An Exact Duality Theory For Semidefinite Programming Based On Sums of Squares
An Exact Duality Theory For Semidefinite Programming Based On Sums of Squares
1. Introduction
A linear matrix inequality (LMI) is a condition of the form
n
X
A(x) = A0 + xi Ai 0 (x ∈ Rn )
i=1
where the Ai are symmetric matrices of the same size and one is interested in the
solutions x ∈ Rn making A(x) positive semidefinite (A(x) 0). The solution
set to such an inequality is a closed convex semialgebraic subset of Rn called a
spectrahedron or an LMI domain. Optimization of linear objective functions over
spectrahedra is called semidefinite programming (SDP) [BV96, To01, WSV00], and
is a subfield of convex optimization. In this article, we are concerned with the
duality theory of SDP from a viewpoint of a real algebraic geometer, and with the
important SDP feasibility problem: When is an LMI feasible; i.e., when is there an
x ∈ Rn satisfying A(x) 0?
A diagonal LMI, where all Ai are diagonal matrices, is just a system of linear
inequalities, and its solution set is a polyhedron. Optimization of linear objective
functions over polyhedra is called linear programming (LP). The ellipsoid method
developed by Shor, Yudin, Nemirovskii and Khachiyan showed at the end of the
1970s for the first time that the LP feasibility problem (and actually the problem
of solving LPs) can be solved in (deterministically) polynomial time (in the bit
model of computation assuming rational coefficients) [Sr86, Chapter 13]. Another
breakthrough came in the 1980s with the introduction of the more practical interior
use quadratic modules from real algebraic geometry. As a side product we intro-
duce a hierarchy for infeasibility of LMIs, whose first stage coincides with strong
infeasibility. Subsection 4.4 contains certificates for boundedness of spectrahedra
which are used to give a Putinar-Schmüdgen-like Positivstellensatz for polynomi-
als positive on bounded spectrahedra. Finally, the article concludes with two brief
sections containing examples illustrating our results, and an application to positive
linear functionals.
2.2. Linear pencils and spectrahedra. We use the term linear pencil as a syn-
onym and abbreviation for symmetric linear matrix polynomial. A linear pencil
A ∈ R[x]α×α can thus be written uniquely as
A = A0 + x1 A1 + · · · + xn An
with Ai ∈ SRα×α . If A ∈ R[x]α×α is a linear pencil, then the condition A(x) 0
(x ∈ Rn ) is called a linear matrix inequality (LMI) and its solution set
SA := {x ∈ Rn | A(x) 0}
is called a spectrahedron (or also an LMI set ). We say that A is infeasible if SA = ∅,
and A is feasible if SA 6= ∅.
Obviously, each spectrahedron is a closed convex semialgebraic subset of Rn .
Optimization of linear objective functions over spectrahedra is called semidefinite
programming (SDP) [BV96, To01, WSV00]. If A ∈ R[x]α×α is a diagonal linear pen-
cil, then A(x) 0 (x ∈ Rn ) is just a (finite) system of (non-strict) linear inequalities
and SA is a (closed convex) polyhedron. Optimization of linear objective functions
over polyhedra is called linear programming (LP). The advantage of LMIs over sys-
tems of linear inequalities (or of spectrahedra over polyhedra, and SDP over LP, re-
spectively) is a considerable gain of expressiveness which makes LMIs an important
tool in many areas of applied and pure mathematics [BEFB94, Go97, Par00, Las10].
SDPs can be solved efficiently using interior point methods [NN94, St00, dK02].
4 IGOR KLEP AND MARKUS SCHWEIGHOFER
√ √
An ideal I ⊆ R[x] is called radical if I = I and real radical if I = R I. We refer
the reader to [BCR98] for further details.
3.1. Weakly feasible and weakly infeasible linear pencils. Recall that the
linear pencil A(x) ∈ R[x]1α×α is infeasible if SA = ∅. We call A strongly infeasible
if
dist {A(x) | x ∈ Rn }, SRα×α
0 > 0,
and weakly infeasible if it is infeasible but is not strongly infeasible. A feasible linear
pencil A is strongly feasible if there is an x ∈ Rn such that A(x) ≻ 0, and weakly
feasible otherwise. To A we associate the convex cone
n X o
CA := c + u∗i Aui | c ∈ R≥0 , ui ∈ Rα
i
n o
= c + tr(AS) | c ∈ R≥0 , S ∈ SRα×α
0 ⊆ R[x]1 .
Proof. Suppose
n
X
A = A0 + xi Ai ∈ R[x]α×α
1
i=1
is strongly infeasible. Then the non-empty convex sets {A(x) | x ∈ Rn } and SRα×α0
can be strictly separated by an affine hyperplane (since their Minkowski sums with a
small ball are still disjoint and can therefore be separated [Ba02, Theorem III.1.2]).
