Poincar e Inequalities On Intervals Quadrature
Poincar e Inequalities On Intervals Quadrature
Franck Barthe
Institut de Mathématiques de Toulouse
Université Paul Sabatier, 31062 Toulouse Cedex 9, France
e-mail: franck.barthe@math.univ-toulouse.fr
and
Bertrand Iooss
Electricité de France R&D
6 quai Watier, Chatou, F–78401, France
&
Institut de Mathématiques de Toulouse
Université Paul Sabatier, 31062 Toulouse Cedex 9, France
e-mail: bertrand.iooss@edf.fr
Abstract: The development of global sensitivity analysis of numerical
model outputs has recently raised new issues on 1-dimensional Poincaré in-
equalities. Typically two kinds of sensitivity indices are linked by a Poincaré
type inequality, which provides upper bounds of the most interpretable in-
dex by using the other one, cheaper to compute. This allows performing
a low-cost screening of unessential variables. The efficiency of this screen-
ing then highly depends on the accuracy of the upper bounds in Poincaré
inequalities.
The novelty in the questions concern the wide range of probability dis-
tributions involved, which are often truncated on intervals. After providing
an overview of the existing knowledge and techniques, we add some theory
about Poincaré constants on intervals, with improvements for symmetric
intervals. Then we exploit the spectral interpretation for computing exact
value of Poincaré constants of any admissible distribution on a given inter-
val. We give semi-analytical results for some frequent distributions (trun-
cated exponential, triangular, truncated normal), and present a numerical
method in the general case.
Finally, an application is made to a hydrological problem, showing the
benefits of the new results in Poincaré inequalities to sensitivity analysis.
Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3082
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3082
1.2 Aim and plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3084
2 Background on Poincaré inequalities . . . . . . . . . . . . . . . . . . . 3085
2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3085
2.2 A general criterion . . . . . . . . . . . . . . . . . . . . . . . . . . 3088
2.3 Perturbation and transport . . . . . . . . . . . . . . . . . . . . . 3089
2.4 Spectral interpretation . . . . . . . . . . . . . . . . . . . . . . . . 3091
2.5 More techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . 3095
3 Poincaré inequalities on intervals . . . . . . . . . . . . . . . . . . . . . 3096
3.1 General results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3096
3.2 Improvements for symmetric settings . . . . . . . . . . . . . . . . 3097
4 Exact values of Poincaré constants . . . . . . . . . . . . . . . . . . . . 3100
4.1 Semi-analytical results . . . . . . . . . . . . . . . . . . . . . . . . 3100
4.2 A numerical method . . . . . . . . . . . . . . . . . . . . . . . . . 3108
4.3 Illustrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3109
5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3111
5.1 First study on a simplified flood model . . . . . . . . . . . . . . . 3112
5.2 Application on a 1D hydraulic model . . . . . . . . . . . . . . . . 3114
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3116
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3116
1. Introduction
1.1. Motivation
for all i ∈ I and all xI . Such conditions imply orthogonality, and lead to the
variance decomposition:
Var(f (x)) = Var(fI (xI )).
I⊆{1,...,d}
The “total Sobol index” SiT is then defined as the ratio of variance of f (x)
explained by xi (potentially with other variables):
T
Si = Var fI (xI ) /D, (1.2)
Ii
where D = Var(f (x)). Thus one can decide that xi is not influential if SiT is
less than (say) 5%.
Despite their nice interpretation, Sobol indices require numerous computa-
tions for their estimation. Then another global sensitivity index, called DGSM
(Derivative-based Global Sensitivity Measure), can advantageously be used as
a proxy [34, 23]. Defined by
2
∂f (x)
νi = dμ(x) (1.3)
∂xi
for i = 1, . . . , d, they are cheaper to compute, especially when the gradient of
f is available (e.g. as output of an evaluation of a complex numerical code).
As shown in [24], Sobol indices and DGSM are connected by a 1-dimensional
Poincaré-type inequality:
SiT ≤ C(μi )νi /D, (1.4)
where C(μi ) is a Poincaré constant for μi . Recall that μi satisfies a Poincaré
inequality if the energy of any centered
function is controlled by the energy of
its derivative: For all g satisfying gdμi = 0 there exists C(μi ) > 0 s.t.
g 2 dμi ≤ C(μi ) (g )2 dμi . (1.5)
Inequality 1.4 is easily obtained by applying the Poincaré inequality 1.5 to the
functions xi → Ii fI (xI ), which are centered for any choice of xI with i ∈ I
(condition 1.1), and integrating with respect to the other variables xj (j = i).
This allows performing a “low-cost” screening based on the upper bound of
Sobol indices in 1.4 instead of Sobol indices directly. Hence, one can decide that
xi is not influential if C(μi ) νDi is less than (say) 5%.
The efficiency of this low-cost screening strongly depends on the accuracy
of the Poincaré-type inequality 1.4, and this motivates the investigation of 1-
dimensional Poincaré inequalities. Notice that there is a large variety of prob-
ability distributions used in sensitivity analysis. They are often linked to prior
3084 O. Roustant, F. Barthe and B. Iooss
On the other hand, the DGSM νf of f is equal to 1 for each xi , and the corre-
ν
sponding upper bound is CP (N (0, 1)) Df = 12 , which is here exactly equal to the
total Sobol index of xi .
To conclude about motivations in sensitivity analysis, it is worth mentioning
that the idea of low-cost screening can be extended to higher-order interactions,
and also depends on the accuracy of 1-dimensional Poincaré inequalities [31]. It
allows screening out useless interactions and discovering additive structures in f .
Finally notice that measuring the influence of variables has useful applications
in other fields of probability. For instance, assuming a negligibility condition in
terms of variance ratios, [14] proves a central limit theorem for homogeneous
sums in Sobol-Hoeffding decomposition. Similarly, controlling the influence of
variables allows [25] to establish an invariance principle on the probability dis-
tribution of multilinear polynomials.
