Citation: Claudia Würtz Jürgensen, Anne Mette Madsen. Influence of everyday activities and presence of people in common indoor environments on exposure to airborne fungi[J]. AIMS Environmental Science, 2016, 3(1): 77-95. doi: 10.3934/environsci.2016.1.77
[1] | Samantha B. Gacasan, Daniel L. Baker, Abby L. Parrill . G protein-coupled receptors: the evolution of structural insight. AIMS Biophysics, 2017, 4(3): 491-527. doi: 10.3934/biophy.2017.3.491 |
[2] | Anna Kahler, Heinrich Sticht . A modeling strategy for G-protein coupled receptors. AIMS Biophysics, 2016, 3(2): 211-231. doi: 10.3934/biophy.2016.2.211 |
[3] | Alyssa D. Lokits, Julia Koehler Leman, Kristina E. Kitko, Nathan S. Alexander, Heidi E. Hamm, Jens Meiler . A survey of conformational and energetic changes in G protein signaling. AIMS Biophysics, 2015, 2(4): 630-648. doi: 10.3934/biophy.2015.4.630 |
[4] | Tika Ram Lamichhane, Hari Prasad Lamichhane . Structural changes in thyroid hormone receptor-beta by T3 binding and L330S mutational interactions. AIMS Biophysics, 2020, 7(1): 27-40. doi: 10.3934/biophy.2020003 |
[5] | Hayet Saidani, Daria Grobys, Marc Léonetti, Hanna Kmita, Fabrice Homblé . Towards understanding of plant mitochondrial VDAC proteins: an overview of bean (Phaseolus) VDAC proteins. AIMS Biophysics, 2017, 4(1): 43-62. doi: 10.3934/biophy.2017.1.43 |
[6] | Sebastian Kube, Petra Wendler . Structural comparison of contractile nanomachines. AIMS Biophysics, 2015, 2(2): 88-115. doi: 10.3934/biophy.2015.2.88 |
[7] | Shivakumar Keerthikumar . A catalogue of human secreted proteins and its implications. AIMS Biophysics, 2016, 3(4): 563-570. doi: 10.3934/biophy.2016.4.563 |
[8] | Nicholas Spellmon, Xiaonan Sun, Wen Xue, Joshua Holcomb, Srinivas Chakravarthy, Weifeng Shang, Brian Edwards, Nualpun Sirinupong, Chunying Li, Zhe Yang . New open conformation of SMYD3 implicates conformational selection and allostery. AIMS Biophysics, 2017, 4(1): 1-18. doi: 10.3934/biophy.2017.1.1 |
[9] | Stephanie H. DeLuca, Samuel L. DeLuca, Andrew Leaver-Fay, Jens Meiler . RosettaTMH: a method for membrane protein structure elucidation combining EPR distance restraints with assembly of transmembrane helices. AIMS Biophysics, 2016, 3(1): 1-26. doi: 10.3934/biophy.2016.1.1 |
[10] | Ateeq Al-Zahrani, Natasha Cant, Vassilis Kargas, Tracy Rimington, Luba Aleksandrov, John R. Riordan, Robert C. Ford . Structure of the cystic fibrosis transmembrane conductance regulator in the inward-facing conformation revealed by single particle electron microscopy. AIMS Biophysics, 2015, 2(2): 131-152. doi: 10.3934/biophy.2015.2.131 |
Dedicated to Italo Capuzzo Dolcetta with friendship, respect, and admiration on the occasion of his retirement.
We consider the weakly coupled $ m $-system of Hamilton-Jacobi equations
$ \lambda v^ \lambda+Bv^ \lambda+H[v^ \lambda] = 0 \ \ \text{ in } \mathbb{T}^n,\;\;\;\;\left( {{{\rm{P}}_\lambda }} \right) $ |
where $ m\in \mathbb{N} $, $ \lambda $ is a nonnegative constant, called the discount factor in terms of optimal control. Here $ \mathbb{T}^n $ denotes the $ n $-dimensional flat torus, $ H = (H_i)_{i\in \mathbb{I}} $ is a family of Hamiltonians given by
$ Hi(x,p)=maxξ∈Ξ[−gi(x,ξ)⋅p−Li(x,ξ)],(H) $
|
where $ \mathbb{I} = \{1, \ldots, m\} $, $ \Xi $ is a given compact metric space, $ g = (g_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n \times \Xi, \mathbb{R}^n)^m $ and $ L = (L_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n \times \Xi)^m $. The unknown in $\left({{{\rm{P}}_\lambda }} \right)$ is an $ \mathbb{R}^m $-valued function $ v^ \lambda = (v^ \lambda_i)_{i\in \mathbb{I}} $ on $ \mathbb{T}^n $, $ B \, :\, C(\mathbb{T}^n)^m \to C(\mathbb{T}^n)^m $ is a linear map represented by a matrix $ B = (b_{ij})_{i, j\in \mathbb{I}}\in C(\mathbb{T}^n)^{m \times m} $, that is,
$ (Bu)_i(x) = (B(x)u(x))_i: = \sum\limits_{j\in \mathbb{I}}b_{ij}(x)u_j(x) \ \ \text{ for } (x,i)\in \mathbb{T}^n \times \mathbb{I}. $ |
We use the abbreviated notation $ H[v^ \lambda] $ to denote $ (H_i(x, Dv_i^ \lambda(x))_{i\in \mathbb{I}} $. The system is called weakly coupled since the $ i $-th equation depends on $ Dv^ \lambda $ only through $ Dv_i^ \lambda $ but not on $ Dv_j^ \lambda $, with $ j\not = i $. Problem $\left({{{\rm{P}}_\lambda }} \right)$ can be stated in the component-wise style as
$ \lambda v_i^ \lambda+\sum\limits_{j\in \mathbb{I}}b_{ij}(x)v_j^ \lambda+H_i(x,Dv_i^ \lambda) = 0 \ \ \text{ in } \mathbb{T}^n,\ i\in \mathbb{I}. $ |
We are mainly concerned with the asymptotic behavior of the solution $ v^ \lambda $ of $\left({{{\rm{P}}_\lambda }} \right)$ as $ \lambda \to 0+ $. Asymptotic problems in this class are called the vanishing discount problem, in view that the constant $ \lambda $ in $\left({{{\rm{P}}_\lambda }} \right)$ appears as a discount factor in the dynamic programming PDE in optimal control.
Recently, there has been a keen interest in the vanishing discount problem concerned with Hamilton-Jacobi equations and, furthermore, fully nonlinear degenerate elliptic PDEs. We refer to [1,7,10,12,19,20,23,24,25,27] for relevant work. The asymptotic analysis in these papers relies heavily on Mather measures or their generalizations and, thus, it is considered part of the Aubry-Mather and weak KAM theories. For the development of these theories we refer to [14,16,17] and the references therein.
We are here interested in the case of systems of Hamilton-Jacobi equations and, indeed, Davini and Zavidovique in [12] have established a general convergence result for the vanishing discount problem for $\left({{{\rm{P}}_\lambda }} \right)$. We establish a result (Theorem 9 below) similar to the main result of [12]. In establishing our convergence result, we adapt the argument in [23] (see also [18]) to the case of systems, especially, to construct generalized Mather measures for $\left({{{\rm{P}}_\lambda }} \right)$. Regarding the recent developments of the weak KAM theory and asymptotic analysis in its influence for systems, we refer to [5,6,26,28,29,30,33].
The new argument, which is different from that of [12], makes it fairly easy to build a generalized Mather measure for systems in a wide class. One advantage of our argument is that it allows us to treat the case where the coupling matrix $ B $ in $\left({{{\rm{P}}_\lambda }} \right)$ depends on the space variable $ x\in \mathbb{T}^n $. As in [20,23], our approach is applicable to the system with nonlinear coupling of fully nonlinear second-order elliptic PDEs, but we restrict ourselves in this paper to the case of the linearly coupled system of first-order Hamilton-Jacobi equations. Another possible approach for constructing generalized Mather measures is the so-called adjoint method (see [5,15,19,27,33]).
This paper is part 1 of our study of the vanishing discount problem for weakly coupled systems of Hamilton-Jacobi equations and deals only with the linear coupling and with compact control sets $ \Xi $. These restrictions make the presentation of our results clear and transparent. In part 2 [20], we remove these restrictions and establish a general convergence result extending Theorem 9 below. Sections 5 and 6 are devoted to the study of ergodic problems of the form $ Bu+H[u] = c $, where $ c\in \mathbb{R}^m $ is an unknown as well. Also, thanks to the linearity of the coupling, our results on the ergodic problems are applied to extend the scope of Theorem 9. On the other hand, the role of the ergodic problem, with general right-hand side $ c $, is not clear at least for the author in the vanishing discount problem for the systems with the nonlinear coupling.
In this paper, we adopt the notion of viscosity solution to $\left({{{\rm{P}}_\lambda }} \right)$, for which the reader may consult [2,4,8,31].
To proceed, we give our main assumptions on the system $\left({{{\rm{P}}_\lambda }} \right)$.
We assume that $ H $ is coercive, that is, for any $ i\in \mathbb{I} $,
$ lim|p|→∞minx∈TnHi(x,p)=∞.(C) $
|
This is a convenient assumption, under which any upper semicontinuous subsolution of $\left({{{\rm{P}}_\lambda }} \right)$ is Lipschitz continuous on $ \mathbb{T}^n $.
We assume that $ B(x) = (b_{ij}(x)) $ is a monotone matrix for every $ x\in \mathbb{T}^n $, that is, it satisfies
$\text{for}\ \text{any}\ x\in {{\mathbb{T}}^{n}},\ \text{if}\ u={{({{u}_{i}})}_{i\in \mathbb{I}}}\in {{\mathbb{R}}^{m}}\ \text{and}\ {{u}_{k}}=\underset{i\in \mathbb{I}}{\mathop{\max }}\,{{u}_{i}}\ge 0,\ \text{then}\ {{(B(x)u)}_{k}}\ge 0.\ \ \ \ \left( \text{M} \right)$ |
This is a natural assumption that $\left({{{\rm{P}}_\lambda }} \right)$ should possess the comparison principle between a subsolution and a supersolution.