This means that there is a non-zero linear functional
ℓ : SRα×α → R
α×α
and γ ∈ R, with ℓ(SR0 ) ⊆ R>γ and ℓ({A(x) | x ∈ Rn }) ⊆ R<γ . Choose
α×α
B ∈ SR such that
ℓ(A) = tr(AB)
α×α α×α
for all A ∈ SR . Since ℓ(SR0 ) is bounded from below, by the self-duality of
the convex cone of positive semidefinite matrices, 0 6= B 0. Similarly, we obtain
ℓ(Ai ) = 0 for i ∈ {1, . . . ,n}. Note that γ <P0 since 0 = ℓ(0) ∈ R>γ so we can
assume ℓ(A0 ) = −1 by scaling. Writing B = i ui u∗i with ui ∈ Rα , we obtain
X X
−1 = ℓ(A0 ) = ℓ(A(x)) = tr(A(x) ui u∗i ) = u∗i A(x)ui .
i i
n ∗
P
for all x ∈ R . Hence −1 = i ui Aui
∈ CAP
.
Conversely, if −1 ∈ CA , i.e., −1 = c + i u∗i Aui for some c ≥ 0 and ui ∈ Rα ,
α×α
then with B := i ui u∗i ∈ SR0
P
we obtain a linear form
ℓ : SRα×α → R, A 7→ tr(AB)
satisfying ℓ(SRα×α
0 ) ⊆ R≥0 and ℓ({A(x) | x ∈ Rn }) = {−1 − c} ⊆ R≤−1 . So A is
strongly infeasible.
Lemma 3.1.2. Let A ∈ SR[x]α×α 1 be an infeasible linear pencil. Then the following
are equivalent:
(i) A is weakly infeasible;
(ii) SA+εIα 6= ∅ for all ε > 0.
Proof. Since all norms on a finite-dimensional vector space are equivalent, we can
without loss of generality use the operator norm on Rα×α .
Suppose that (i) holds and ε > 0 is given. Choose B ∈ SRα×α n
0 and x ∈ R with
kB − A(x)k < ε. Then A(x) + εIα 0, i.e., x ∈ SA+εIm .
Conversely, suppose that (ii) holds. To show that
dist {A(x) | x ∈ Rn }, SRα×α
0 } = 0,
kA(x) − Bk ≤ ε.
n
But this is easy: choose x ∈ R with A(x) + εIα 0, and set B := A(x) + εIα .
The following lemma is due to Bohnenblust [Bo48] (see also [Ba01, Theorem 4.2]
for an easier accessible reference). While Bohnenblust gave a non-trivial bound
on the number of terms that are really needed to test condition (i) below, we will
not need this improvement and therefore take the trivial bound α. Then the proof
becomes easy and we include it for the convenience of the reader.
Lemma 3.1.3 (Bohnenblust). For A1 , . . . ,An ∈ SRα×α the following are equiva-
lent:
Pα
(i) Whenever u1 , . . . ,uα ∈ Rα with i=1 u∗i Aj ui = 0 for all j ∈ {1, . . . ,n}, then
u1 = · · · = uα = 0;
6 IGOR KLEP AND MARKUS SCHWEIGHOFER
Proof. Assume that the conclusion is false. By Lemma 3.1.3, we find x0 ,x1 , . . . ,xn ∈
R such that
x0 A0 + x1 A1 + · · · + xn An ≻ 0.
Of course it is impossible that x0 > 0 since otherwise A xx10 , . . . , xxn0 ≻ 0. Also
To see that this corresponds (up to some minor technicalities) to the formulation
in the literature, just write the polynomial constraint ℓ − a = tr(AS) of the dual as
n + 1 linear equations by comparing coefficients.
A SUMS OF SQUARES DUAL FOR SDP AND INFEASIBLE LMI 7
For the induction step, we now suppose that n ∈ N and that we know already
how to find the required data for linear pencils in n − 1 variables. We distinguish
two cases and will use the induction hypothesis only in the second one.
Case 1. SA contains an interior point.
In this case, we set all ℓi and Si to zero so that (3) is trivially satisfied. Property
(4) can be fulfilled by Proposition 3.2.1.
Case 2. The interior of SA is empty.
In this case, we apply Proposition 3.3.1 to obtain 0 6= ℓ1 ∈ R[x]1 and a quadratic
sos-matrix S1 ∈ SR[x]α×α with
ℓ21 + tr(AS1 ) = 0. (5)
The case where ℓ1 is constant is trivial. In fact, in this case we can choose all
remaining data being zero since (ℓ1 , . . . ,ℓi ) = (ℓ1 ) = R[x] for all i ∈ {1, . . . ,n}.
From now on we therefore assume ℓ1 to be non-constant. But then the reader
easily checks that there is no harm carrying out an affine linear variable trans-
formation which allows us to assume ℓ1 = xn . We then apply the induction hy-
pothesis to the linear pencil A′ := A(x1 , . . . ,xn−1 ,0) and the linear polynomial
f ′ := f (x1 , . . . ,xn−1 ,0) in n − 1 variables to obtain ℓ2 , . . . ,ℓn ∈ R[x]1 , quadratic
sos-matrices S2 , . . . ,Sn ∈ SR[x]α×α2 , a matrix S ∈ SRα×α 0 and a constant c ≥ 0
such that
ℓ2i + tr(A′ Si ) ∈ (ℓ2 , . . . ,ℓi−1 ) for i ∈ {2, . . . ,n} and (6)
′ ′
f − c − tr(A S) ∈ (ℓ2 , . . . ,ℓn ). (7)
′ ′ ′
Noting that both f − f and tr(ASi ) − tr(A Si ) = tr((A − A )Si ) are contained in
the ideal (xn ) = (ℓ1 ), we see that (6) together with (5) implies (3). In the same
manner, (7) yields (4).