Our aim is to bridge the gap between industrial needs coming from sensitivity
analysis problems and the theory of Poincaré inequalities, by providing an ac-
cessible introduction to Poincaré inequalities for non-specialists, and by using
and developing the theory in order to deal with the specific situations motivated
by low-cost screening.
In Section 2, we present the general background on Poincaré inequalities,
together with the main techniques available to establish them (Muckenhoupt’s
criterion, perturbation and transportation methods, spectral methods). Most
of these techniques provide upper bounds on the Poincaré constant for large
classes of measures.
Poincaré inequalities on intervals – application to sensitivity analysis 3085
In this section, we provide a quick survey of the main simple techniques allowing
to derive Poincaré inequalities for probability measures on the real line. We often
make regularity assumptions on the measures. This allows to avoid technicalities,
without reducing the scope for realistic applications. Indeed, all the measures
we are interested in are supported on an open interval (a, b), and have a positive
continuous, piecewise continuously differentiable density. Moreover, when a is
finite, they are monotonic on a neighborhood of a (and similarly for b).
2.1. Definitions
Consider an open interval of the real line Ω = (a, b) with −∞ ≤ a < b ≤ +∞.
A locally integrable function f : Ω → R is weakly differentiable if there exists a
locally integrable function g : Ω → R such that for all functions φ of class C ∞
with compact support in Ω:
f (t)φ (t)dt = − g(t)φ(t)dt.
Ω Ω
Then g is a.e. uniquely determined (more precisely two functions with this prop-
erty coincide almost everywhere), it is called the weak derivative of f and de-
noted by f .
Let μ be a probability measure on Ω, and f : Ω → R be a Borel measurable
function. Recall that the variance of f for μ is defined as
Varμ (f ) = inf (f − a)2 dμ.
a∈R Ω
3086 O. Roustant, F. Barthe and B. Iooss
Obviously Varμ (f ) = +∞ if f 2 dμ = +∞. When f 2 dμ < +∞, it holds
2 2
Varμ (f ) = f dμ −
2
f dμ = f − f dμ dμ.
In this case, the smallest possible constant C above is denoted CP (μ), it is re-
ferred to as the Poincaré constant of the measure μ.
Observe that the above integrals are always defined, with values in [0, +∞].
Roughly speaking, a Poincaré inequality expresses in a quantitative way that
a function with a small weak derivative, measured in the sense of μ, has to be
close to a constant function again in the sense of μ.
Remark 1. Weakly differentiable functions are exactly the functions which ad-
mit (in their equivalence class for a.e. equality) a continuous version which
satisfies (see e.g. [2], [11])
y
∀x, y ∈ Ω, f (y) = f (x) + f (t) dt.
x
Such functions are also called (locally) absolutely continuous. Their variations
can be recovered by integrating their weak derivatives. It is therefore plain that
they provide a good setting for Poincaré inequalities. On the contrary, it is not
possible to work just with a.e. differentiable functions: for instance the famous
Cantor function (a.k.a. the Devil’s stairs) increases from 0 to 1 but is a.e.
differentiable with zero derivative. However, everywhere differentiable functions
with locally integrable derivative are weakly differentiable, see e.g. [32].
To be very precise, we should have denoted the Poincaré constant as CP (μ, Ω).
Indeed, we could also consider the probability measure μ on Ω as acting on any
larger open interval Ω . The restriction to Ω of weakly differentiable functions
on Ω are specific weakly differentiable function on Ω (they are continuous at
boundary points). Therefore, satisfying a Poincaré inequality on Ω is formally a
more demanding property. From now on, we will assume that Ω is the interior
of the convex hull of the support of μ, which is consistent with the notation
CP (μ).
Obviously, some measures cannot satisfy a Poincaré inequalities, for instance
the uniform measure on (0, 1) ∪ (2, 3) (one can choose a differentiable
function
f on (0, 3) which is equal to 0 on (0, 1) and to 1 on (2, 3). Then (f )2 dμ = 0).
Next let us present an equivalent definition of Poincaré inequalities, in a
more convenient analytic setting. Let L2 (μ) be the set of (equivalence classes
for a.e. equality of) measurable functions on Ω such that Ω f 2 dμ < +∞. We
Poincaré inequalities on intervals – application to sensitivity analysis 3087
1/2
denote . L2 (μ) , or simply . , its associated norm: f = Ω
f 2 dμ . The
first weighted Sobolev space Hμ1 (Ω) is defined by
where f is the weak derivative of f . It is well known that Hμ1 (Ω) is an Hilbert
space with the norm f 2H1 (Ω) = f 2 + f 2 . More generally, for an integer
μ
∀t ∈ Ω, 0 < m ≤ ρ(t) ≤ M.
Then if Leb denotes the Lebesgue measure on Ω, it holds that L2 (μ) = L2 (Leb)
and Hμ (Ω) = HLeb
(Ω) for all positive integers , and the norms on these spaces
are equivalent when μ is replaced by Leb. In other words, the weighted Sobolev
spaces are equivalent to the usual Sobolev spaces. This remark will allow us
to use several results which are available for Leb and a bounded Ω. Firstly,
Hμ1 (μ) then contains functions which are continuous on Ω and piecewise C 1 on
Ω. Secondly, the spectral theory of elliptic problems (see e.g. [2], Chap. 7) will
then be valid.
We now come back to the general case.
Definition 2. Let μ(dt) = ρ(t)dt be an absolutely continuous probability mea-
sure on an open interval Ω, such that ρ > 0 almost everywhere on Ω. We say
that μ admits a Poincaré inequality on Ω if there exists a constant C < +∞
such that for all f in Hμ1 (Ω) verifying Ω f dμ = 0, we have:
f 2 dμ ≤ C (f )2 dμ. (2.2)
Ω Ω
The best possible constant C is denoted CP (μ). If there exists fopt , a centered
function of Hμ1 (Ω), such that (2.2) is an equality for C = CP (μ), we say that
the inequality is saturated by fopt .