In what follows we set, for $ \lambda\geq 0 $,
$ B^ \lambda = \lambda I+B, $ |
and $\left({{{\rm{P}}_\lambda }} \right)$ can be written as
$ B^ \lambda v^ \lambda+H[v^ \lambda] = 0 \ \ \text{ in } \mathbb{T}^n. $ |
We use the symbol $ u\leq v $ (resp., $ u\geq v $) for $ m $-vectors $ u, v\in \mathbb{R}^n $ to indicate $ u_i\leq v_i $ (resp., $ u_i\geq v_i $) for all $ i\in \mathbb{I} $.
The following theorem is well-known: see [13,22] for instance.
Theorem 1. Assume (C) and (M). Let $ \lambda > 0 $. Then the exists a unique solution $ v^ \lambda\in \operatorname{Lip}(\mathbb{T}^n)^m $ of $\left({{{\rm{P}}_\lambda }} \right)$. Also, if $ v = (v_i), w = (w_i) $ are, respectively, upper and lower semicontinuous on $ \mathbb{T}^n $ and a subsolution and a supersolution of $\left({{{\rm{P}}_\lambda }} \right)$, then $ v\leq w $ on $ \mathbb{T}^n $.
Henceforth, let $ {\bf{1}} $ denote the vector $ (1, \ldots, 1)\in \mathbb{R}^m $.
Outline of proof. We follow the line of the arguments in [22]. Although [22] is concerned with the case when the domain is an open subset of a Euclidean space, the results in [22] is valid in the case when the domain is $ \mathbb{T}^n $.
Choose a large constant $ C > 0 $ so that the constant functions $ \pm C {\bf{1}} $ are a supersolution and a subsolution of $\left({{{\rm{P}}_\lambda }} \right)$, respectively. (See also (2.3) below.) According to [22,Theorems 3.3,Lemma 4.8], there is a function $ v^ \lambda = (v_i^ \lambda)_{i\in \mathbb{I}} \, :\, \mathbb{T}^n \to \mathbb{R}^m $ such that the upper and lower semicontinuous envelopes $ (v^ \lambda)^* $ and $ v^ \lambda_* $ are a subsolution and a supersolution of $\left({{{\rm{P}}_\lambda }} \right)$, respectively. By the coercivity assumption (C), we find (see [9,Theorem I.14], [21,Example 1]) that the functions $ (v_i^ \lambda)^* $ are Lipschitz continuous on $ T^n $. Let $ R_1 > 0 $ be a Lipschitz bound of the functions $ (v_i^ \lambda)^* $. To take into account the Lipschitz property of $ (v_i^ \lambda)^* $, we modify the Hamiltonian $ H $. Fix any $ M > 0 $ so that
$ max(x,ξ,i)∈Tn×Ξ×I|gi(x,ξ)|<M, $
|
(1.1) |
and choose constants $ N > 0 $ and $ R_2 > 0 $ so that
$ H_i(x,p)\geq M|p|-N \ \ \text{ for } (x,p,i) \in \mathbb{T}^n \times B_{R_1} \times \mathbb{I}, $ |
and, in view of (1.1),
$ H_i(x,p)\leq M|p|-N \ \ \text{ for } (x,p,i) \in \mathbb{T}^n \times B_{R_2} \times \mathbb{I}. $ |
Define $ G = (G_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n \times \mathbb{R}^n)^m $ by
$ G_i(x,p) = H_i(x,p)\vee (M|p|-N). $ |
By the choice of $ R_1 $, it is easy to see that $ (v^ \lambda)^* $ is a subsolution of
$ λu+Bu+G[u]=0 Tn. $
|
(1.2) |
Also, since $ G\geq H $, $ v_*^ \lambda $ is a supersolution of (1.2). Observe furthermore that, if $ |p|\geq R_2 $, then
$ G_i(x,p) = M|p|-N \ \ \text{ for } (x,i)\in \mathbb{T}^n \times \mathbb{I}, $ |
the functions $ G_i $ are uniformly continuous on $ \mathbb{T}^n \times B_{R_2} $, and hence, for some continuous function $ \omega $ on $ [0, \, \infty) $, with $ \omega(0) = 0 $,
$ |G_i(x,p)-G_i(y,p)|\leq \omega(|x-y|) \ \ \text{ for } (x,y,p)\in( \mathbb{T}^n)^2 \times \mathbb{R}^n,\,i\in \mathbb{I}. $ |
The last inequality above shows that $ G $ satisfies [22,(A.2)], which allows us to apply [22,Theorem 4.7], to conclude that $ (v^ \lambda)^*\leq v^ \lambda_* $ on $ \mathbb{T}^n $ and, moreover, that $ v^ \lambda\in \operatorname{Lip}(\mathbb{T}^n)^* $. Similarly, we deduce that the comparison assertion is valid. Thus, $ v^ \lambda $ is a unique solution of $\left({{{\rm{P}}_\lambda }} \right)$.
Regarding the coercivity (C), the following proposition is well-knwon.
Proposition 2. The function given by (H) satisfies (C) if and only if there exists $ \delta > 0 $ such that
$ Bδ⊂co{gi(x,ξ):ξ∈Ξ} for(x,i)∈Tn×I, $
|
(1.3) |
where $ \, \operatorname{co} $ designates "convex hull" and $ B_ \delta $ denotes the open ball with origin at the origin and radius $ \delta $.
Outline of proof. Set $ C(x, i) = \operatorname{co} \{g_i(x, \xi) \, :\, \xi\in \Xi\}. $ Assume that (1.3) is valid for some $ \delta > 0 $ and observe that
$ Hi(x,p,u)≥maxξ∈Ξ−gi(x,ξ)⋅p−max(x,i,ξ)∈Tn×I×ΞLi(x,ξ)=maxq∈C(x,i)−q⋅p−max(x,i,ξ)∈Tn×I×ΞLi(x,ξ)≥supq∈Bδ−q⋅p−max(x,i,ξ)∈Tn×I×ΞLi(x,ξ)=δ|p|−max(x,i,ξ)∈Tn×I×ΞLi(x,ξ), $
|
which shows that (C) holds.
Next, assume that (1.3) does not hold for any $ \delta > 0 $. Then there exists $ (x_k, i_k)\in \mathbb{T}^n \times \mathbb{I} $ for each $ k\in \mathbb{N} $ such that
$ B_{1/k} \setminus C(x_k,i_k)\not = \emptyset. $ |
For each $ k\in \mathbb{N} $ select $ q_k\in B_{1/k} \setminus C(x_k, i_k) $ and $ r_k\in C(x_k, i_k) $ so that $ r_k $ is the point of $ C(x_k, i_k) $ closest to $ q_k $. (Notice that $ C(x_k, i_k) $ is a compact convex set.) Setting $ \nu_k = (q_k-r_k)/|q_k-r_k| $, we find that
$ \nu_k\cdot (q-r_k)\leq 0 \ \ \text{ for } q\in C(x_k,i_k). $ |
Sending $ k\to\infty $ along an appropriate subsequence, say $ (k_j)_{j\in \mathbb{N}} $, we find that there are a unit vector $ \nu = \lim_{j\to\infty} \nu_{k_j} $ of $ \mathbb{R}^n $, $ r = \lim_{j\to\infty}r_{k_j}\in \mathbb{R}^n $ and $ (x, i)\in \mathbb{T}^n \times \mathbb{I} $ such that
$ r\in C(x,i) \ \ \text{ and } \ \ \nu\cdot (q-r)\leq 0 \ \ \text{ for } q\in C(x,i). $ |
If $ r\not = 0 $, then we have $ \nu = -r/|r| $, since $ \lim_{k\to\infty}q_k = 0 $, and the inequality above reads
$ \nu\cdot q\leq -|r| \lt 0 \ \ \text{ for } q\in C(x,i). $ |
These observations imply that for $ t > 0 $,
$ H_{i}(x,-t\nu) = \max\limits_{\xi\in \Xi} tg_i(x,\xi)\cdot \nu -\min\limits_{\xi\in \Xi} L_i(x,\xi) \leq -\min\limits_{\xi\in \Xi} L_i(x,\xi), $ |
which shows that (C) does not hold. This completes the proof.
The rest of this paper is organized as follows. In Section 2, we recall some basic facts concerning monotone matrices. In Section 3, we study viscosity Green-Poisson measures for our system, which are crucial in our asymptotic analysis. We establish the main result for the vanishing discount problem in Section 4. We study the ergodic problem (P$ _0 $) in the cases when $ B $ is irreducible, and $ B $ is a constant matrix, respectively, in Sections 5 and 6, and combine the results with the analysis on the vanishing discount problem of Section 4.
Here we are concerned with $ m \times m $ real matrix $ B = (b_{ij})_{i, j\in \mathbb{I}} $.
Let $ e_i $ denote the vector $ (e_{i1}, \ldots, e_{im}) $, with $ e_{ii} = 1 $ and $ e_{ij} = 0 $ if $ i\not = j $.
Lemma 3. Let $ B = (b_{ij}) $ be a real $ m \times m $ matrix. It is monotone if and only if
$ bij≤0 ifi≠j and ∑j∈Ibij≥0 fori∈I. $
|
(2.1) |
We remark that if $ B $ satisfies (2.1), then
$ bii=∑j∈Ibij−∑j≠ibij≥0. $
|
(2.2) |
Proof. We assume first that $ B $ is monotone. Since
$ {\bf{1}}_{i} = 1 = \max\limits_{j} {\bf{1}}_{j} \gt 0, $ |
By the monotonicity of $ B $, we have
$ 0≤(B1)i=m∑j=1bij1j=m∑j=1bij for i∈I. $
|
(2.3) |
Similarly, if $ i\not = j $ and $ t\geq 0 $, then we have $ 1 = (e_i-te_j)_i = \max_{k\in \mathbb{I}}(e_i-te_j)_k $ and hence,
$ 0\leq (B(e_i-te_j))_i = b_{ii}-tb_{ij}, $ |
from which we find by sending $ t\to \infty $ that
$ b_{ij}\leq 0. $ |
Hence, (2.1) is satisfied.
Next, we assume that (2.1) holds. Let $ u\in \mathbb{R}^m $ satisfy
$ u_k = \max\limits_{i\in \mathbb{I}}u_i\geq 0. $ |
Then we observe that, since $ u_k\geq u_j $ for all $ j\in \mathbb{I} $,
$ (Bu)_k = \sum\limits_{j\in \mathbb{I}}b_{kj}u_j = b_{kk}u_k +\sum\limits_{j\neq k}b_{kj}u_j = b_{kk}u_k+\sum\limits_{j\neq k}b_{kj}u_k = u_k\sum\limits_{j\in \mathbb{I}}b_{kj}\geq 0. $ |
Thus, $ B $ is monotone.