3.5. Constructing SDPs for sums of squares problems. The (coefficient tu-
ples of) sos-polynomials in R[x] of bounded degree form a projection of a spec-
trahedron. In other words, the condition of being (the coefficient tuple of) an
sos-polynomial in R[x] of bounded degree can be expressed with an LMI by means
of additional variables. This is the well-known Gram matrix method [Lau09, Ma08].
As noted by Kojima [Ko03] and nicely described by Hol and Scherer [SH06], the
Gram matrix method extends easily to sos-matrices (see Example (??) below).
3.6. Real radical computations. Let A ∈ SR[x]α×α 1 be a linear pencil and q ∈
R[x]2 a (quadratic) sos-polynomial such that −q = tr(AS) for some (quadratic)
sos-matrix S like in (2) above. In order to resolve the third issue mentioned in
Subsection 3.3, we would like to get our hands on (cubic) polynomials vanishing on
{q = 0}. That is, we want to implement the ideals appearing in (3) and (4) in an
SDP.
By the Real Nullstellensatz [BCR98, Ma08,
p PD01], each polynomial vanishing
on the real zero set {q = 0} of q lies in R (q). This gives us a strategy of how to
find the cubic polynomials vanishing on {q = 0}, cf. Proposition 3.6.1 and Lemma
3.6.2 below. The Real Nullstellensatz playsp only a motivating role for us; we only
use its trivial converse: Each element of R (q) vanishes on {q = 0}.
The central question is how to model the search for elements in the real radical
ideal using SDP. The key to this will be to represent polynomials by matrices as is
done in the Gram matrix method mentioned in Section 3.5. For this we introduce
some notation.
For each d ∈ N0 , let s(d) := dim R[x]d = d+n
n denote the number of monomials
#„ s(d)
of degree at most d in n variables and [x]d ∈ R[x] the column vector
#„ 2
∗
[x]d := 1 x1 x2 . . . xn x1 x1 x2 . . . . . . xdn
A SUMS OF SQUARES DUAL FOR SDP AND INFEASIBLE LMI 11
consisting of these monomials ordered first with respect to the degree and then
lexicographic.
The following proposition shows how to find elements
p of degree at most d + e
(represented by a matrix W ) in the real radical I := R (q) of the ideal generated by
#„ ∗ #„
a polynomial q ∈ R[x]2d (represented by a symmetric matrix U , i.e., q = [x]d U [x]d ).
We will later use it with d = 1 and e = 2 since q will be quadratic and we will be
interested in cubic polynomials in I. Note that
∗ I W
U W W ⇐⇒ 0
W∗ U
Proposition 3.6.1. Let d,e ∈ N0 , let I be a real radical ideal of R[x] and U ∈
#„ ∗ #„
SRs(d)×s(d) be such that [x]d U [x]d ∈ I. Suppose W ∈ Rs(e)×s(d) with U W ∗ W .
#„ ∗ #„
Then [x]e W [x]d ∈ I.
since I is an ideal.
The following lemma is a weak converse to Proposition 3.6.1. Its proof relies
heavily on the fact that only linear and quadratic polynomials are involved.
Lemma 3.6.2. Let ℓ1 , . . . ,ℓt ∈ R[x]1 , and q1 , . . . ,qt ∈ R[x]2 . Suppose that U ∈
SRs(1)×s(1) satisfies
#„ ∗ #„
[x]1 U [x]1 = ℓ21 + · · · + ℓ2t . (8)
Then there exists λ > 0 and W ∈ Rs(2)×s(1) satisfying λU W ∗ W and
#„ ∗ #„
[x]2 W [x]1 = ℓ1 q1 + · · · + ℓt qt .
Proof. Note that the U satisfying (8) is unique and hence positive semidefinite.
Suppose that at least one qi 6= 0 (otherwise take W = 0). Choose column vectors
#„
ci ∈ Rs(2) such that c∗i [x]2 = qi . Now let W ∈ Rs(2)×s(1) be the matrix defined by
#„ Pt
W [x]1 = i=1 ℓi ci , so that
t t t
#„ ∗ #„ X #„ ∗ X #„ X
[x]2 W [x]1 = ℓi [x]2 ci = ℓi c∗i [x]2 = ℓi qi .
i=1 i=1 i=1
Moreover, we get
t
#„ ∗ #„ #„ #„ X
[x]1 W ∗ W [x]1 = (W [x]1 )∗ W [x]1 = (ℓi ci )∗ (ℓj cj ),
i,j=1
12 IGOR KLEP AND MARKUS SCHWEIGHOFER
|α|≤d
1≤α1
where
1 −1
αX
!
X
p1 := aα xi1 ℓ′α
1
1 −1−i
xα αn
2 · · · xn ∈ R[x]d−1 .
2
|α|≤d i=0
1≤α1
Moreover, g ∈ (ℓ2 , . . . ,ℓt ) and therefore g = p2 ℓ2 + · · · + pt ℓt for some p2 , . . . ,pt ∈
R[x]d−1 by the induction hypothesis. Now
f = (f − g) + g = p1 ℓ1 + · · · + pt ℓt .