This definition means Varμ (f ) ≤ C(μ) Ω (f )2 dμ for weakly differentiable
functions f such that f and f are square integrable for μ. Hence Definition 1
is formally stronger as it involves general weakly differentiable functions. Nev-
ertheless, the Poincaré inequality of Definition 2 implies the one of Definition 1:
first it is enough to consider functions f with square integrable weak derivative
f , then one can apply the Poincaré inequality to certain truncations of f , which
belong to Hμ1 (Ω), in order to show that f is also necessarily square integrable
for μ.
Poincaré inequalities are often stated in terms of the ratio of energies of a
function and its derivative:
3088 O. Roustant, F. Barthe and B. Iooss
Definition 3. Let f in Hμ1 (Ω) with f 2 dμ > 0. The Rayleigh ratio of f is:
2
f 2 f dμ
J(f ) = 2
= Ω 2 . (2.3)
f Ω
f dμ
Thus μ admits a Poincaré inequality if and only if the Rayleigh ratio admits
lower bound over the subspace of centered functions Hμ (Ω) = {f ∈
1,c
a positive
Hμ (Ω), Ω f dμ = 0}. In that case,
1
−1
CP (μ) = inf
1,c
J(f ) = sup J(f )−1 .
f ∈Hμ (Ω)−{0} 1,c
f ∈Hμ (Ω)−{0}
Proof. The lower bound is obtained for f (x) = x − Eμ . The equality case is
well-known (see e.g. [3], [5]).
The class of probability measures that admit a Poincaré inequality has been
completely characterized in the work of Muckenhoupt [26]. See also [8] for re-
finements.
1
max(A− , A+ ) ≤ CP (μ) ≤ 4 max(A− , A+ ). (2.4)
2
Given a measure verifying a Poincaré inequality, one may try to deform it into
another measure still having a finite Poincaré constant. Many such results exist
in the literature. Here we mention two fundamental ones: the bounded pertur-
bation principle, and the Lipschitz transportation principle. They hold in very
general settings, but for the purpose of this article, it is enough to state them
on R.
Lemma 1. Let μ be a probability measure on Ω ⊂ R. Let ψ : Ω → R be
a bounded measurable function and let μ̃ be the probability measure given by
dμ̃ = eψ dμ/Z where Z is the normalizing constant. Then
CP (μ̃) ≤ esup ψ−inf ψ CP (μ).
This fact is easily proved at the level of Poincaré inequalities by using Varμ̃ =
inf a∈R (f −a)2 dμ̃ and applying obvious bounds on ψ. See e.g. Proposition 4.2.7
of [5] and related comments.
Lemma 2. Let μ be an absolutely continuous probability measure on an open
interval Ω ⊂ R. Let T : Ω → R be a Lipschitz map, (i.e. verifying that there
exists L ∈ R such that for all x, y ∈ Ω, |T (x) − T (y)| ≤ L|x − y|). Denote by T μ
the image measure of μ by T , i.e. the measure on R such that for every Borel
set B, T μ(B) = μ(T −1 (B)). Assume that T μ is absolutely continuous. Then
CP (T μ) ≤ T 2
Lip CP (μ),
In the case of probability measures μ, ν on the real line, and when μ has
no atoms, a natural map which pushes μ forward to ν is the monotonic map
T := Fν−1 ◦ Fμ (here Fν−1 stands for the generalized left inverse). It remains to
estimate the Lipschitz norm of T .
Lemma 3. Let μ and ν be probability measures on R. Assume that μ(dt) =
ρμ (t)dt where ρμ is positive and continuous on (aμ , bμ ) (−∞ ≤ aμ , bμ ≤ +∞)
and vanishes outside. Let us make the same structural assumption for ν. Then
T := Fν−1 ◦ Fμ is well defined and differentiable on (aμ , bμ ), with T = ρμ /ρν ◦
Fν−1 ◦ Fμ . Consequently its Lipschitz norm is
ρμ ρμ ◦ Fμ−1 ρμ ◦ Fμ−1 ◦ Fν
T Lip = sup −1 = sup −1 = sup .
(aμ ,bμ ) ρν ◦ Fν ◦ Fμ (0,1) ρν ◦ Fν (aν ,bν ) ρν
The previous two lemmas can be applied when μ(dx) = exp(−|x|)dx/2, the
double exponential measure. In this case ρν ◦Fν−1 (t) = min(t, 1−t) and CP (μ) =
4. This is how Bobkov and Houdré [10] established the following estimate:
2
min(Fν (x), 1 − Fν (x))
CP (ν) ≤ 4 sup . (2.5)
x∈(aν ,bν ) ρν (x)
These authors also deduced from this approach that for log-concave probability
measures ν on R, with median m, CP (ν) ≤ 1/ρν (m)2 .
Actually, we can improve on (2.5) by choosing the logistic measure μ(dx) =
ex
(1+ex )2 dx instead of the double exponential measure in the proof. Indeed, in
this case ρμ ◦ Fμ−1 (t) = t(1 − t) is smaller than min(t, 1 − t), while the Poincaré
constant is the same CP (μ) = 4, as shown in [6]. Consequently
Corollary 1. Let ν be a probability measure on (a, b) ⊂ R, with a positive
continuous density on (a, b). Then
2
Fν (x)(1 − Fν (x))
CP (ν) ≤ 4 sup .
x∈(a,b) ρν (x)
Remark 2. If we apply the previous two lemmas with μ being the uniform
probability measure on [− 12 ; 12 ], for which CP (μ) = π −2 (see Table 1 below), we
get that
2
1
CP (ν) ≤ .
π inf x∈(a,b) ρν (x)
This inequality is meaningful only when the density of ν is bounded from below by
a positive constant (which implies in particular that ν is compactly supported).
This is related to the fact (put forward in the introduction) that rewriting sen-
sitivity analysis questions in terms of uniform variables may not provide good
results for our purposes. The present section shows that rewriting the problem
in terms of logistic variables is more effective in general.
Poincaré inequalities on intervals – application to sensitivity analysis 3091
When μ has a continuous density that does not vanish on a compact interval
[a, b], then μ admits a Poincaré constant on [a, b], as follows from the Muck-
enhoupt condition (Theorem 1) or from the bounded perturbation principle
(Lemma 1). Furthermore, modulo a small additional regularity condition, the
Poincaré inequality is saturated and the Poincaré constant is related to a spec-
tral problem, as detailed now.
Theorem 2. Let Ω = (a, b) be a bounded open interval of the real line, and
assume that μ(dt) = ρ(t)dt = e−V (t) dt where V is continuous and piecewise C 1
on Ω = [a, b]. Consider the three following problems:
2
(P1) Find f ∈ Hμ1 (Ω) s.t. J(f ) = ff 2 is minimum under f dμ = 0.
(P2) Find f ∈ Hμ (Ω) s.t. f , g = λf, g ∀g ∈ Hμ (Ω).
1 1
∀t ∈ Ω, 0 < m ≤ ρ(t) ≤ M.
3092 O. Roustant, F. Barthe and B. Iooss
Then, as mentioned in § 2.1, L2 (μ) and all weighted Sobolev spaces Hμ (Ω) are
topologically unchanged when we replace μ by the Lebesgue measure on Ω. This
allows using the spectral theory of elliptic problems (see e.g. [2], Chap. 7) on
usual Sobolev spaces with bounded Ω.
More precisely, let us consider problem (P2), and denote H = L2 (μ) and
V = Hμ1 (Ω). Then the injection V ⊂ H is compact, and V is dense in H. (P2)
can be written as an eigenvalue problem of the form:
a(f, g) = λf, g ∀g ∈ V, (2.7)
where a(f, g) = f , g and ., . denotes the scalar product on H. In order to
apply spectral theory, a should be coercive with respect to V, i.e. ∃C > 0 s.t.
a(f, f ) ≥ Cf, f H1μ (Ω) for all f , which is not true here. A convenient way to
overcome this issue is to consider the equivalent shifted problem (see [13], §8.2.):
α(f, g) = (λ + 1)f, g ∀g ∈ V, (2.8)
where α(f, g) = f , g + f, g is the usual scalar product on V. It is coercive
on V. Then we can apply Theorem 7.3.2 in [2]: The possible eigenvalues form
an increasing sequence (λk + 1)k≥0 of non-negative values that tends to infinity
and the eigenvectors (uk )k≥0 form a Hilbert basis of L2 (μ). Further, from (2.7)
with f = g, we have λk ≥ 0. Now remark that 0 is an eigenvalue and the
corresponding eigenspace is spanned by the constant function u0 = 1. Indeed,
b
taking g = f in (2.2) leads to a f (x)2 e−V (x) dx = 0 and f is a constant function.
Now let us prove the equivalence between (P2) and (P3). We restrict the
presentation to λ > 0 since the case λ = 0 is direct, using a Sturm-Liouville
form of (P3), i.e. (f V ) = −λf . Formally the link comes from an integration
by part of the left hand side of (P2):
b b
f g ρ = f (b)ρ(b)g(b) − f (a)ρ(a)g(a) − (f ρ) g. (2.9)
a a
b
Since this must be equal to λ a f gρ for all g ∈ Hμ1 (Ω), we should have f (a) =
f (b) = 0 and (f ρ) = −λf ρ which is problem (P3). Actually this method, read
in the reverse sense, shows that (P3) implies (P2). Indeed if f ∈ Hμ2 (Ω) then
f ρ ∈ Hμ1 (Ω) and the integration by part is valid. However, to see that (P2)
implies (P3), an argument of regularity must be used since f is only assumed
to belong to Hμ1 (Ω). This is achieved by proving that, μ-almost surely,
b
λ
f (x) = f (t)ρ(t)dt. (2.10)
ρ(x) x
Indeed, this implies that f ρ is C 1 on [a, b] and (2.9) is valid. Furthermore, (2.10)
also implies f (a) = f (b) = 0 and this gives (P3). Let us now come xback to the
proof of (2.10). Following [8], Lemma 4.3., by using g(x) = g(a) + a g (t)dt and
Fubini’s theorem, it holds:
b b b b
f (x)g(x)ρ(x)dx = g(a) f (x)ρ(x)dx + f (t)ρ(t)dt g (x)dx.
a a a x
Poincaré inequalities on intervals – application to sensitivity analysis 3093
b
Recall that for λ > 0, a
f ρ = 0. Thus (P2) can be rewritten:
b b
b
f (x)g (x)ρ(x)dx = λ f (t)ρ(t)dt g (x)dx.
a a x
f 2 +∞ 2
α(f, f ) k=1 λk fk
J(f ) = 2
= −1= +∞ 2 ≥ λ1 ,
f f 2 k=1 fk
= |f | and thus (g )2 dμ = (f )2 dμ. Secondly, using the
Firstly, we have g
J(f + g, λ, β) = J(f, λ, β) +
b
∂L ∂L
g(t) (t, f (t), f (t)) + g (t) (t, f (t), f (t)) dt + O( 2 ).
a ∂f ∂f
The first order condition leads to cancel the integral, which gives here:
b b
(−λf g + f g ) dμ = β gdμ.
a a
b
Now, with g = 1 and using a
f dμ = 0, we get β = 0. This gives (P2).