Lemma 4. Let $ u\in \mathbb{R}^m $ and $ C\geq 0 $ be a constant. Let $ B $ be an $ m \times m $ real monotone matrix. Then we have
$ B(u-C {\bf{1}})\leq Bu\leq B(u+C {\bf{1}}). $ |
Proof. Using Lemma 3, we see that
$ (B {\bf{1}})_i = \sum\limits_{j\in \mathbb{I}}b_{ij}\geq 0 \ \ \text{ for } i\in \mathbb{I}, $ |
which states that $ B {\bf{1}}\geq 0 $. It is then obvious to compute that
$ B(u+C {\bf{1}})-Bu = CB {\bf{1}}, \quad Bu-B(u-C {\bf{1}}) = CB {\bf{1}} \ \ \text{ and } \ \ CB {\bf{1}}\geq 0 $ |
and therefore,
$ B(u+C {\bf{1}})\geq Bu\geq B(u-C {\bf{1}}). $ |
For $ \lambda\geq 0 $ we write $ \mathcal{F}(\lambda) $ for the set of all $ (\phi, u)\in C(\mathbb{T}^n \times \Xi)^m \times C(\mathbb{T}^n)^m $ such that $ u $ is a subsolution of
$ B^ \lambda u+H_{\phi}[u] = 0 \ \ \text{ in } \mathbb{T}^n, $ |
where $ H_{\phi} = (H_{\phi, i})_{i\in \mathbb{I}} $ and
$ H_{\phi.i}(x,p) = \max\limits_{\xi\in \Xi}(-g_i(x,\xi)\cdot p -\phi_i(x,\xi)). $ |
In the above, since $ \phi $ is bounded on $ \mathbb{T}^n \times \Xi $, if $ H $ satisfies (C), then $ H_\phi $ satisfies (C).
Lemma 5. The set $ \mathcal{F}(\lambda) $ is a convex cone in $ C(\mathbb{T}^n \times \Xi)^m \times C(\mathbb{T}^n)^m $ with vertex at the origin.
Proof. Recall [3,Remark 2.5] that for any $ u\in Lip(\mathbb{T}^n)^m $, $ u $ is a subsolution of
$ B^ \lambda u+H[u] = 0 \ \ \text{ in } \mathbb{T}^n $ |
if and only if for any $ i\in \mathbb{I} $,
$ (B^ \lambda u)_i(x)+H_i(x,Du_i(x))\leq 0 \ \ \text{ a.e. in } \mathbb{T}^n, $ |
and by the coercivity (C) that for any $ (\phi, u)\in \mathcal{F}(\lambda) $, we have $ u\in \operatorname{Lip}(\mathbb{T}^n)^m $.
Fix $ (\phi, u), (\psi, v)\in \mathcal{F}(\lambda) $ and $ t, s\in[0, \infty) $. Fix $ i\in \mathbb{I} $ and observe that
$ (Bλu)i(x)+Hϕ,i(x,Dui(x))≤0 a.e. in Tn,(Bλv)i(x)+Hψ,i(x,Dvi(x))≤0 a.e. in Tn, $
|
which imply that there is a set $ N\subset \mathbb{T}^n $ of Lebesgue measure zero such that
$ (Bλu)i(x)≤g(x,ξ)⋅Dui(x)+ϕi(x.ξ) for all (x,ξ)∈Tn∖N×Ξ,(Bλv)i(x)≤gi(x,ξ)⋅Dvi(x)+ψi(x,ξ) for all (x,ξ)∈Tn∖N×Ξ. $
|
Multiplying the first and second by $ t $ and $ s $, respectively, adding the resulting inequalities and setting $ w = tu+sv $, we obtain
$ (B^ \lambda w)_i(x)\leq g(x,\xi)\cdot Dw_i(x) +(t\phi_i+s\psi)(x.\xi) \ \ \text{ for all } (x,\xi) \in \mathbb{T}^n \setminus N\, \times\, \Xi, $ |
which readily implies that $ t(\phi, u)+s(\psi, v)\in \mathcal{F}(\lambda) $.
We refer the reader to [23,Lemma 2.2] for another proof of the above lemma.
We establish a representation formula for the solution of $\left({{{\rm{P}}_\lambda }} \right)$, with $ \lambda > 0 $, by modifying the argument in [23] (see also [18]).
For any nonnegative Borel measure $ \nu $ on $ \mathbb{T}^n \times \Xi $ and $ \phi\in C(\mathbb{T}^n \times \Xi) $, we write
$ \langle{\nu,\phi}\rangle = \int_{ \mathbb{T}^n \times \Xi}\phi(x,\xi)\nu(dx,\,d\xi). $ |
Similarly, for any collection $ \nu = (\nu_i)_{i\in \mathbb{I}} $ of nonnegative Borel measures on $ \mathbb{T}^n \times \Xi $ and $ \phi = (\phi_i)\in C(\mathbb{T}^n \times \Xi)^m $, we write
$ \langle{\nu,\phi}\rangle = \sum\limits_{i\in \mathbb{I}}\langle{\nu_i,\phi_i}\rangle\in \mathbb{R}. $ |
Note that any collection $ \nu = (\nu_i)_{i\in \mathbb{I}} $ of nonnegative Borel measures on $ \mathbb{T}^n \times \Xi $ is regarded as a nonnegative Borel measure on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $ and vice versa.
We set
$ \rho_i(x): = \sum\limits_{j\in \mathbb{I}}b_{ij}(x) \ \ \text{ for } i\in \mathbb{I}. $ |
Note that
$ B1=(b11(x)⋯b1m(x)⋮⋮bm1(x)⋯bmm(x))(1⋮1)=(ρ1(x)⋮ρm(x)) and Bλ1=(λ+ρ1(x)⋮λ+ρm(x)). $
|
(3.1) |
By assumption (M) and Lemma 3, we have $ \rho_i\geq 0 $ on $ \mathbb{T}^n $ for all $ i\in \mathbb{I} $.
Given a constant $ \lambda > 0 $, let $ \operatorname{\mathbb{P}}_{B^ \lambda} $ denote the set of of nonnegative Borel measures $ \nu = (\nu_i)_{i\in \mathbb{I}} $ on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $ such that
$ \langle{\nu,B^ \lambda}\rangle = 1. $ |
The last condition reads
$ \sum\limits_{i\in \mathbb{I}}( \lambda |\nu_i|+\langle{\nu_i,\rho_i}\rangle) = 1, $ |
where $ |\nu_i| $ denotes the total mass of $ \nu_i $ on $ \mathbb{T}^n \times \Xi $. Note as well that $ \operatorname{\mathbb{P}}_{B^ \lambda} $ can be identified with the space of Borel probability measures on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $ by the correspondence between $ \nu = (\nu_i)_{i\in \mathbb{I}} $ and $ \sum_{i\in \mathbb{I}}(\lambda+\rho_i)\nu_i\otimes \delta_i $, where $ \otimes $ indicates the product of two measures and $ \delta_i $ denotes the Dirac measure at $ i $. If we set $ \mu: = \sum_{i\in \mathbb{I}}(\lambda+\rho_i)\nu_i\otimes \delta_i $ and consider $ \mu $ as a collection $ (\mu_i) $ of measures on $ \mathbb{T}^n \times \Xi $, then $ \nu_i = (\lambda+\rho_i)^{-1}\mu_i $. We denote simply by $ \operatorname{\mathbb{P}} $ the space of Borel probability measures on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $.
For $ \lambda\geq 0 $ and $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $ we set
$ \mathcal{G}(z,k, \lambda): = \{\phi-u_k(z) B^ \lambda {\bf{1}} \,:\, (\phi,u)\in \mathcal{F}( \lambda)\}\subset C( \mathbb{T}^n \times \Xi)^m, $ |
and
$ \mathcal{G} \,^\prime(z,k, \lambda) = \{\nu = (\nu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}_{B^ \lambda} \,:\, \langle{\nu,f}\rangle \geq 0 \ \ \text{ for } \ f = (f_i)\in \mathcal{G}(z,k, \lambda)\}. $ |
Theorem 6. Assume (H), (C) and (M). Let $ \lambda > 0 $ and $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. Let $ v^ \lambda\in C(\mathbb{T}^n \times \mathbb{I}) $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$. Then there exists a $ \nu^{z, k, \lambda} = (\nu^{z, k, \lambda}_i)_{i\in \mathbb{I}}\in \mathcal{G} \, ^\prime(z, k, \lambda) $ such that
$ vλk(z)=⟨νz,k,λ,L⟩. $
|
(3.2) |
We remark that for any $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $ we have $ \langle{\nu, L}\rangle\geq v_k^ \lambda(z)\langle{\nu, B^ \lambda}\rangle = v_k^ \lambda(z) $ and, accordingly, in the theorem above, the measures $ \nu^{z, k, \lambda} $ has the minimizing property:
$ vλk(z)=⟨νz,k,λ,L⟩=minν∈G′(z,k,λ)⟨ν,L⟩. $
|
(3.3) |
We call any minimizing family $ (\nu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}_{B^ \lambda} $ of the optimization problem above a viscosity Green-Poisson measure for $\left({{{\rm{P}}_\lambda }} \right)$.
Proof. Note first that $ (L, v^ \lambda)\in \mathcal{F}(\lambda) $ and hence, for any $ \nu\in \mathcal{G} \, ^\prime (z, k, \lambda) $,
$ 0≤⟨ν,L−vλk(z)Bλ⟩=⟨ν,L⟩−vλk(z)⟨ν,Bλ⟩=⟨ν,L⟩−vλk(z). $
|
(3.4) |
Next, we show that
$ sup(ϕ,u)∈F(λ)infν∈PBλ⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩=0. $
|
(3.5) |
Note that for $ z\in \mathbb{T}^n $,
$ sup(ϕ,u)∈F(λ)infν∈PBλ⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩≥infν∈PBλ⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩|(ϕ,u)=(L,vλ)=0. $
|
Hence, in order to prove (3.5), we only need to show that
$ sup(ϕ,u)∈F(λ)infν∈PBλ⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩≤0. $
|
(3.6) |
We postpone the proof of (3.6) and, assuming temporarily that (3.5) is valid, we prove that there exists $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $ such that
$ vλk(z)=⟨ν,L⟩, $
|
(3.7) |
which, together with (3.4), completes the proof.