Theorem 3.7.3 (Sums of squares SDP duality). Let A ∈ SR[x]α×α 1 be a linear
pencil and f ∈ R[x]1 . Then
f ≥ 0 on SA
if and only if there exist quadratic sos-matrices S1 , . . . ,Sn ∈ SR[x]α×α 2 , matrices
U1 , . . . ,Un ∈ SRs(1)×s(1) , W1 , . . . ,Wn ∈ Rs(2)×s(1) , S ∈ SRα×α
0 and c ∈ R≥0 such
that
#„ ∗ #„ #„ ∗ #„
[x]1 Ui [x]1 + [x]2 Wi−1 [x]1 + tr(ASi ) = 0 (i ∈ {1, . . . ,n}), (9)
Ui Wi∗ Wi (i ∈ {1, . . . ,n}), (10)
#„ ∗ #„
f − c + [x]2 Wn [x]1 − tr(AS) = 0, (11)
where W0 := 0 ∈ Rs(2)×s(1) .
Proof. We first prove that existence of the above data implies f ≥ 0 on SA . All
we will use about the traces appearing in (9) and (11) is that they are polynomials
nonnegative on SA . Let I denote the real radical ideal of all polynomials vanishing
#„ ∗ #„
on SA . It is clear that (11) gives f ≥ 0 on SA if we show that [x]2 Wn [x]1 ∈ I. In
#„ ∗ #„
fact, we prove by induction that [x]2 Wi [x]1 ∈ I for all i ∈ {0, . . . ,n}.
The case i = 0 is trivial since W0 = 0 by definition. Let i ∈ {1, . . . ,n} be given
#„ ∗ #„ #„ ∗ #„
and suppose that [x]2 Wi−1 [x]1 ∈ I. Then (9) shows [x]1 Ui [x]1 ≤ 0 on SA . On the
#„ ∗ #„
other hand, (10) implies in particular Ui 0 and therefore [x]1 Ui [x]1 ≥ 0 on SA .
#„ ∗ #„ #„ ∗ #„
Combining both, [x]1 Ui [x]1 ∈ I. Now Proposition 3.6.1 implies [x]2 Wi [x]1 ∈ I by
(10). This ends the induction and shows f ≥ 0 on SA as claimed.
Conversely, suppose now that f ≥ 0 on SA . By Theorem 3.4.1 and Lemma 3.7.2,
we can choose ℓ1 , . . . ,ℓn ∈ R[x]1 , quadratic sos-matrices S1′ , . . . ,Sn′ ∈ SR[x]α×α ,
S ∈ SRα×α
0 and qij ∈ R[x]2 (1 ≤ j ≤ i ≤ n) such that
i−1
X
ℓ21 + · · · + ℓ2i + tr(ASi′ ) = q(i−1)j ℓj (i ∈ {1, . . . ,n}) and (12)
j=1
Xn
f − c − tr(AS) = qnj ℓj . (13)
j=1
14 IGOR KLEP AND MARKUS SCHWEIGHOFER
There are two little arguments involved in this: First, (3) can trivially be rewritten
as ℓ21 + · · · + ℓ2i + tr(ASi′ ) ∈ (ℓ1 , . . . ,ℓi−1 ) for i ∈ {1, . . . ,n}. Second, in Lemma 3.7.2
applied to ℓ1 , . . . ,ℓi−1 (i ∈ {1, . . . ,n + 1}) we might fall into case (b). But then we
may set ℓi = · · · = ℓn = 0 and Si′ = · · · = Sn′ = S = 0.
#„ ∗ #„
Now define Ui′ ∈ SRs(1)×s(1) by [x]1 Ui′ [x]1 = ℓ21 + · · · + ℓ2i for i ∈ {1, . . . ,n}.
Using Lemma 3.6.2, we can then choose λ > 0 and W1′ , . . . ,Wn′ ∈ Rs(2)×s(1) such
that
λUi′ Wi′∗ Wi′ (14)
#„ ∗ ′ #„ Pi
and [x]2 Wi [x]1 = − j=1 qij ℓj for i ∈ {1, . . . ,n}. Setting Wn := Wn′ , equation
(13) becomes (11). Moreover, (12) can be rewritten as
#„ ∗ #„ #„ ∗ ′ #„
[x]1 Ui′ [x]1 + [x]2 Wi−1 [x]1 + tr(ASi′ ) = 0 (i ∈ {1, . . . ,n}) (15)
To cope with the problem that λ might be larger than 1 in (14), we look for
λ1 , . . . ,λn ∈ R>0 such that Ui Wi∗ Wi for all i ∈ {1, . . . ,n} if we define Ui := λi Ui′
′
and Wi−1 := λi Wi−1 for all i ∈ {1, . . . ,n} (in particular W0 = W0′ = 0). With
this choice, the desired linear matrix inequality (10) is now equivalent to λi Ui′
λ2i+1 Wi′∗ Wi′ for i ∈ {1, . . . ,n − 1} and λn Un′ Wn′2 . Looking at (14), we therefore
see that any choice of the λi satisfying λi ≥ λλ2i+1 for i ∈ {1, . . . ,n − 1} and λn ≥ λ
ensures (10). Such a choice is clearly possible. Finally, equation (15) multiplied by
λi yields (9) by setting Si := λi Si′ for i ∈ {1, . . . ,n}.
We get the same interpretation for the certificates from Theorem 3.7.3.
Proposition 4.2.2. If A ∈ SR[x]α×α 1 is a linear pencil, f ∈ R[x]1 is a linear polyno-
mial, S1 , . . . ,Sn ∈ SR[x]2α×α are quadratic sos-matrices, U1 , . . . ,Un ∈ SRs(1)×s(1) ,
α×α
W1 , . . . ,Wn ∈ Rs(2)×s(1) , S ∈ SR0 and c ∈ R≥0 are such that (9), (10) and (11)
√
hold, then f ∈ CA + supp MA .
√
Proof. Set I := supp MA . It is clear that constraint (11) gives f ∈ CA + I if we
#„ #„ #„ #„
show that [x]2 Wn [x]1 ∈ I. In fact, we show by induction that [x]2 Wi [x]1 ∈ I for all
i ∈ {0, . . . ,n}.
The case i = 0 is trivial since W0 = 0 by definition. Let i ∈ {1, . . . ,n} be
#„ ∗ #„
given and suppose that we know already [x]2 Wi−1 [x]1 ∈ I. Then (9) shows
#„ ∗ #„
[x]1 Ui [x]1 ∈ I − MA . On the other hand (10) implies in particular Ui 0 and
#„ ∗ #„ #„ ∗ #„
R[x]2 ⊆ MA . But then [x]1 Ui [x]1 ∈ (I − MA ) ∩ MA ⊆ I
P
therefore [x]1 Ui [x]1 ∈
#„ ∗ #„
by (17). Now (10) yields [x]2 Wi [x]1 ∈ I by Proposition 3.6.1 since I is real radical
by Proposition 4.1.2. This ends the induction.
The following corollary is now a generalization of Proposition 3.2.1 working also
for low-dimensional SA (note that supp MA = (0) if SA has non-empty interior).
Corollary 4.2.3. Let A ∈ SR[x]α×α
1 be a linear pencil. Then
p
f ≥ 0 on SA ⇐⇒ f ∈ CA + supp MA
for all f ∈ R[x]1 .
Proof. Combine either Theorem 3.4.1 with Proposition 4.2.1, or Theorem 3.7.3 with
Proposition 4.2.2.
Corollary 4.2.4. Let A ∈ SR[x]α×α
1 be a linear pencil. Then
f ≥ 0 on SA ⇐⇒ ∀ε > 0 : f + ε ∈ MA
for all f ∈ R[x]1 .
Proof. To prove the non-trivial implication, let f ∈ R[x]1 with f ≥ 0 on SA be given.
It suffices to show f + ε ∈ MA for the special case ε = 1 (otherwise replace f by εf
and divide by ε). By Corollary 4.2.3, there exists g ∈ CA , p ∈ R[x] and k ∈ N such
that f = g+p and pk ∈ I := supp MA . Now f +ε = f +1 = g+(f −g)+1 = g+(p+1)
and it is enough to show that p + 1 ∈ MA . This will follow from the fact that the
image of p + 1 is a square in the quotient ring R[x]/I. Indeed, since the image of p
in R[x]/I is nilpotent (in fact the image of pk is zero), we can simply write down
a square root of this element using the finite Taylor expansion at 1 of the square
root function in 1 given by the binomial series:
k−1
!2
X 1
2 i
p+1≡ p mod I.
i=0
i
2
By Lemma 3.1.4, there is an u ∈ Rα r {0} with u∗ (Iα ⊗ A)u = 0. Replacing A
(k)
by Iα ⊗ A changes neither MA nor SA = ∅. Without loss of generality, we assume
therefore that there is u ∈ Rα r {0} with u∗ Au = 0. Writing A = (ℓij )1≤i,j≤α and
performing a linear coordinate change on Rα , we can moreover assume ℓ11 = 0.
Furthermore, without loss of generality, ℓ12 6= 0. Setting ℓ′ := 12 (−1 − ℓ22 ), v :=
[ℓ′ ℓ12 0 . . . 0]∗ and S := vv ∗ , we have
(1)
tr(AS) = v ∗ Av = 2ℓ′ ℓ212 + ℓ212 ℓ22 = ℓ212 (ℓ22 + 2ℓ′ ) = −ℓ212 ∈ MA . (18)
If ℓ12 ∈ R, we are done. Otherwise after possibly performing an affine linear change
of variables on Rn , we may assume ℓ12 = xn .
Now A′ := A(x1 , . . . ,xn−1 ,0) is an infeasible linear pencil in n − 1 variables. By
(2n−1 −1)
our induction hypothesis, −1 ∈ MA′ . In particular, there are
pi ∈ R[x]2n−1 −1 and vj ∈ R[x]α
2n−1 −1
satisfying X X
−1 = p2i + vj∗ A′ vj .
i j
vj∗ An vj ∈ R[x]2n −2 . Then
P
Let q := j
X X
−1 = 2 p2i + 2 vj∗ A′ vj + 1
i j
X X
=2 p2i +2 vj∗ Avj − 2qxn + 1 (19)
i j
X X
=2 p2i + 2 vj∗ Avj + (1 − qxn )2 + q 2 (−x2n ).
i j
(2n −1)
Since deg q ≤ 2n − 2, we have q 2 (−xn ) ∈ MA by (18). Taken together with
2 (2n −1) (2n −1)
(1 − qxn ) ∈ MA , (19) implies −1 ∈ MA .