Remark 3. Another solution to make coercive the bilinear form a(f, g) =
f , g is to project onto the space of centered functions Hμ1,c (Ω), as done in
[13]. Indeed, on that space, coercivity is equivalent to the existence of a Poincaré
constant, which can be proved independently by the bounded perturbation prin-
ciple as mentioned above. However, this seems a less convenient setting for the
numerical method developed in Section 4.2 since the ‘hat’ functions used do not
belong to Hμ1,c (Ω).
Definition 4 (Spectral gap). The smallest strictly positive eigenvalue λ1 of
Lf = f − V f with Neumann boundary conditions f (a) = f (b) = 0 is called
Neumann spectral gap, or simply spectral gap, and denoted by λ(μ).
We conclude this section by a proposition showing that, under regularity
conditions on V , the Neumann problem (P3) can be written as a Dirichlet
problem.
Proposition 2. If V is of class C 2 , then (P3) is equivalent to:
(P4) Find h ∈ Hμ2 (Ω) s.t. h − V h − V h = −λh and h(a) = h(b) = 0.
Poincaré inequalities on intervals – application to sensitivity analysis 3095
This section gathers tools for the study of the Poincaré constant of the restric-
tions of a given measure to subintervals.
The same applies if β < β. This allows to build a weakly differentiable function
f˜ on Ω which coincides with f on I and has zero derivative outside I. The result
easily follows:
dμ dμ
Varμ|I (f ) = inf (f − a)2 ≤ inf (f˜ − a)2
a μ(I) a μ(I)
I
Ω
2 dμ dμ
≤ CP (μ) (f )˜ = CP (μ) (f )2
μ(I) μ(I)
Ω I
= CP (μ) (f )2 dμ|I .
I
Poincaré inequalities on intervals – application to sensitivity analysis 3097
The previous result helps proving a continuity type property with respect to
the support:
Lemma 5. Consider again the framework of Lemma 4, and let I be a family
and increasing subintervals such that I ↑ Ω when → 0. Typically, I = (a +
, b − ) for Ω = (a, b), I = (a, a + 1/ ) for Ω = (a, +∞) and so on. Then the
Poincaré constant on Ω is the limit of the Poincaré constant on subintervals:
CP (μ) = lim CP (μ|I ).
→0
Since it is true for all f ∈ Hμ1 (Ω), we get CP (μ) ≤ lim CP (μ|I ).
→0
ρμ (x) ρμ (x)
T (x) = = μ(I) ≤ μ(I),
ρν (T (x)) ρμ (T (x))
Proof. Let (−b, b) denote the interior of the support of μ. By hypothesis, f (0)
is well defined and by the variational definition of the variance
2 0 2 b
(f )2 dμ (f ) dμ −b
(f ) dμ + 0 (f )2 dμ
≥ = 0 b .
Varμ (f ) (f − f (0))2 dμ (f − f (0))2 dμ + 0 (f − f (0))2 dμ
−b
Using the inequality u+vx+y ≥ min( x , y ), which is valid for non-negative numbers
u v
u, v, x, y such that x + y > 0 (and with the convention that v/0 = +∞), we get
that 0 2 b 2
2
(f ) dμ −b
(f ) dμ (f ) dμ
≥ min 0 , b 0 .
Varμ (f ) (f − f (0))2 dμ
−b (f − f (0))2 dμ
0
In order to conclude, we consider the odd functions g, h defined on (−b, b), such
that for all x ∈ [0, b), g(x) = f (x) − f (0), and for all x ∈ (−b, 0], h(x) =
f (x) − f (0). By symmetry of μ,
0 b
−b
(f )2 dμ (h )2 dμ (f )2 dμ (g )2 dμ
0 = , and b
0
= ·
−b
(f − f (0))2 dμ Varμ (h)
0
(f − f (0))2 dμ Varμ (g)
Poincaré inequalities on intervals – application to sensitivity analysis 3099
Next, we derive from the above proposition a perturbation principle for the
Neumann spectral gap. It should be compared with the “bounded perturbation
principle” that we recalled in the previous section.
Proposition 3. Let dμ(x) = 1(−b,b) e−V (x) dx be an even probability measure,
with continuous potential V . Let ϕ : (−b, b) → R+ be an even function. Assume
that ϕ is non-increasing on [0, b) and that μ̃ defined by dμ̃ = ϕ dμ is a probability
measure. Then
CP (μ̃) ≤ CP (μ).
is true for all t ≥ 0. But, by definition, pointwise, ϕ = R+ 1It . Thus integrating
the latter inequality over t yields
f 2 ϕ dμ ≤ CP (μ) (f )2 ϕ dμ.
This can be rewritten as Varμ̃ f ≤ CP (μ) (f )2 dμ̃.
This section presents some examples where the spectral gap is obtained in a
closed-form expression or by numerically searching a zero of a function ex-
pressed analytically. They are summarized in Table 1. The most famous one
is about the uniform distribution on [a, b], corresponding to a spectral prob-
lem for the Laplacian operator, f = −λf . Another well-known result is about
the truncated normal, for the intervals formed by consecutive zeros of Hermite
polynomials. After recalling this result, we show how to extend it for a general
interval and develop the case of truncated (double) exponential and triangular
distributions. Without loss of generality, we will consider only one value for the
parameters, remarking that the general result is obtained by rescaling (Lemma
2 applied to T and T −1 when T is an affine function).
equation u − V u = −λu. Then the spectral gap on [a, b] is the first zero of
fλ (a) gλ (a)
d(λ) = det = fλ (a)gλ (b) − fλ (b)gλ (a).
fλ (b) gλ (b)
Furthermore if V is even and a = −b, the spectral gap is the first zero of λ →
gλ (b).