To prove (3.7), we observe that $ \operatorname{\mathbb{P}}_{B^ \lambda} $ and, by Lemma 5, $ \mathcal{F}(\lambda) $ are convex,
$ \operatorname{\mathbb{P}}_{B^ \lambda}\ni \nu\mapsto \langle{\nu,L-\phi+(u_k(z)-v_k^ \lambda(z))B^ \lambda}\rangle $ |
is convex and continuous, in the topology of weak convergence of measures, for any $ (\phi, u)\in \mathcal{F}(\lambda) $ and
$ \mathcal{F}( \lambda)\ni (\phi,u)\mapsto \langle{\nu,L-\phi+(u_k(z)-v_k^ \lambda(z))B^ \lambda}\rangle $ |
is concave and continuous for any $ \nu\in \operatorname{\mathbb{P}}_{B^ \lambda} $. Hence, noting moreover that $ \mathbb{T}^n \times \Xi \times \mathbb{I} $ is a compact set, we apply the minimax theorem ([34,32]), to find from (3.5) that
$ 0=sup(ϕ,u)∈F(λ)minν∈PBλ⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩=minν∈PBλsup(ϕ,u)∈F(λ)⟨ν,L−ϕ+(uk(z)−vλk(z))Bλ⟩. $
|
(3.8) |
Observe by using the cone property of $ \mathcal{F}(\lambda) $ that
$ \sup\limits_{(\phi,u)\in \mathcal{F}( \lambda)}\langle{\nu, u_k(z)B^ \lambda -\phi}\rangle = {0 if ν∈G′(z,k,λ),∞ if ν∈PBλ∖G′(z,k,λ). $
|
This and (3.8) yield
$ 0=minν∈PBλsup(ϕ,u)∈F(λ)⟨ν,L−ϕ+(uk(z)−vλk(z)Bλ⟩=minν∈G′(z,k,λ)sup(ϕ,u)∈F(λ)⟨ν,L−vλk(z)Bλ⟩=minν∈G′(z,k,λ)⟨ν,L−vλk(z)Bλ⟩=minν∈G′(z,k,λ)(⟨ν,L⟩−vλk(z)⟨ν,Bλ⟩)=minν∈G′(z,k,λ)⟨ν,L⟩−vλk(z), $
|
which proves (3.7).
It remains to show (3.6). For this, we argue by contradiction and thus suppose that (3.6) does not hold. Accordingly, we have
$ \sup\limits_{(\phi,u)\in \mathcal{F}( \lambda)}\inf\limits_{\nu\in \operatorname{\mathbb{P}}_{B^ \lambda}} \langle{\nu,L-\phi+(u_k(z)-v_k^ \lambda(z))B^ \lambda}\rangle \gt \varepsilon $ |
for some $ \varepsilon > 0 $. We may select $ (\phi, u)\in \mathcal{F}(\lambda) $ so that
$ \inf\limits_{\nu\in \operatorname{\mathbb{P}}_{B^ \lambda}}\langle{\nu,L-\phi+(u_k(z)-v_k^ \lambda(z))B^ \lambda}\rangle \gt \varepsilon. $ |
That is, for any $ \nu\in \operatorname{\mathbb{P}}_{B^ \lambda} $, we have
$ \langle{\nu,L-\phi+(u_k(z)-v_k^ \lambda(z))B^ \lambda}\rangle \gt \varepsilon = \langle{\nu, \varepsilon B^ \lambda}\rangle. $ |
Plugging $ \nu = (\lambda+\rho_i)^{-1} \delta_{(x, \xi, i)}\in \operatorname{\mathbb{P}}_{B^ \lambda} $, with any $ (x, \xi, i)\in \mathbb{T}^n \times \Xi \times \mathbb{I} $, into the above, we find that
$ (L_i-\phi_i)(x,\xi)-(v_k^ \lambda(z)-u_k(z)- \varepsilon)(B^ \lambda {\bf{1}})_i \gt 0. $ |
Hence, we have
$ \phi(x,\xi) \lt L(x,\xi)+(u_k(z)-v_k^ \lambda(z)- \varepsilon)B^ \lambda {\bf{1}} \ \ \text{ for } (x.\xi)\in \mathbb{T}^n \times \mathbb{R}^n. $ |
This ensures that $ u $ is a subsolution of
$ B^ \lambda u+H[u] = (u_k(z)-v_k^ \lambda(z)- \varepsilon)B^ \lambda {\bf{1}} \ \ \text{ in } \mathbb{T}^n, $ |
which implies that $ u-(u_k(z)-v_k^ \lambda(z)- \varepsilon) {\bf{1}} $ is a subsolution of $\left({{{\rm{P}}_\lambda }} \right)$. By comparison (Theorem 1), we get
$ u(x)-(u_k(z)-v_k^ \lambda(z)- \varepsilon)\leq v^ \lambda(x) \ \ \text{ for } x\in \mathbb{T}^n. $ |
The $ k $-th component of the above, evaluated at $ x = z $, yields an obvious contradiction. Thus we conclude that (3.6) holds.
We have the following characterization of $ \mathcal{G} \, ^\prime(z, k, \lambda) $.
Proposition 7. Assume (H), (C) and (M) hold. Let $ \nu = (\nu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}_{B^ \lambda} $ and $ (z, k, \lambda)\in \mathbb{T}^n \times \mathbb{I} \times(0, \infty) $. Then we have $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $ if and only if
$ ∑i∈I⟨νi,(Bλψ)i−gi⋅Dψi⟩=ψk(z) for ψ=(ψi)i∈I∈C1(Tn)m. $
|
(3.9) |
Proof. Assume first that $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $. Fix any $ \psi = (\psi_i)_{i\in \mathbb{I}} \in C^1(\mathbb{T}^n)^m $ and define $ \phi = (\phi_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n \times \mathbb{I})^m $ by
$ \phi_i(x,\xi) = (B^ \lambda \psi)_i(x) -g_i(x,\xi)\cdot D\psi_i(x). $ |
Observe that $ u: = \pm\psi $ satisfy, respectively,
$ B^ \lambda u+H_{\pm\phi}[u] = 0 \ \ \text{ in } \mathbb{T}^n, $ |
and, hence,
$ \pm(\phi,\psi)\in \mathcal{F}( \lambda). $ |
Since $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $, we have
$ \pm \psi_k(z)\leq \langle{\nu,\pm\phi}\rangle = \pm\langle{\nu,\phi}\rangle, $ |
respectively, which shows that (3.9) is valid.
Now, assume that (3.9) is satisfied. Fix any $ (u, \phi)\in \mathcal{F}(\lambda) $. As noted in the proof of Theorem 1, we have $ u\in \operatorname{Lip}(\mathbb{T}^n) $. By the standard mollification technique, given a positive constant $ \varepsilon > 0 $, we can approximate $ u $ by a smooth function $ u^ \varepsilon $ so that
$ \max\limits_{ \mathbb{T}^n}|u-u^ \varepsilon| \lt \varepsilon \ \ \text{ and } \ \ B^ \lambda u^ \varepsilon+H_\phi[u^ \varepsilon]\leq \varepsilon B^ \lambda {\bf{1}} \ \ \text{ in } \mathbb{T}^n. $ |
The last inequality reads
$ B^ \lambda u^ \varepsilon_i(x)-g_i(x,\xi)\cdot Du^ \varepsilon_i(x)-\phi_i(x,\xi)\leq \varepsilon (B^ \lambda {\bf{1}})_i(x) \ \ \text{ for } (x,\xi,i)\in \mathbb{T}^n \times \mathbb{R}^n \times \mathbb{I}. $ |
Integrating the above by $ \nu_i $, summing up in $ i\in \mathbb{I} $ and using (3.9), we get
$ u^ \varepsilon_k(z)-\langle{\nu,\phi}\rangle\leq \varepsilon\langle{\nu,B^ \lambda}\rangle = \varepsilon. $ |
Sending $ \varepsilon\to 0 $ shows that $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $.
It is convenient to restate the theorem above as follows. For $ \mu = (\mu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}} $ and $ \lambda > 0 $, consider $ \nu = (\nu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}_{B^ \lambda} $ given by
$ \nu_i: = ( \lambda+\rho_i)^{-1}\mu_i = \frac{1}{(B^ \lambda {\bf{1}})_i}\mu_i. $ |
(Notice by the above definition that $ \langle{\nu, B^ \lambda}\rangle = \langle{\mu, }\rangle = 1 $.) Observe that for $ \phi = (\phi_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n \times \Xi)^m $,
$ \langle{\nu,\phi}\rangle = \sum\limits_{i\in \mathbb{I}}\langle{\nu_i,\phi_i}\rangle = \sum\limits_{i\in \mathbb{I}}\langle{\mu_i, ( \lambda+\rho_i)^{-1}\phi_i}\rangle, $ |
and that for any $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $, we have $ \nu\in \mathcal{G} \, ^\prime(z, k, \lambda) $ if and only if
$ ∑i∈I⟨μi,(λ+ρi)−1ϕi⟩≥uk(z) for (ϕ,u)∈F(λ). $
|
(3.10) |
The condition above is stated in the spirit of Proposition 7 as
$ \sum\limits_{i\in \mathbb{I}}\langle{\mu_i,( \lambda+\rho_i)^{-1}((B^ \lambda\psi)_i -g_i\cdot D\psi_i)}\rangle = \psi_k(z) \ \ \text{ for } \ \psi = (\psi_i)_{i\in \mathbb{I}}\in C^1( \mathbb{T}^n)^m. $ |
We define
$ \operatorname{\mathbb{P}}(z,k, \lambda) = \{\mu = (\mu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}} \,:\, \mu \text{ satisfies (3.10)}\}. $ |
The following proposition is an immediate consequence of Theorem 6.
Corollary 8. Assume (H), (C) and (M). Let $ \lambda > 0 $ and $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. Let $ v^ \lambda\in C(\mathbb{T}^n \times \mathbb{I}) $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$. Then there exists a $ \mu^{z, k, \lambda} = (\mu^{z, k, \lambda}_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}(z, k, \lambda) $ such that
$ vλk(z)=∑i∈I⟨μz,k,λi,(λ+ρi)−1Li⟩=minμ=(μi)i∈I∈P(z,k,λ) ∑i∈I⟨μi,(λ+ρi)−1Li⟩. $
|
(3.11) |
We study the asymptotic behavior of the solution $ v^ \lambda $ of $\left({{{\rm{P}}_\lambda }} \right)$, with $ \lambda > 0 $, as $ \lambda\to 0 $.
We make a convenient assumption on the system (P$ _0 $):
$ \text{problem}\ ({{\text{P}}_{0}})\ \text{has}\ \text{a}\ \text{solution}\ {{v}_{0}}\in \operatorname{Lip}({{\mathbb{T}}^{n}}).\text{ } $ |
If $ \rho_i > 0 $ for all $ i\in \mathbb{I} $, then Theorem 1 assures that there exists a unique solution $ v_0 $ of (E). In this situation, it is not difficult to show that the uniform convergence, as $ \lambda \to 0+ $, of $ v^ \lambda $ to the unique solution $ v_0 $ on $ \mathbb{T}^n $. In general, existence and uniqueness of a solution of (P$ _0 $) may fail. In fact, one can prove at least in the case when the $ b_{ij} $ are constants (see Theorem 18) that there exists $ c\in \mathbb{R}^m $ such that
$ Bu+H[u]=c in Tn $
|
(4.1) |
has a solution $ v_0\in \operatorname{Lip}(\mathbb{T}^n) $ and possibly multiple solutions. If such a $ c = (c_i) $ exists, then the introduction of a new family of Hamiltonians,
$ \widetilde H = (\widetilde H_i)_{i\in \mathbb{I}}, \quad \text{ with }\widetilde H_i(x,p) = H_i(x,p)-c_i, $ |
allows us to view (4.1) as in the form of (P$ _0 $). The link between two vanishing discount problems for the original $\left({{{\rm{P}}_\lambda }} \right)$ and for $\left({{{\rm{P}}_\lambda }} \right)$, with $ \widetilde H $ in place of $ H $, is discussed in Sections 5 and 6.
Theorem 9. Assume (H), (C), (M) and (E). Let $ v^ \lambda $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$ for $ \lambda > 0 $. Then there exists a solution $ v^0\in \operatorname{Lip}(\mathbb{T}^n)^m $ of (P$ _0 $) such that the functions $ v_i^ \lambda $ converge to $ v_i^0 $ uniformly on $ \mathbb{T}^n $ as $ \lambda\to 0 $ for all $ i\in \mathbb{I} $.
Lemma 10. Under the hypotheses of Theorem 9, there exists a constant $ C_0 > 0 $ such that for any $ \lambda > 0 $,
$ |vλi(x)|≤C0 for(x,i)∈Tn×I. $
|
(4.2) |
Proof. Let $ v_0 = (v_{0, i})_{i\in \mathbb{I}}\in \operatorname{Lip}(\mathbb{T}^n)^m $ be the solution of (P$ _0 $). Choose a constant $ C_0 > 0 $ so that
$ |v_{0,i}(x)|\leq C_1 \ \ \text{ for } (x,i)\in \mathbb{T}^n \times \mathbb{I}, $ |
and observe by the monotonicity of $ B $ (Lemma 4) that $ v_0+C_1 {\bf{1}} $ and $ v_0-C_1 {\bf{1}} $ are a supersolution and a subsolution of (P$ _0 $), respectively. Noting that $ v_0+C_1 {\bf{1}}\geq 0 $ and $ v_0-C_1 {\bf{1}}\leq 0 $, we deduce that $ v_0+C_1 {\bf{1}}\geq 0 $ and $ v_0-C_1 {\bf{1}}\leq 0 $ are a supersolution and a subsolution of $\left({{{\rm{P}}_\lambda }} \right)$ for any $ \lambda > 0 $, respectively. By comaprison (Theorem 1), we see that, for any $ \lambda > 0 $, $ v_0-C_1 {\bf{1}}\leq v^ \lambda\leq v_0+C_1 {\bf{1}} $ on $ \mathbb{T}^n $ and, moreover, $ -2C_1 {\bf{1}}\leq v^ \lambda\leq 2C_1 {\bf{1}} $ on $ \mathbb{T}^n $. Thus, (4.2) holds with $ C_0 = 2C_1 $.
Lemma 11. Under the hypotheses of Theorem 9, the family $ (v^ \lambda)_{ \lambda\in(0, \, 1)} $ is equi-Lipschitz continuous on $ \mathbb{T}^n $.
Indeed, the family $ (v^ \lambda)_{ \lambda > 0} $ is equi-Lipschitz continuous on $ \mathbb{T}^n $, which we do not need here.
Proof. By Lemma 10, there is a constant $ C_0 > 0 $ such that
$ |(B^ \lambda v^ \lambda(x))_i|\leq C_0 \ \ \text{ for } (x,i, \lambda)\in \mathbb{T}^n \times \mathbb{I} \times(0,\,1). $ |
Hence, as $ v^ \lambda $ is a solution of $\left({{{\rm{P}}_\lambda }} \right)$, we deduce by (C) that there exists a constant $ C_1 > 0 $ such that the $ v_i^ \lambda $ are subsolutions of $ |Du|\leq C_1 $ in $ \mathbb{T}^n $. It is a standard fact that the $ v_i^ \lambda $ are Lipschitz continuous on $ \mathbb{T}^n $ with $ C_1 $ as their Lipschitz bound.
In the proof of Theorem 9, Corollary 8 has a crucial role. We need also results for $ \lambda = 0 $ similar to the corollary.
We consider the condition for $ \mu\in \operatorname{\mathbb{P}} $,
$ ⟨μ,ϕ⟩≥0 for (ϕ,u)∈F(0). $
|
(4.3) |
We denote by $ \operatorname{\mathbb{P}}(0) $ the subset of $ \operatorname{\mathbb{P}} $ consisting of those $ \mu $ which satisfy (4.3).
Theorem 12. Assume (H), (C), (M) and (E). Assume that $ \rho_i = 0 $ on $ \mathbb{T}^n $ for every $ i\in \mathbb{I} $. Then there exists a $ \mu^0 = (\mu_i^0)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}(0) $ such that
$ 0=⟨μ0,L⟩=minμ∈P(0)⟨μ,L⟩. $
|
(4.4) |
Proof. We fix a $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. By Corollary 8, for each $ \lambda > 0 $ there exists $ \mu^ \lambda = (\mu_i^ \lambda)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}}(z, k, \lambda) $ such that
$ λvλk(z)=∑i∈Iλ⟨μλi,λ−1Li⟩=⟨μλ,L⟩. $
|
(4.5) |
Since $ (\mu^ \lambda)_{ \lambda > 0} $ is a family of Borel probability measures on a compact space $ \mathbb{T}^n \times \Xi \times \mathbb{I} $, there exists a sequence $ (\lambda_j)_{j\in \mathbb{N}}\subset (0, \, 1) $ converging to zero such that the sequence $ (\mu^{ \lambda_j})_{j\in \mathbb{N}} $ converges weakly in the sense of measures to a Borel probability measure $ \mu^0 $ on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $. It follows from (4.5) and Lemma 10 that
$ 0 = \langle{\mu^0,L}\rangle. $ |
Observe that if $ (\phi, u)\in \mathcal{F}(0) $, then, for any $ \lambda > 0 $, $ u $ is a subsolution of
$ B^ \lambda u+H_\phi[u] = \lambda u \ \ \text{ in } \mathbb{T}^n, $ |
and hence, $ (\psi, u) \in \mathcal{F}(\lambda) $, with $ \psi(x, \xi) = \phi(x, \xi)+ \lambda u(x) $. Hence, the inclusion $ \mu^ \lambda\in \mathcal{G} \, ^\prime(z, k, \lambda) $ yields
$ u_k(z)\leq \sum\limits_{i\in \mathbb{I}}\langle{\mu_i^ \lambda, \lambda^{-1}(\phi_i+ \lambda u_i)}\rangle = \lambda\langle{\mu^ \lambda,\phi}\rangle+\langle{\mu^ \lambda,u}\rangle. $ |
Multiplying the above by $ \lambda $ and sending $ \lambda = \lambda_j \to 0 $, in view of Lemma 10, we get
$ 0\leq \langle{\mu^0,\phi}\rangle. $ |
This shows that $ \mu^0\in \operatorname{\mathbb{P}}(0) $. These observations together with (4.3) for $ \mu\in \operatorname{\mathbb{P}}(0) $ guarantee that
$ 0 = \langle{\mu^0,L}\rangle = \min\limits_{\mu\in \operatorname{\mathbb{P}}(0)}\langle{\mu,L}\rangle. $ |
We state a characterization of $ \operatorname{\mathbb{P}}(0) $ in the next, similar to Proposition 7, which we leave to the reader to verify.
Proposition 13. Assume (H), (C) and (M). Let $ \mu = (\mu_i)_{i\in \mathbb{I}}\in \operatorname{\mathbb{P}} $. We have $ \mu\in \operatorname{\mathbb{P}}(0) $ if and only if
$ \sum\limits_{i\in \mathbb{I}}\langle{\mu_i,(B\psi)_i-g_i\cdot D\psi_i}\rangle = 0 \ \ \mathit{\text{for}} \ \psi = (\psi_i)_{i\in \mathbb{I}}\in C^1( \mathbb{T}^n)^m. $ |
We call any minimizer $ \mu\in \operatorname{\mathbb{P}}(0) $ of the optimization problem (4.4) a viscosity Mather measure.
We denote by $ \mathbb{M}_+ $ the set of all Borel nonnegative measures $ \mu = (\mu_i)_{i\in \mathbb{I}} $ on $ \mathbb{T}^n \times \Xi \times \mathbb{I} $. We set
$ \mathbb{M}_+(0) = \{\mu\in \mathbb{M}_+ \,:\, \mu \text{ satisfies (4.3)}\}. $ |
Theorem 14. Let $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. Assume (H), (C), (M) and (E). For any $ \lambda > 0 $, let $ v^{ \lambda} $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$ and $ \mu^ \lambda \in \operatorname{\mathbb{P}}(z, k, \lambda) $ be a minimizer of (3.11). Then there exists a subsequence of $ (\lambda_j) $, which is denoted again by the same symbol, such that, as $ j\to\infty $,
$ \frac{ \lambda_j}{ \lambda_j+\rho_i}\mu_i^{ \lambda_j} \to \mu_i^0 $ |
weakly in the sense of measures for some $ \mu^0 = (\mu^0_i)_{i\in \mathbb{I}}\in \mathbb{M}_+(0) $, and $ \mu^0 $ satisfies
$ ⟨μ0,L⟩=0. $
|
(4.6) |
In particular,
$ 0=⟨μ0,L⟩=minμ∈M+(0)⟨μ,L⟩. $
|
(4.7) |
Notice that the minimization problem (4.7) is trivial since $ \mu^0 = 0 $ is a minimizer.
Proof. The proof is similar to that of Theorem 12.
We fix a $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. For each $ \lambda > 0 $, we have
$ λvλk(z)=∑i∈Iλ⟨μλi,(λ+ρi)−1Li⟩. $
|
(4.8) |
Observe that
$ \langle{ \lambda( \lambda+\rho_i)^{-1}\mu^ \lambda_i,}\rangle\leq \langle{\mu^ \lambda_i,}\rangle = \sum\limits_{i\in \mathbb{I}}|\mu_i^ \lambda| = 1. $ |
Accordingly, since $ \mathbb{T}^n \times \Xi \times \mathbb{I} $ is a compact metric space, the families $ (\lambda(\lambda+\rho_i)^{-1}\mu_i^ \lambda)_{ \lambda = \lambda_j, j\in \mathbb{N}} $ have a common subsequence, along which all the families converge to some Borel nonnegative measures $ \mu^0_i $ weakly in the sense of measures. We may assume by replacing the original sequence $ (\lambda_j) $ by its subsequence that
$ \frac{ \lambda_j}{ \lambda_j+\rho_i}\mu_i^{ \lambda_j} \to \mu_i^0 $ |
weakly in the sense of measures. Combining this with (4.8) yields
$ 0 = \sum\limits_{i\in \mathbb{I}}\langle{\mu^0_i,L_i}\rangle = \langle{\mu^0,L}\rangle. $ |
It is obvious to see that $ \mu^0\in \mathbb{M}_+ $.
Let $ (\phi, u)\in \mathcal{F}(0) $. As before, we have $ (\psi, u) \in \mathcal{F}(\lambda) $, with $ \psi(x, \xi) = \phi(x, \xi)+ \lambda u(x) $ and moreover
$ u_k(z)\leq \sum\limits_{i\in \mathbb{I}}\langle{\mu_i^ \lambda,( \lambda+\rho_i)^{-1}(\phi_i+ \lambda u_i)}\rangle = \langle{\mu^ \lambda,( \lambda+\rho_i)^{-1}\phi}\rangle+ \lambda\langle{\mu^ \lambda,( \lambda+\rho_i)^{-1}u}\rangle. $ |
Multiplying the above by $ \lambda $ and sending $ \lambda = \lambda_j \to 0 $, we get
$ 0\leq \langle{\mu^0,\phi}\rangle. $ |
This shows that $ \mu^0\in \mathbb{M}_+(0) $.
Proof of Theorem 9. Let $ \mathcal{V} $ denote the set of accumulation points $ v = (v_i)\in C(\mathbb{T}^n)^m $ in the space $ C(\mathbb{T}^n)^m $ of $ v^ \lambda $ as $ \lambda\to 0 $. In view of the Ascoli-Arzela theorem, Lemmas 4.2 and 4.3 guarantee that the family $ (v^ \lambda)_{ \lambda\in(0, \, 1)} $ is relatively compact in $ C(\mathbb{T}^n)^m $. In particular, the set $ \mathcal{V} $ is nonempty. Note by the stability of the viscosity property under uniform convergence that any $ v\in \mathcal{V} $ is a solution of (P$ _0 $).
If $ \mathcal{V} $ is a singleton, then it is obvious that the whole family $ (v^ \lambda)_{ \lambda > 0} $ converges to the unique element of $ \mathcal{V} $ in $ C(\mathbb{T}^n)^m $ as $ \lambda\to 0 $.
We need only to show that $ \mathcal{V} $ is a singleton. It is enough to show that for any $ v, w\in \mathcal{V} $ and $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $, the inequality $ w_k(z)\leq v_k(z) $ holds.
Fix any $ v, w\in \mathcal{V} $ and $ (z, k)\in \mathbb{T}^n \times \mathbb{I} $. Select sequences $ (\lambda_j) $ and $ (\delta_j) $ converging to zero so that
$ v^{ \lambda_j} \to v,\ \ v^{ \delta_j} \to w \ \ \text{ in } C(T^n)^m \ \ \text{ as } j\to\infty. $ |
By Corollary 8, there exists a sequence $ (\mu^j)_{j\in \mathbb{N}} $ such that
$ μj∈G′(z,k,λj) and vλjk(z)=∑i∈I⟨μji,(λj+ρi)−1Li⟩ for j∈N. $
|
(4.9) |
In view of Theorem 14, we may assume by passing to a subsequence if necessary that, as $ j\to\infty $,
$ \frac{ \lambda_j}{ \lambda_j+\rho_i}\mu_i^j \to \mu_i^0 \ \ \text{ weakly in the sense of measures} $ |
for all $ i\in \mathbb{I} $ and for some $ \mu^0 = (\mu^0_i)_{i\in \mathbb{I}}\in \mathbb{M}_+(0) $ and, moreover,
$ 0=⟨μ0,L⟩. $
|
(4.10) |
Since $ (L- \lambda v^ \lambda, v^ \lambda)\in \mathcal{F}(0) $ and $ \mu^0\in \mathbb{M}_+(0) $, in view of (4.10), we have
$ 0\leq \langle{\mu^0,L- \lambda v^ \lambda}\rangle = \langle{\mu^0,L}\rangle-\langle{\mu^0, \lambda v^ \lambda}\rangle = - \lambda\langle{\mu^0,v^ \lambda}\rangle, $ |
which yields after dividing by $ \lambda > 0 $ and then sending $ \lambda \to 0 $ along $ \lambda = \delta_j $
$ ⟨μ0,w⟩≤0. $
|
(4.11) |
Now, note that $ w $ is a solution of
$ B^ \lambda w+H[w] = \lambda w \ \ \text{ in } \mathbb{T}^n, $ |
and thus, $ (L+ \lambda w, w)\in \mathcal{F}(\lambda) $ and infer by (4.9) that
$ w_k(z)\leq \sum\limits_{i\in \mathbb{I}}\langle{\mu^j_i,( \lambda_j+\rho_i)^{-1}(L_i+ \lambda_j w_i)}\rangle = v_k^{ \lambda_j}(z)+ \lambda_j\sum\limits_{i\in \mathbb{I}}\langle{\mu^j_i,( \lambda_j+\rho_i)^{-1}w_i}\rangle. $ |
Sending $ j\to\infty $ now yields
$ w_k(z)\leq v_k(z)+\langle{\mu^0,w}\rangle. $ |
This together with (4.11) shows that $ w_k(z)\leq v_k(z) $, which completes the proof.
We consider the problem of finding $ c = (c_i)_{i\in \mathbb{I}}\in \mathbb{R}^m $ and $ v = (v_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n)^m $ such that $ v $ is a solution of
$ Bv+H[v]=c in Tn. $
|
(5.1) |
The pair of such $ c $ and $ v $ is also called a solution of (5.1). This problem is called the ergodic problem in this paper although the term, ergodic problem, should be used only when the condition that $ \sum_{j\in \mathbb{I}}b_{ij}(x) = 0 $ holds for some $ (i, x)\in \mathbb{I} \times \mathbb{T}^n $.
Henceforth, $ D(x) $ denotes the diagonal matrix
$ D(x) = \operatorname{diag}(\rho_1(x),\ldots,\rho_m(x)) \ \ \ \text{ for } x\in \mathbb{T}^n, $ |
where, as before, $ \rho_i(x) = \sum_{j\in \mathbb{I}}b_{ij}(x) $.
Throughout this section, we treat the case when
$ B(x) is irreducible. $
|
(5.2) |
The irreducibility of $ B(x) $ is stated as follows: for any nonempty subset $ I $ of $ \mathbb{I} $, which is not identical to $ \mathbb{I} $, there exists a pair of $ i\in I $ and $ j\in \mathbb{I} \setminus I $ such that $ b_{ij}(x)\not = 0 $.
The following result has been established in Davini-Zavidovique [11,Theorem 2.10] (see also [6,30]).
Proposition 15. Assume (H), (C), (M), (5.2), and that
$ ∑j∈Ibij(x)=0 for all(i,x)∈I×Tn. $
|
(5.3) |
Then there exist $ c_0\in \mathbb{R} $ and $ v_0\in \operatorname{Lip}(\mathbb{T}^n)^m $ such that the pair $ (c_0 {\bf{1}}, v_0) $ is a solution of (5.1).
We remark that (5.3) is satisfied if and only if $ B(x) {\bf{1}} = 0 $ for all $ x\in \mathbb{T}^n $, which holds if and only if $ \rho_i(x) = 0 $ for all $ (i, x)\in \mathbb{I} \times \mathbb{T}^n $.
The next theorem states the central result of this section.
Theorem 16. Assume (H), (C), (M), (5.2), and (5.3). Let $ v^ \lambda $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$ for $ \lambda > 0 $. Then there exists a constant $ c^0\in \mathbb{R} $ and a function $ v^0\in \operatorname{Lip}(\mathbb{T}^n)^m $ such that the functions $ v^ \lambda+ \lambda^{-1}c^0 {\bf{1}} $ converge to $ v^0 $ uniformly on $ \mathbb{T}^n $ as $ \lambda\to 0 $. Moreover, the pair $ (c^0 {\bf{1}}, v^0) $ is a solution of (5.1).
Proof. Thanks to Proposition 15, there exists a solution $ (c_0, v_0)\in \mathbb{R}^m \times C(\mathbb{T}^n)^m $ of (5.1). We set $ \widetilde H = H-c_0 {\bf{1}} $, and note that, since $ B(x) {\bf{1}} = 0 $ for all $ x\in \mathbb{T}^n $, the function $ w^ \lambda: = v^ \lambda+ \lambda^{-1}c_0 {\bf{1}} $ satisfies, in the viscosity sense,
$ \lambda w^ \lambda+Bw^ \lambda+ \widetilde H[w^ \lambda] = \lambda v^ \lambda +c_0 {\bf{1}}+Bv^ \lambda+H[v^ \lambda]-c_0 {\bf{1}} = 0. $ |
By Theorem 9, there exists a solution $ v^0\in \operatorname{Lip}(\mathbb{T}^n)^m $ of $ Bv^0+ \widetilde H[v^0] = 0 $ in $ \mathbb{T}^n $ such that, as $ \lambda\to 0+ $, $ w^ \lambda\to v^0 $ in $ C(\mathbb{T}^n)^m $. Noting that $ (c_0 {\bf{1}}, v^0) $ is a solution of (5.1), we finish the proof.
The condition (5.3) in Proposition 15 can be removed and the following theorem is valid.
Theorem 17. Assume (H), (C), (M), and (5.2). Then there exist $ c^0\in \mathbb{R} $ and $ v^0 = (v^0_i)_{i\in \mathbb{I}}\in \operatorname{Lip}(\mathbb{T}^n)^m $ such that the pair $ (c^0 {\bf{1}}, v^0) $ is a solution of (5.1).
Proof. For $ x\in \mathbb{I} \times \mathbb{T}^n $, we set
$ B^0(x) = (b^0_{ij}(x)): = B(x)-D(x). $ |
and note that $ B^0(x) $ is irreducible and (5.3) holds with $ b_{ij}(x) $ replaced by $ b^0_{ij}(x) $. Note also that $ \rho_i(x)\geq 0 $ for all $ (i, x)\in \mathbb{I} \times \mathbb{T}^n $.
Thanks to Proposition 15, there exist $ c^0\in \mathbb{R} $ and $ v = (v_i)\in \operatorname{Lip}(\mathbb{T}^n)^m $ which solve
$ B^0v+H[v] = c^0 {\bf{1}} \ \ \text{ in } \mathbb{T}^n. $ |
We choose a constant $ C > 0 $ so that $ \max_{(i, x)\in \mathbb{I} \times \mathbb{T}^n}|v_i(x)|\leq C $ and set $ v^\pm(x) = v(x)\pm C {\bf{1}} $, respectively. Observe that, since $ v^+_i(x)\geq 0 $ and $ v^-_i(x)\leq 0 $ for all $ (i, x)\in \mathbb{I} \times \mathbb{T}^n $, the functions $ u = v^+ $ and $ u = v^- $ are a supersolution and subsolution of
$ B^0u+Pu+H[u] = c^0 {\bf{1}} \ \ \text{ in } \mathbb{T}^n, $ |
that is, $ Bu+H[u] = c^0 {\bf{1}} \ \ \text{ in } \mathbb{T}^n $, respectively. In view of the Perron method, the function $ v^0 = (v^0_i)_{i\in \mathbb{I}}\in \operatorname{Lip}(\mathbb{T}^n) $ given by
$ v0i(x)=sup{ui(x):u=(ui)∈C(Tn)m is a subsolution of Bu+H[u]=c01 in Tn,v−≤u≤v+ in Tn}, $
|
is a solution of (5.1), with $ c = c^0 {\bf{1}} $.
Even without the assumption (5.3), it is immediate from Theorem 9 that, under the hypotheses of Theorem 17, if $ c^0 = 0 $, then the convergence holds for the whole family of the solutions $ v^ \lambda $ of $\left({{{\rm{P}}_\lambda }} \right)$, with $ \lambda > 0 $. A typical case when $ c^0 = 0 $ is given by [6,Theorem 4.2] (see also [11,28]).
Throughout this section we assume that $ B $ is a constant matrix, that is, independent of $ x\in \mathbb{T}^n $.
The main results in this section are as follows.
Theorem 18. Assume (H), (C), (M), and that $ B $ is a constant matrix. Then (5.1) has a solution $ (c, v)\in \mathbb{R}^m \times C(\mathbb{T}^n)^m $.
Theorem 19. Under the same hypotheses of Theorem 18, let $ (c, v_0)\in \mathbb{R}^m \times C(\mathbb{T}^n)^m $ be a solution of (5.1) and let $ v^ \lambda $ be the unique solution of $\left({{{\rm{P}}_\lambda }} \right)$ for $ \lambda > 0 $. Then there exists a function $ v^0\in C(\mathbb{T}^n)^m $ such that the functions $ v^ \lambda+(\lambda I+B)^{-1}c $ converge to $ v^0 $ uniformly on $ \mathbb{T}^n $ as $ \lambda\to 0 $. Moreover, the pair $ (c, v^0) $ is a solution of (5.1).
Proof. It is well-known (and easily checked) that due to the monotonicity of $ B $, $ (\lambda I+B) $ is invertible for any $ \lambda > 0 $. We set $ \widetilde H(x, p) = H(x, p)-c $ for $ (x, p)\in \mathbb{T}^n \times \mathbb{R}^n $ and also $ w^ \lambda(x) = v^ \lambda(x)+(\lambda I+B)^{-1}c $ for $ x\in \mathbb{T}^n $. Observe that, in the viscosity sense,
$ λwλ(x)+Bwλ(x)+˜H[wλ]=λvλ+Bvλ+H[vλ]−c+λ(λI+B)−1c+B(λI+B)−1c=0 in Tn. $
|
It is clear that $ \widetilde H $ satisfies (H) and (C) and that $ v_0 $ is a solution of $ Bu+ \widetilde H[u] = 0 $ in $ \mathbb{T}^n $. By Theorem 9, we conclude that there exists a solution $ v^0\in C(\mathbb{T}^n)^m $ of $ Bu+ \widetilde H[u] = 0 $ in $ \mathbb{T}^n $ such that $ w^ \lambda \to v^0 $ in $ C(\mathbb{T}^n)^m $ as $ \lambda\to 0+ $. Noting that $ (c, v^0) $ is a solution of (5.1), we finish the proof.
For the proof of Theorem 18, we begin with a preliminary remark on the permutations.
For a given permutation $ \pi \, :\, \mathbb{I} \to \mathbb{I} $, we define the $ m \times m $ matrix $ P $ by
$ P=(δπ(i),j)i,j∈I, $
|
(6.1) |
where $ \delta_{ij} = \delta_{i, j}: = 1 $ if $ i = j $ and $ = 0 $ otherwise. Note that $ P^{-1} = (\delta_{i, \pi(j)})_{i, j\in \mathbb{I}} = P^ \mathrm{T} $ and that for any $ u = (u_i)_{i\in \mathbb{I}} $,
$ Pu = P (u1⋮um) = (uπ(1)⋮uπ(m)) . $
|
The system of Hamilton-Jacobi equations
$ λu+Bu+H[u]=0 $
|
(6.2) |
can be written component-wise as
$ \lambda u_{\pi(i)}+(Bu)_{\pi(i)}+H_{\pi(i)}[u_{\pi(i)}] = 0 \ \ \text{ for } i\in \mathbb{I}. $ |
By the use of $ P $, the system above is expressed as
$ \lambda (Pu)_i+(PBu)_i+(PH)_{i}[(Pu)_i] = 0, $ |
and furthermore, if $ v = Pu $,
$ λ(v)i+(PBPTv)i+(PH)i[vi]=0. $
|
(6.3) |
Set $ A = (a_{ij})_{i, j\in \mathbb{I}} = PBP^ \mathrm{T} $ and observe that if $ B $ is monotone, then
$ a_{ij} = \sum\limits_{k,l\in \mathbb{I}} \delta_{i,\pi(k)}b_{kl} \delta_{\pi(l),j} = b_{\pi^{-1}(i),\pi^{-1}(j)} {≥0 if i=j,≤0 if i≠j, $
|
and
$ \sum\limits_{j\in \mathbb{I}}a_{ij} = \sum\limits_{j\in \mathbb{I}}b_{\pi^{-1}(i),\pi^{-1}(j)} = \sum\limits_{j\in \mathbb{I}}b_{\pi^{-1}(i),j}\geq 0. $ |
Consequently, if $ B $ is monotone, then $ PBP^ \mathrm{T} $ is monotone as well, and the system (6.2), by using the permutation matrix $ P $, is converted to (6.3).
Proof of Theorem 18. It is well-known (see for instance [35,Section 2.3]) that, given a monotone matrix $ B $, one can find a permutation $ \pi \, :\, \mathbb{I}\to \mathbb{I} $ such that
$ PBPT=(B(1)0⋯0∗B(2)⋱⋮⋮⋱⋱0∗⋯∗B(rp)), $
|
(6.4) |
where, $ P $ is given by (6.1), $ B^{(1)} $ is a diagonal matrix of order $ r_1 $ and, for $ 1 < i\leq p $, $ B^{(i)} $ are irreducible matrices of order $ r_i $. In view of the preliminary remark before this proof, to seek for a solution of (5.1), we may and do assume henceforth $ B $ has the normal form of the right hand side of (6.4).
Set
$ s_k = \sum\limits_{1\leq i \lt k} r_i \ \ \text{ and } \ \ \mathbb{I}_k = \{s_k+1,\ldots,s_k+r_k\} \ \ \text{ for } k\in\{1,\ldots,p\}. $ |
Notice that $ s_1 = 0 $. If $ r_1\geq 1 $, then we first show that there exist an $ r_1 $-vector $ c^{(1)} = (c^{(1)}_i)_{i\in \mathbb{I}_1}\in \mathbb{R}^{r_1} $ and a function $ v^{(1)} = (v^{(1)}_i)_{i\in \mathbb{I}_1}\in C(\mathbb{T}^n)^{r_1} $ such that $ v^{(1)} $ is a solution of
$ B(1)v(1)+H(1)[v(1)]=c(1) in Tn, $
|
(6.5) |
where $ H^{(1)} = (H_i)_{i\in \mathbb{I}_1} $. The system is, in fact, a collection of single equations
$ biiv(1)i+H(1)i[v(1)i]=c(1)i in Tn, with i∈I1, $
|
(6.6) |
and thus the existence of a solution $ (c^{(1)}, v^{(1)}) $ of (6.5) is a classical result. Indeed, for each $ i\in \mathbb{I}_1 $, if $ b_{ii}^{(1)} > 0 $, then (6.6) has a (unique) solution $ v_i^{(1)}\in \operatorname{Lip}(\mathbb{T}^n) $ for any choice of $ c_i^{(1)} $. If $ b_{ii}^{(1)} = 0 $, then (6.6) has a solution $ (c_i^{(1)}, v_i^{(1)})\in \mathbb{R} \times \operatorname{Lip}(\mathbb{T}^n) $ (see unpublished work by Lions PL, Papanicolaou G, and Varadhan S: Homogenization of Hamilton-Jacobi equations). If $ r_1 = m $, then we are done.
Next, assume that $ r_1 < m $ (and equivalently, $ 1 < p $) and we show that there exist a vector $ c^{(2)} = (c^{(2)}_i)_{i\in \mathbb{I}_2}\in \mathbb{R}^{r_2} $ and a function $ v^{(2)} = (v^{(2)}_i)_{i\in \mathbb{I}_2}\in C(\mathbb{T}^n)^{r_2} $ such that $ v^{(2)} $ is a solution of the system
$ B(2)v(2)+H(2)[v(2)]=c(2) in Tn, $
|
(6.7) |
where
$ H(2)i(x,p)=Hi(x,p)−∑j∈I1bi,jv(1)j(x) for i∈I2. $
|
(6.8) |
According to Proposition 15, there exist $ c^{(2)} = (c^{(2)}_i)_{i\in \mathbb{I}_2}\in \mathbb{R}^{r_2} $ and $ v^{(2)} = (v^{(2)}_i)_{i\in \mathbb{I}_2}\in C(\mathbb{T}^n)^{r_2} $ which satisfy (6.7). This way (by induction), we find $ c^{(1)}, \ldots, c^{(p)} $ and $ v^{(1)}, \ldots, v^{(p)} $ such that
$ c^{(k)}\in \mathbb{R}^{r_k} \ \ \text{ and } \ \ v^{(k)}\in C( \mathbb{T}^n)^{r_k} \ \ \text{ for } k\in\{1,\ldots,p\}, $ |
and $ v^{(k)} $ satisfies
$ B(k)v(k)+H(k)[v(p)]=c(k) in Tn, for k∈{1,…,p}. $
|
(6.9) |
where
$ H(k)i(x,p)=Hi(x,p)−∑1≤j<k∑q∈Ijbi,qv(j)q(x) for i∈Ik. $
|
(6.10) |
We define $ c = (c_i)_{i\in \mathbb{I}}\in \mathbb{R}^m $ and $ v = (v_i)_{i\in \mathbb{I}}\in C(\mathbb{T}^n)^m $ by setting
$ c_i = c_i^{(k)} \ \ \text{ and } \ \ v_i = v_i^{(k)} \ \ \text{ for } i\in \mathbb{I}_k,\, k\in\{1,\ldots,p\}, $ |
and observe that
$ Bv+H[v] = c \ \ \text{ in } \mathbb{T}^n. $ |
This completes the proof.
The author would like to thank the anonymous referee for useful and critical comments on the original version of this paper, which have helped significantly to improve the presentation. This work is partially supported by the JSPS KAKENHI #16H03948, #18H00833, #20K03688, and #20H01817.
The author declares no conflicts of interest in this paper.
[1] | Cruz AA, Mantzouranis E, Matricardi PM, et al. Global surveillance, prevention and control of Chronic Respiratory Diseases—A comprehensive approach. World Health Organization; 2007. |
[2] | Afshari A, Anderson HR, Cohen A, et al. Dampness and Mould. WHO Regional Office for Europe; 2009. |
[3] | Reponen T, Vesper S, Levin L, et al. (2011) High environmental relative moldiness index during infancy as a predictor of asthma at 7 years of age. Ann Allerg Asthma Im 107: 120-126. |
[4] | Dannemiller KC, Mendell MJ, Macher JM, et al. (2014) Next generation DNA sequencing reveals that low fungal diversity in house dust is associated with childhood asthma development. Indoor Air 24: 236-247. |
[5] | Behbod B, Sordillo JE, Hoffman EB, et al. (2013) Wheeze in infancy: protection associated with yeasts in house dust contrasts with increased risk associated with yeasts in indoor air and other fungal taxa. Allergy 68: 1410-1418. |
[6] | Lötvall J, Akdis CA, Bacharier LB, et al. (2011) Asthma endotypes: a new approach to classification of disease entities within the asthma syndrome. J Allergy Clin Immunol 127: 355-360. |
[7] | Radon K (2006) The two sides of the “endotoxin coin”. Occup Environ Med 63: 73-78. |
[8] | Hamilos DL (2010) Allergic fungal rhinitis and rhinosinusitis. Proc Am Thora Soc 7: 245-252. |
[9] | Pringle A (2013) Asthma and the diversity of fungal spores in air. Plos Pathogens 9: e1003371. |
[10] | Denning DW, Pashley C, Hartl D, et al. (2014) Fungal allergy in asthma-state of the art and research needs. Clinical and Translational Allergy 4: 14. |
[11] | Vesper S, Barnes C, Ciaccio CE, et al. (2013) Higher environmental relative moldiness index (ERMI) values measured in homes of asthmatic children in Boston, Kansas City, and San Diego. J Asthma 50: 155-161. |
[12] | Reponen T, Lockey J, Bernstein DI, et al. (2012) Infant origins of childhood asthma associated with specific molds. J Allergy Clin Immunol 130: 639-644. |
[13] | O’Connor GT (2005) Allergen avoidance in asthma: What do we do now? J Allergy Clin Immunol 116: 26-30. |
[14] | Tovey E, Ferro A (2012) Time for new methods for avoidance of house dust mite and other allergens. Curr allergy asthma rep 12: 465-477. |
[15] | Burge HA, Solomon WR, Muilenberg ML (1982) Evaluation of indoor plantings as allergen exposure sources. J Allergy Clin Immunol 70: 101-108. |
[16] | Twaroch TE, Curin M, Valenta R, et al. (2015) Mold allergens in respiratory allergy: from structure to therapy. Allergy Asthma Immunol Res 7: 205-220. |
[17] | Trail F (2007) Fungal cannons: explosive spore discharge in the Ascomycota. FEMS Microbiol Lett 276: 12-18. |
[18] | Frankel M, Hansen EW, Madsen AM (2014) Effect of relative humidity on the aerosolization and total inflammatory potential of fungal particles from dust-inoculated gypsum boards. Indoor Air 24: 16-28. |
[19] | Madsen AM (2012) Effects of air flow and changing humidity on the aerosolisation of respirable fungal fragments and conidia of Botrytis cinerea. Appl Environ Microbiol 78: 3999-4007. |
[20] | Górny RL, Reponen T, Grinshpun SA, et al. (2001) Source strength of fungal spore aerosolization from moldy building material. Atm Environ 35: 4853-4862. |
[21] | Pasanen A-L, Pasanen P, Jantunen MJ, et al. (1991) Significance of air humidity and air velocity for fungal spore release into the air. Atm Environ 25a: 459-462. |
[22] | Zoberi MH (1961) Take-off of mould spores in relation to wind speed and humidity. Annals of Botany 25: 53-64. |
[23] | Eduard W (2009) Fungal spores: a critical review of the toxicological and epidemiological evidence as a basis for occupational exposure limit setting. Crit Rev Toxicology 39: 799-864. |
[24] | Sigler L, Abbott SP, Gauvreau H (1996) Assessment of worker exposure to airborne molds in honeybee overwintering facilities. Am Ind Hyg Assoc J 57: 484-490. |
[25] | Madsen AM, Tendal K, Frederiksen MW (2014) Attempts to reduce exposure to fungi, ß-glucan, bacteria, endotoxin and dust in vegetable greenhouses and a packaging unit. Sci Total Environ 468: 1112-1121. |
[26] | Visser MJ, Spaan S, Arts HJ, et al. (2006) Influence of different cleaning practices on endotoxin exposure at sewage treatment plants. Ann Occup Hyg 50: 731-736. |
[27] | Madsen AM. Identification of Work Tasks Causing High occupational Exposure to Bioaerosols at Biofuel plants Converting Straw or Wood Chips. In: Dr.Marco Aurelio DosSantos Bernardes, editor. Environmental Impact of biofuels.InTech; 2011. 251-270. |
[28] | Simon-Nobbe B, Denk U, Pöll V, et al. (2008) The spectrum of fungal allergy. Int Arch Allergy Immunol 145: 58-86. |
[29] | Hunter CA, Grant C, Flannigan B, et al. (1988) Mould in buildings: the air spora of domestic dwellings. Inte Biodeter 24: 81-101. |
[30] | Reponen T, Lehtonen M, Raunemaa T, et al. (1992) Effect of indoor sources on fungal spore concentrations and size distributions. J Aerosol Sci 23: 663-666. |
[31] | Lehtonen M, Reponen T, Nevalainen A (1993) Everyday activities and variation of fungal spore concentrations in indoor air. Int Biodeterior Biodegradation 31: 25-39. |
[32] | Buttner MP, Stetzenbach LD (1993) Monitoring airborne fungal spores in an experimental indoor environment to evaluate sampling methods and the effects of human activity on air sampling. Appl Environ Microbiol 59: 219-226. |
[33] | Buttner MP, Cruz-Perez P, Stetzenbach LD, et al. (2002) Measurement of airborne fungal spore dispersal from three types of flooring materials. Aerobiologia 18: 1-11. |
[34] | Millington WM, Corden JM (2005) Long term trends in outdoor Aspergillus/Penicillium spore concentrations in Derby, UK from 1970 to 2003 and a comparative study in 1994 and 1996 with the indoor air of two local houses. Aerobiologia 21: 105-113. |
[35] | Veillette M, Knibbs LD, Pelletier A, et al. (2013) Microbial Contents of Vacuum Cleaner Bag Dust and Emitted Bioaerosols and Their Implications for Human Exposure Indoors. Appl Environ Microbiol 79: 6331-6336. |
[36] | Hegarty JM, Rouhbakhsh S, Warner JA, et al. (1995) A comparison of the effect of conventional and filter vacuum cleaners on airborne house dust mite allergen. Respiratory Medicine 89: 279-284. |
[37] | Siracusa A, De BF, Folletti I, et al. (2013) Asthma and exposure to cleaning products - a European Academy of Allergy and Clinical Immunology task force consensus statement. Allergy 68: 1532-1545. |
[38] | Bernstein JA, Brandt D, Rezvani M, et al. (2009) Evaluation of cleaning activities on respiratory symptoms in asthmatic female homemakers. Ann Allerg Asthma Im 102: 41-46. |
[39] | Mäkelä R, Kauppi P, Suuronen K, et al. (2011) Occupational asthma in professional cleaning work: a clinical study. Occup Med 61: 121-126. |
[40] | Balasubramanian R, Nainar P, Rajasekar A (2012) Airborne bacteria, fungi, and endotoxin levels in residential microenvironments: a case study. Aerobiologia 28: 375-390. |
[41] | Summerbell RC, