
(1990), pp. 487507.
D. A. Cox, J. B. Little, and D. O'Shea, Ideals, varieties, and algorithms: an introduction to computational algebraic geometry and commutative algebra, vol. 10, Springer Verlag,
2007.
F. Deutsch, Rate of convergence of the method of alternating projections, Parametric optimization and approximation, (1985), pp. 96107.
F. Deutsch, Best approximation in inner product spaces, CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC, 7, Springer-Verlag, New York, 2001.
F. Deutsch and H. Hundal, The rate of convergence for the method of alternating projections, ii, Journal of Mathematical Analysis and Applications, 205 (1997), pp. 381405.
R. L. Dykstra, An algorithm for restricted least squares regression, Journal of the American
Statistical Association, (1983), pp. 837842.
C. Eckart and G. Young, The approximation of one matrix by another of lower rank,
Psychometrika, 1 (1936), pp. 211218.
J. N. Franklin, Matrix theory., Dover Publications in Mineola, N.Y., 2000.
K. Friedrichs, On certain inequalities and characteristic value problems for analytic functions and for functions of two variables, Trans. Amer. Math. Soc, 41 (1937), pp. 321364.
D. Gilbarg and N. Trudinger, Elliptic partial dierential equations of second order,
vol. 224, Springer Verlag, 2001.
K. M. Grigoriadis, A. E. Frazho, and R. E. Skelton, Application of alternating convex
projection methods for computation of positive Toeplitz matrices, Signal Processing, IEEE
Transactions on, 42 (1994), pp. 1873 1875.
I. Grubi²i¢ and R. Pietersz, Ecient rank reduction of correlation matrices, Linear algebra and its applications, 422 (2007), pp. 629653.
L. G. Gubin, B. T. Polyak, and E. V. Raik, The method of projections for nding the
common point of convex sets, USSR Computational Mathematics and Mathematical Physics,
7 (1967), pp. 124.
N. J. Higham, Computing the nearest correlation matrix a problem from nance, IMA
Journal of Numerical Analysis, 22 (2002), pp. 329343.
L. Hörmander, The analysis of linear partial dierential operators. III, Springer, Berlin,
2007.
C. Badea, S. Grivaux, and V. M�ller,
ALTERNATING PROJECTIONS ON NON-TANGENTIAL MANIFOLDS.
33
Error bounds for the method of alternating projections,
Mathematics of Control, Signals, and Systems (MCSS), 1 (1988), pp. 4359.
[30] K. Kendig, Elementary algebraic geometry, Springer-Verlag, 1977.
[31] S. G. Krantz and H. R. Parks, Distance to C k hypersurfaces, Journal of Dierential
Equations, 40 (1981), pp. 116120.
[32] P. D. Lax, Linear algebra and its applications, Pure and Applied Mathematics (Hoboken),
Wiley-Interscience [John Wiley & Sons], Hoboken, NJ, second ed., 2007.
[33] A. Levi and H. Stark, Signal restoration from phase by projections onto convex sets, J.
Opt. Soc. Am., 73 (1983), pp. 810822.
[34] A. S. Lewis, D. Luke, and J. Malick, Local linear convergence for alternating and averaged nonconvex projections, Foundations of Computational Mathematics, 9 (2009), pp. 485
513.
[35] A. S. Lewis and J. Malick, Alternating projections on manifolds, Math. Oper. Res., 33
(2008), pp. 216234.
[36] Y. Li, K. Liu, and J. Razavilar, A parameter estimation scheme for damped sinusoidal
signals based on low-rank hankel approximation, Signal Processing, IEEE Transactions on,
45 (1997), pp. 481 486.
[37] B. Lu, D. Wei, B. L. Evans, and A. C. Bovik, Improved matrix pencil methods, in Signals,
Systems Computers, 1998. Conference Record of the Thirty-Second Asilomar Conference on,
vol. 2, nov 1998, pp. 1433 1437 vol.2.
[38] I. Markovsky, Structured low-rank approximation and its applications, Automatica, 44
(2008), pp. 891909.
[39] R. J. Marks II, Alternating projections onto convex sets, in Deconvolution of images and
spectra (2nd ed.), P. A. Jansson, ed., Academic Press, Inc., 1997, pp. 476501.
[40] S. Montiel, A. Ros, and D. G. Babbitt, Curves and surfaces, vol. 69, American Mathematical Society, 2009.
[41] J. R. Munkres, Topology, Prentice Hall, Upper Saddle River, 2nd ed., 2000.
[42] N. K. Nikolski, Operators, functions, and systems: an easy reading. Vol. 1, vol. 92 of
Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI,
2002. Hardy, Hankel, and Toeplitz, Translated from the French by Andreas Hartmann.
[43] V. U. Prabhu and D. Jalihal, An improved ESPRIT based time-of-arrival estimation
algorithm for vehicular OFDM systems, in Vehicular Technology Conference, 2009. VTC
Spring 2009. IEEE 69th, april 2009, pp. 1 4.
[44] R. T. Rockafellar and R. J.-B. Wets, Variational analysis, vol. 317 of Grundlehren
der Mathematischen Wissenschaften, Springer-Verlag, Berlin, 1998.
[45] A. Sard, The measure of the critical values of dierentiable maps, Bull. Amer. Math. Soc,
48 (1942), pp. 883890.
[46] E. Schmidt, Zur theorie der linearen und nichtlinearen integralgleichungen. iii. teil., Math.
Ann., 65 (1908), pp. 370399.
[47] I. R. Shafarevich, Basic algebraic geometry, Springer-Verlag, New York, 1974. Translated
from the Russian by K. A. Hirsch, Die Grundlehren der mathematischen Wissenschaften,
Band 213.
[48] K. C. Toh, M. J. Todd, and R. H. Tütüncü, SDPT3 a MATLAB software package
for semidenite programming, version 1.3, Optimization Methods and Software, 11 (1999),
pp. 545581.
[49]
, On the implementation and usage of SDPT3a matlab software package for
semidenite-quadratic-linear programming, version 4.0, Handbook on Semidenite, Conic
and Polynomial Optimization, (2012), pp. 715754.
[50] D. Trotman, Singularity Theory, Scientic Publishing Co. Pte. Ltd., Oxford, 2007, ch. Lectures on Real Stratication Theory.
[51] R. H. Tütüncü, K. C. Toh, and M. J. Todd, Solving semidenite-quadratic-linear programs using SDPT3, Mathematical programming, 95 (2003), pp. 189217.
[52] J. von Neumann, Functional Operators, Volume II: The Geometry of Orthogonal Spaces,
Princeton University Press, 1950.
[53] H. Whitney, Elementary structure of real algebraic varieties, The Annals of Mathematics,
66 (1957), pp. 545556.
[29]
S. Kayalar and H. L. Weinert,
34
FREDRIK ANDERSSON AND MARCUS CARLSSON
The method of alternating projections and the method of subspace corrections in Hilbert space, Journal of the American Mathematical Society, 15 (2002),
pp. 573598.
[55] W. I. Zangwill, Nonlinear Programming, Prentice Hall, Englewood Clis, N. J., 1969.
[54]
J. Xu and L. Zikatanov,
Centre for Mathematical Sciences, Lund University, Sweden
E-mail address : [email protected]
Centre for Mathematical Sciences, Lund University, Sweden
E-mail address : [email protected]