Remark 4. This result is immediately adapted to the equivalent spectral problem
with Dirichlet conditions (P4), u −V u −V u = −λu. In that case, with similar
notations, the spectral gap is the first zero of
fλ (a) gλ (a)
d(λ) = det = fλ (a)gλ (b) − fλ (b)gλ (a)
fλ (b) gλ (b)
f − f = −λf.
Now the Neumann conditions f (a) = f (b) = 0 provide the linear system
M (A B)T = 0, with:
1 − a/2 −1/2
M= .
1 + b/2 −1/2 − b/2
Triangular distribution
1
y (x) + y (x) + λy(x) = 0,
x
with initial conditions y(1) = y ( ) = 0. This is a Bessel-type differential equa-
tion, which solution has the form (see e.g. [1], Chapter 9):
√ √
y(x) = AJ0 x λ + BY0 x λ ,
where J0 and Y0 are the Bessel functions of first and second type. J0 is an even
function of class C ∞ on R, whereas Y0 is C ∞ on R∗+ with infinite limit at 0+ .
Poincaré inequalities on intervals – application to sensitivity analysis 3105
By using the same argument as in Proposition 4, λ(μ|I ) is then the first zero of
⎛ √ √ ⎞
J0 λ Y0 λ
˜ = det ⎝ √
d(λ) √ ⎠ .
J0 λ Y0 λ
Now J (0) = 0, but Y0 is not defined at 0 and xY0 (x) → π2 as x tends to zero
([1], Chapter 9, Eq. 9.1.9 and 9.1.28). So it is convenient to consider:
⎛ √ √ ⎞
J0 λ Y0 λ
˜ = det ⎝
d(λ, ) = d(λ) √ √ ⎠ .
J0 λ Y0 λ
The expression of f has been deduced from y(x) = J0 (r1 x) by applying the
transformation x = 1 − t, followed by a symmetrization. By construction, f and
f are continuous at 0. Thus f is C 1 on Ω and piecewise C 2 on Ω. Furthermore,
it verifies the following equation on [0, 1)
f − V f + r12 f = 0, (4.1)
with boundary conditions f (0) = f (1) = 0. Since μ is even and f is odd, the
1 1
Rayleigh quotient of f is simply 0 (f )2 ρ/ 0 f 2 ρ. In order to calculate this
quotient, we multiply Eq. 4.1 by f ρ and integrate by parts on [0, 1). However,
since V = − ln(ρ) has a singularity at 1, we rather fix > 0 and work on
I = [0, 1 − ]. Observe that f and f ρ are C 1 on I . Starting from Eq. 4.1 and
then integrating by parts, we obtain:
1− 1− 1− 1−
f 2 ρ−[f f ρ]0 .
1−
2
r1 f ρ=−
2
(f −V f )ρf = − (ρf ) f =
0 0 0 0
Let us consider the standard Gaussian distribution dγ(x) = √1 e−V (x) with
2π
2
V (x) = x2 . The spectral problem (P3) associated to Poincaré inequality is
relative to the operator Lf = f − V f :
Lf = f − xf .
The spectral gap is known on every interval formed by successive roots of Her-
mite polynomials (see e.g. [13]), as detailed in Prop. 7. Recall that Hermite
polynomials Hn are the orthonormal basis of polynomials for L2 (dγ) with uni-
∞
tary highest coefficient: −∞ Hn (x)Hm (x)dγ(x) = δn,m . The first ones are:
and the following can be obtained by the recursive relation: Hn+1 (x) = xHn (x)−
Hn (x). They are also shown to verify the identity Hn+1
(x) = (n + 1)Hn (x)
which proves, together with the recursion formula, that Hn is solution of the
differential equation Lf = −nf .
Proposition 7 (Spectral gap of the normal distribution, specific case). Let
n ≥ 1, x1 < · · · < xn the roots of the n-th Hermite polynomial Hn . Let I
be one of the intervals: (−∞, x1 ], [x1 , x2 ], . . . , [xn−1 , xn ], [xn , +∞). Then the
(Neumann) spectral gap on I is n + 1.
Proof. On each I, the function Hn is of constant sign and vanishes on the
boundary. Since Hn+1 = (n+1)Hn then inside I, Hn+1 is strictly monotonic and
Hn+1 = 0 on the boundary. Furthermore, Hn+1 is solution of Lf = −(n + 1)f .
Hence, by Corollary 2, n + 1 is the spectral gap.
In the general case, the spectral gap of the truncated normal distribution is
related to hypergeometric series, and more specifically to confluent hypergeo-
metric series or Kummer’s function ([1], Section 13). The Kummer’s function
Ma1 ,b1 (z) = 1 F1 (a1 ; b1 ; z) is a series p≥0 xp with x0 = 1 satisfying
xp+1 (p + a1 )z
= . (4.2)
xp (p + b1 )(p + 1)
Denote N|[a,b] the standard normal distribution N (0, 1) truncated on [a, b].
Then its spectral gap λ(N|[a,b] ) is the first zero of the function
d(.) = bh0 (., a)h1 (., b) − ah0 (., b)h1 (., a).
Furthermore:
• If there exists λ such that a, b are two successive zeros of h0 (λ, .) or two
successive zeros of h1 (λ, .) then λ(N|[a,b] ) = λ.
• If a = −b then λ(N|[a,b] ) is the first zero of h0 (., a).
• If a = 0 (resp. b = 0), then λ(N|[a,b] ) is the first zero of h1 (., b) (resp.
h1 (., a)).
Proof. Here it is easier to consider the spectral problem with Dirichlet conditions
(Problem (P4), see Prop. 2). It is given by:
Let us split f into even an odd part, f (t) = f0 (t)+tf1 (t) with f0 (t) = p≥0 up t2p
2p
and f1 (t) = p≥0 vp t , where up = c2p and vp = c2p+1 . Then by defini-
t2
tion of Kummer series (4.2), we recognize f0 (t) = c0 M 1−λ ; 1 2 and f1 (t) =
2 2 2
Fig 1. Basis of finite elements P1 on [0, 1]. The gi ’s are hat functions for i = 1, . . . , n − 1,
truncated at the boundaries (i = 0 and i = n).
In this section we present a numerical method to estimate the spectral gap. The
technique is well-known in the literature of elliptic problems (see e.g. [30] or
[2]), and we adapt it to our case. We consider the framework and notations of
Section 2.4. The spectral problem (P3),
where α(f, g) = f , g +f, g is coercive on Hμ1 (Ω), and to show its equivalence
to (P3). This was done in Theorem 2.
The second step is to observe that (4.5) can be solved algebraically in a
finite-dimensional space. Denote Gh a finite-dimensional space of Hμ1 (Ω), where
h > 0 is a discretization parameter, and let (gi )0≤i≤n be a basis of Gh . Typically,
Gh = P1 , the space of Lagrange finite elements, composed of piecewise linear
functions on [a, b]. A basis is formed by the ‘hat’ functions gi represented in
Figure 1. A solution in Gh of (4.5) is given by
n
fh = fh,i gi
i=0
Kh fh = (λ + 1)Mh fh , (4.6)
where fh is the vector of (fh,i )0≤i≤n , and Kh and Mh are called respectively
‘rigidity matrix’ and ‘mass matrix’, defined by:
Notice that Mh and Kh are symmetric and positive definite. Then the problem
can be reexpressed in a standard form by using the Choleski decomposition of
h =
Mh = Lh LTh , where Lh is a lower triangular matrix. Indeed, denoting K
−1 T −1 T
Lh Kh (Lh ) and fh = L fh , (4.6) is written:
h fh = (λ + 1)fh .
K
Thus λ and fh are obtained by performing the eigen decomposition of the sym-
metric matrix K h . Finally fh is deduced from the relation fh = (LT )−1 fh .
h
The last step is to observe that the solutions of the finite-dimensional weak
formulation (4.6) converge to the solutions of (P2), and equivalently (P3), when
h → 0, as stated in [30] (Theorem 6.5.1.), [2] (§ 6.2.2.), or [4] (§ 8, Eq. 8.43 and
§ 10.1.2. Eq. 10.22), with a speed of convergence linked to the regularity of the
solutions. These results are expressed for the Lebesgue measure on [a, b], but are
still valid under the assumptions of Theorem 2 since for all integer , Hμ (Ω) is
then equal to HLeb
(Ω) with equivalent norms (see the proof of Theorem 2). In
our situation, the space spanned by the first two eigenfunctions lies in Hμ+1 (Ω),
with ≥ 1 (see Theorem 2). Hence the smallest strictly positive eigenvalue of
(4.6), i.e. the second one, converges to the spectral gap at the speed O(h2 ), and
– since it is a simple eigenvalue – the corresponding eigenvector converges to a
function saturating the Poincaré inequality at the speed O(h ).
Remark 6 (Numerical improvements). We briefly mention two well-known
improvements for finite elements algorithms, which apply to our case (see e.g.
[2]). Firstly, it is possible to replace the computations of integrals in Mh by
quadrature formulas (‘mass lumping’ speed-up technique). This allows obtain-
ing a diagonal matrix and Lh is simply the square root of its diagonal terms.
The convergence result is still valid. Secondly, if the solution is regular enough
(typically f ∈ Hμ3 (Ω)), more regular basis functions can be used (such as P2
finite elements). The speed of convergence is higher, at the price of a higher
computational cost.
4.3. Illustrations
Fig 2. Poincaré constant of the double exponential distribution μ truncated on [a, b]. Left:
π 2 −1
Contour plot as a function of Fμ (a) and 1 − Fμ (b), with the upper bound ( 14 + ( b−a ) ) (red
dashed lines). Right: Representation in the symmetric case for I = [−b, b], with two upper
bounds and the lower bound given by the variance.
Similar plots have been produced for the normal distribution N (0, 1), truncated
on I = [a, b], gathered in Figure 3. In the symmetric case, the upper bound
obtained by transport μ(I)2 (Lemma 6) looks globally accurate, especially for
strong truncations. The lower bound given by the variance of the truncated
distribution σI2 = 1 − 2 bφ(b)
μ(I) is even sharper. In order to visualize the link to
Hermite polynomials, we have added the points corresponding to intervals [−b, b]
where −b, b are two successive zeros of Hermite polynomials of degree 2n, up to
degree 100. Recall that in that case CP (μ) = 1/(2n + 1).
Poincaré inequalities on intervals – application to sensitivity analysis 3111
In the general case, we have added the intervals [rn,i−1 , rn,i ] corresponding to
consecutive zeros of Hermite polynomials of degree n (up to n = 100), associated
to Poincaré constants equal to 1/(n + 1). We can observe that the whole set of
Hermite consecutive zeros poorly fill the space of possible intervals, except for
very small symmetric intervals (located around the diagonal x + y = 1). The
Poincaré constants obtained with their extensions, namely Kummer’s functions,
thus provide a useful improvement in many cases.
Fig 3. Poincaré constant of the normal distribution N (0, 1) truncated on [a, b]. Left: Contour
plot as a function of Φ(a) and 1 − Φ(b). Right: Representation in the symmetric case for
I = [−b, b], with the upper bound μ(I)2 and the lower bound given by the variance. Colored
points correspond to Hermite polynomials: For a given degree, the same color is used.
5. Applications
where the output variable S is the maximal annual overflow (in meters), H is
the maximal annual height of the river (in meters), Q is the river flowrate, Ks
the Strickler’s friction coefficient, Zm and Zv the upstream and downstream
riverbed levels (w.r.t. a fixed reference), L and B are the length and width of
the water section, Hd is the dyke height and Cb is the bank level. Table 2 gives
the probability distributions of the model input variables which are supposed to
be independent. Notice that, as in the rest of this paper, we have used here the
notation N (μ, σ 2 ) where σ is the standard deviation, contrarily to [24] which
uses N (μ, σ). Hence the standard deviation of Ks is equal to 8.
Table 3 gives the Poincaré constants and two upper bounds for the different
probability distributions used in this test case (the uniform distribution is not
represented as the result is well known). For the sake of interpretation, we have
considered scaled distributions. Indeed, a simple change of variable (which can
also be viewed as a linear transport) shows that:
a−μ b−μ
CP N (μ, σ )|[a, b] = σ CP N (0, 1)
2 2
, ,
σ σ
Poincaré inequalities on intervals – application to sensitivity analysis 3113
Figure 4 gives the final global sensitivity analysis results for this test case. It
is based on the numerical values obtained in [24]. In particular, the DGSM νi
(i = 1, . . . , 8) have been computed via a sample of the output derivatives com-
ing from a low discrepancy sequence of the inputs of size 10 000. The previous
results of [24] (given here in Fig. 4, left) have shown that the DGSM-based up-
per bounds can be used for a screening purpose (i.e. identifying non influential
inputs which have a total Sobol index close to zero), but are useless for quanti-
fying the effects of the influential inputs (because of their non-informative high
values). In contrary, our new results (Fig. 4, right) show that the DGSM-based
upper bounds give a reliable information on the real influence of the inputs (in
terms of contribution to the model output variance) thanks to their closeness to
total Sobol indices. The hierarchy of influence given by the DGSM-based upper
bounds is the same than for total Sobol indices: the flowrate Q is the most in-
fluential input, followed by Hd , Zv and Ks . The remaining four inputs are not
active in the model. We can also note that screening is now fully efficient as Cb
can be judged as non influential (by choosing for example a threshold of 5% of
the total variance for the sensitivity index).
3114 O. Roustant, F. Barthe and B. Iooss
Fig 4. Comparison between total Sobol indices and DGSM-based upper bounds for the simpli-
fied flood model. Left: Results with double exponential transport; Right: New optimal bounds,
obtained with Poincaré constants.
Remark 7. For the DGSM-based upper bound of Ks , our value (Table 3, left)
is different from the one of [24] (Table 5). Indeed, an error is present in [24]
where the multiplicative factor 2π has been omitted (see also the erratum in [31],
Remark 3). Hence, in [24], 0.198 has to be replaced by 1.244.
Fig 5. Representation of a river cross section, showing the main channel (low flow channel)
and the flood plain (high flow channel). Source: [28].
and n output derivatives are obtained. This very large n has been used for
a demonstrative purpose, while in other industrial studies n ranges from 100
to 1 000 ([22, 36]). From the sample of derivatives, the 37 DGSM νi are then
computed. [28] has shown that 32 inputs have DGSM-based upper bounds close
to zero. Then, amongst the 37 inputs, only 5 are potentially active.
Table 4 gives our new results for these 5 remaining inputs. To compute the
DGSM bound, we have also computed Var(f (x)) = 0.369 whose standard devia-
tion (sd = 3.4e−3) is obtained by bootstrap, i.e. by resampling with replacement
11 12
the Monte Carlo sample of output values. As inputs Ks,c and Ks,c follow a uni-
Inputs 11
Ks,c 12
Ks,c dZ 11 dZ 12 Q
ST 0.456 0.0159 0.293 0.015 0.239
(2e−3) (1e−4) (1e−3) (1e−4) (1e−3)
ν 5.695e−3 2.728e−4 1.089e−1 6.592e−3 3.553e−5
(3e−5) (4e−6) (3e−4) (9e−5) (6e−8)
By double exponential transport
Upper bound for CP (μ) - - 6.249 6.249 15623.26
Upper bound for S T - - 1.844 0.116 1.504
- - (2e−3) (2e−3) (1.5e−2)
By logistic transport
Upper bound for CP (μ) - - 1.562 1.562 3905.815
Upper bound for S T - - 0.461 0.028 0.376
- - (4e−3) (5e−4) (4e−3)
Optimal, with the Poincaré constant
CP (μ) 40.528 40.528 0.976 0.976 2441.071
Optimal bound for S T 0.625 0.029 0.288 0.017 0.235
(2e−4) (1e−5) (3e−3) (3e−4) (2e−3)
Table 4
Sensitivity indices for the Mascaret test case. S T gives the total Sobol index, ν the DGSM.
Standard deviations of all these estimates are obtained by bootstrap and given in
parentheses. Partial results are not shown for Ks,c11 , K 12 , which are uniformly distributed.
s,c
3116 O. Roustant, F. Barthe and B. Iooss
form distribution, we report only the optimal DGSM bounds, already used in
[28]. For the 3 other inputs (dZ 11 , dZ 12 and Q), as in the previous application, a
strong decrease factor is obtained for the bounds by using the logistic constant
(gain factor around 4) and the optimal constant (gain factor around 6) instead
of the double exponential constant. The new bounds now allow to identify the
variable dZ 12 as inactive. Moreover, they provide the same ranking than to-
tal Sobol indices for the influential inputs. Notice that the optimal bounds are
nearly equal to the total Sobol indices. The numerical errors, assessed by the
standard deviations of the estimates, explain that the bound values are some-
times slightly smaller than the Sobol values.
In terms of interpretation, the sensitivity analysis results show that the hy-
draulic engineers have to concentrate the research efforts on the knowledge of
the physical parameters Ks and dZ at the 11th river profile, in order to be able
to reduce the prediction uncertainty of the water level when a flood occurs.
From a methodological point of view, the numerical model users are now able
to perform a global sensitivity analysis, in the sense of variance decomposition,
at a lower cost than before by using the DGSM-based technique.
Acknowledgements
References