Lemma 4.3.5. Let A be an infeasible linear pencil of size α. Then
(2α−1 −1)
−1 ∈ MA .
Proof. We prove this by induction on α. The statement is clear for α = 1. Given
n
X
A = A0 + xi Ai ∈ SR[x]α×α
1
i=1
of size α ≥ 2, we assume the statement has been established for all infeasible linear
(0)
pencils of size α − 1. If A is strongly infeasible, then −1 ∈ CA = MA by Lemma
3.1.1. So we may assume A is weakly infeasible.
Claim. There is an affine linear change of variables after which A assumes the
form
b∗
b
A= 0 ,
b A′
∗
where b0 ∈ R[x]1 , b = b1 · · · bα−1 ∈ R[x]α−1 , A′ is a linear pencil of size α−1,
1
and bj ∈ R[x]1 satisfy
(1)
− b2j ∈ MA for j = 0, . . . ,α − 1. (20)
Furthermore, b0 can be chosen to be either 0 or x1 .
Explanation. By Lemma 3.1.4, there is k ∈ N and u1 , . . . ,uk ∈ Rα r {0} with
Pk ∗
i=1 ui Aui = 0. We distinguish two cases.
Case 1. There is u ∈ Rα r {0} with u∗ Au = 0.
A SUMS OF SQUARES DUAL FOR SDP AND INFEASIBLE LMI 19
Since Case 1 does not apply, ℓ11 6= 0. Furthermore, since A is assumed to be weakly
infeasible, ℓ11 6∈ R. Hence after an affine linear change of variables on Rn , we can
assume ℓ11 = x1 . Thus
b∗
x
A= 1 ,
b A′
∗
where b = b1 · · · bα−1 ∈ R[x]α−1 and A′ is a linear pencil of size α − 1. Note
1
that
−4x21 = (1 − x1 )2 x1 + (1 + x1 )2 (−x1 )
(1)
shows that −x21 ∈ ML . Using this, one gets similarly as above that also each of
(1)
the entries bj of b satisfies −b2j ∈ MA . This proves our claim.
If one of the bj ∈ R r {0}, we are done by (20). Otherwise we consider two cases.
Case a. The linear system b0 (x) = 0, b(x) = 0 is infeasible.
Then we proceed as follows. There are γ0 , . . . ,γα−1 ∈ R satisfying
α−1
X
γj bj = 1. (21)
j=0
Introducing
X 0
qt := vi∗ At vi ∈ R[x]2α−1 −2 and wi := ∈ R[x]α
2α−2 −1
vi
i
we have
X
vi∗ A′′ vi + 1
−1 = 2s + 2
i
r r
X X X 1 √ 2
= 2s + 2 vi∗ A′ vi − 2qt xt + √ − rqt xt + 2qt xt + rqt2 (−x2t )
i t=1 t=1
r
r r
X X 1 √ 2 X
= 2s + 2 wi∗ Awi + √ − rqt xt + rqt2 (−x2t ).
i t=1
r t=1
(22)
(2α−1 −1) √ (2α−1 −1)
Combining qt2 (−x2t ) ∈ MA with ( √1r − rqt xt )2 ∈ MA , (22) implies
(2α−1 −1)
−1 ∈ MA .
Proof of Theorem 4.3.3. Immediate from Lemma 4.3.4 and Lemma 4.3.5.
Remark 4.3.6. With the aid of truncated quadratic modules associated to a linear
pencil A, we can introduce a hierarchy of infeasibility: A is called k-infeasible for
(k) (k−1)
k ∈ N0 , if −1 ∈ MA and −1 6∈ MA . By Lemma 3.1.1, A is strongly infeasible
if and only if it is 0-infeasible, and A is weakly infeasible if and only if it is k-
infeasible for some k ∈ N. Detecting k-infeasibility can be implemented as an SDP,
cf. Subsection 3.5.
In [HL06] Henrion and Lasserre extend Lasserre’s hierarchy [Las01] for optimiz-
ing over scalar polynomial inequalities to polynomial matrix inequalities (PMIs).
Motivated by problems of systems control theory, the authors of [HL06] develop
the primal-moment/dual-sos approach for (non-convex) PMIs, a particular case
of which are the (convex) LMIs treated here. Our Theorem 4.3.3 shows that for
infeasible LMIs, the SDP hierarchy described by Henrion and Lasserre in [HL06]
generates a certificate of infeasibility at a finite relaxation order.
(2) From Corollary 4.4.2 it is easy to see that the quadratic module in the sense
of rings with involution (see [KS10]) associated to a linear pencil A of size α
in the ring of β × β matrix polynomials is archimedean (in the sense of [KS10,
Subsection 3.1] or [HKM, Sections 6, 7]) if the spectrahedron SA defined by A
is bounded (cf. [HKM, Section 7]). Among other consequences, this implies a
suitable generalization of Corollary 4.4.4 for matrix polynomials positive definite
on the bounded spectrahedron SA (cf. [SH06, Corollary 1], [KS10, Theorem 3.7]
and [HKM, Theorem 7.1]).
4.5. An application to positive linear functionals. In this brief subsection we
explain how our results pertain to positive linear functionals.
Definition 4.5.1. Suppose R ⊆ SRα×α is a vector subspace, and let R0 :=
R ∩ SRα×α
0 . A linear functional f : R → R is called positive if f (R0 ) ⊆ R≥0 .
−2 = u∗ Au.
Hence A is 1-infeasible.
Example 4.6.2. Let
0 x1 0
A := x1 x2 1 .
0 1 x1
(1)
Then A is weakly infeasible and −1 6∈ MA .
Assume otherwise, and let
X
−1=s+ vj∗ Avj , (25)
j
where vj ∈ R[x]31 and s ∈ Σ2 ∩ R[x]2 . We shall carefully analyze the terms vj∗ Avj .
Write ∗
vj = q1j q2j q3j , and qij = aij + bij x1 + cij x2
with aij ,bij ,cij ∈ R. Then the x32 coefficient of vj∗ Avj equals c22j , so c2j = 0 for all
j. Next, by considering the x1 x22 terms, we deduce c3j = 0. Now the only terms
A SUMS OF SQUARES DUAL FOR SDP AND INFEASIBLE LMI 23
max a
s.t. S ∈ SR3×3
0 , a ∈ R
S1 ,S2 ∈ SR[x]3×3
2 quadratic sos-matrices (28)
3×3 6×3
U1 ,U2 ∈ SR , W1 ,W2 ∈ R
#„ ∗ #„
[x]1 U1 [x]1 + tr(AS1 ) = 0 (29)
U1 W1∗ W1 (30)
#„ ∗ #„ #„ ∗ #„
[x]1 U2 [x]1 + [x]2 W1 [x]1 + tr(AS2 ) = 0 (31)
U2 W2∗ W2 (32)
#„ ∗ #„
ℓ − a + [x]2 W2 [x]1 − tr(AS) = 0. (33)
Let us use the notation of (34) for the quadratic sos-matrix S1 , and consider the
left hand side of (29). Its constant coefficient is tr(A0 D) + (U1 )1,1 = 0. Since
A0 ,D,U1 0, this implies (U1 )1,j = (U1 )j,1 = 0 and D1,j = Dj,1 = 0. Next,
the x31 term of (29) is tr(A1 F1,1 ) = 0,h whence
i (F1,1 )2,j = (F1,1 )j,2 = 0. From
0 0 0
(35) it follows that Ξ1 is of the form ∗0∗ . Hence (E1 )2,2 = 0. Finally, by
∗0∗
considering the x21 term in (29), we obtain tr(A0 F1,1 ) + tr(A1 E1 h) + (Ui1 )2,2 = 0.
00 0
Since (E1 )2,2 = 0, A1 E1 = 0. As A0 ,F1,1 ,U1 0, we deduce U1 = 0 0 0 for some
00u
u ∈ R≥0 . In particular, from (30) we see the first two columns of W1 are 0. Hence
#„ ∗ #„
[x]2 W1 [x]1 ∈ x2 R[x].
Using this information on W1 , we can analyze (31) as in the previous paragraph,
and deduce that the first two columns of W2 are 0. Next, we turn to (33). All its
#„ ∗ #„
terms of degree ≥ 2 come from [x]2 W2 [x]1 , so (W2 )j,k = 0 for all (j,k) 6= (1,3).
This reduces (33) to the system of linear equations
−a = tr(A0 S)
0 = tr(A1 S)
1 + (W2 )1,3 = tr(A2 S).
It is instructive to compare this to (27) above. Again S2,j = Sj,2 = 0 for all j.
Since tr(A0 S) ≥ 0 and we are maximizing a, we set S1,1 = 0, yielding (W2 )1,3 = −1
and a = 0. It is now easy to see that this W2 and a can be extended to a feasible
(and thus optimal) point for the above sums of squares dual.
A SUMS OF SQUARES DUAL FOR SDP AND INFEASIBLE LMI 25
Finally, we give Ramana’s dual for (26); we refer to [Ra97, p. 142] for details.
max a
s.t. 0 = tr(Aj U1 ), j = 0,1,2
U1 W1 W1∗
0 = tr Aj (U2 + W1 ) , j = 0,1,2
(36)
U2 W2 W2∗
0 = tr Aj (U + W2 ) , j = 1,2
−a = tr A0 (U + W2 )
U 0.
The reader will have no problems verifying that the optimal value of (36) is 0.
Example 4.6.4 demonstrates that Ramana’s dual is generally smaller in size than
the sums of squares dual. However, the advantage of our sums of squares dual is
that it admits a nice real algebraic geometric interpretation, and naturally leads
itself to the Positivstellensätze we presented in Section 4.
Acknowledgments. The authors thank three anonymous referees for their de-
tailed reading and many helpful comments.
References
[Ba01] A. Barvinok: A remark on the rank of positive semidefinite matrices subject to affine
constraints, Discrete Comput. Geom. 25 (2001), no. 1, 23–31 5
[Ba02] A. Barvinok: A course in convexity, Graduate Studies in Mathematics 54, Amer.
Math. Soc., 2002 5, 6
[BCR98] J. Bochnak, M. Coste, M.F. Roy: Real algebraic geometry, Springer-Verlag, 1998 4, 10,
15
[BEFB94] S. Boyd, L. El Ghaoui, E. Feron, V. Balakrishnan: Linear matrix inequalities in system
and control theory, Studies in Applied Mathematics 15, SIAM, 1994 2, 3
[BL04] D.P. Blecher and C. Le Merdy: Operator algebras and their modules—an operator
space approach, London Mathematical Society Monographs. New Series 30, Oxford
University Press, 2004 20, 22
[Bo48] F. Bohnenblust: Joint positiveness of matrices, Technical report, California Institute
of Technology, 1948; available at
http://orion.uwaterloo.ca/~hwolkowi/henry/book/fronthandbk.d/Bohnenblust.pdf 5
[BV96] S. Boyd, L. Vandenberghe: Semidefinite programming, SIAM Rev. 38 (1996), no. 1,
49–95 1, 3
[BW81] J.M. Borwein, H. Wolkowicz: Facial reduction for a cone-convex programming problem,
J. Austral. Math. Soc. Ser. A 30 (1980/81), no. 3, 369–380 12
[Ce10] G. Chesi: LMI techniques for optimization over polynomials in control: a survey, IEEE
Trans. Automat. Control 55 (2010), no. 11, 2500–2510 2
[dK02] E. de Klerk: Aspects of semidefinite programming, Interior point algorithms and se-
lected applications. Applied Optimization 65, Kluwer Acad. Publ., 2002 2, 3
[Du07] B. Dumitrescu: Positive trigonometric polynomials and signal processing applications,
Springer-Verlag, 2007 2
[ER00] E.G. Effros, Z.-J. Ruan: Operator spaces, London Mathematical Society Monographs.
New Series 23, Oxford University Press, 2000 20, 22
[Fa02] J. Farkas: Theorie der einfachen Ungleichungen, J. reine angew. Math. 124 (1902),
1–27 21
[GLR82] I. Gohberg, P. Lancaster, L. Rodman: Matrix polynomials, Computer Science and
Applied Mathematics, Academic Press, 1982 3
[GN11] J. Gouveia, T. Netzer: Positive polynomials and projections of spectrahedra, SIAM J.
Optim. 21 (2011), no. 3, 960–976 20
26 IGOR KLEP AND MARKUS SCHWEIGHOFER
[Sm91] K. Schmüdgen: The K-moment problem for compact semi-algebraic sets, Math. Ann.
289 (1991), no. 2, 203–206 21
[Sr86] A. Schrijver: Theory of linear and integer programming, Wiley-Interscience Series in
Discrete Mathematics, John Wiley & Sons, 1986 1, 2
[St00] J. Sturm: Theory and algorithms of semidefinite programming, High performance op-
timization, 1–194, Appl. Optim. 33, Kluwer Acad. Publ., 2000 3, 4, 22
[To01] M.J. Todd: Semidefinite optimization, Acta Numer. 10 (2001), 515—560 1, 3
[TW] L. Tuncel, H. Wolkowicz, Strong Duality and Minimal Representations for Cone Opti-
mization, to appear in Comput. Optim. Appl.
http://orion.math.uwaterloo.ca/~hwolkowi/henry/reports/regclosegap.pdf 12
[WSV00] H. Wolkowicz, R. Saigal, L. Vandenberghe (editors): Handbook of semidefinite pro-
gramming. Theory, algorithms, and applications, Kluwer Acad. Publ., 2000 1, 3, 22
[Za] A. Zalar: A note on a matrix version of the Farkas lemma, Comm. Algebra 40 (2012)
3420–3429 17
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Reader’s guide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Notation and terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Matrix polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Linear pencils and spectrahedra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3. Sums of squares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.4. Radical ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Duality theory of semidefinite programming. . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1. Weakly feasible and weakly infeasible linear pencils . . . . . . . . . . . . . . . . 4
3.2. An algebraic glimpse at standard SDP duality . . . . . . . . . . . . . . . . . . . . 6
3.3. Certificates for low dimensionality of spectrahedra . . . . . . . . . . . . . . . . . 8
3.4. Linear polynomials positive on spectrahedra . . . . . . . . . . . . . . . . . . . . . . 9
3.5. Constructing SDPs for sums of squares problems . . . . . . . . . . . . . . . . . . 10
3.6. Real radical computations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.7. A new exact duality theory for SDP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4. Positivity of polynomials on spectrahedra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.1. Quadratic module associated to a linear pencil . . . . . . . . . . . . . . . . . . . . 14
4.2. Linear polynomials positive on spectrahedra – revisited . . . . . . . . . . . . . 15
4.3. Infeasibility of linear pencils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.4. Bounded spectrahedra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.5. An application to positive linear functionals . . . . . . . . . . . . . . . . . . . . . . 22
4.6. Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25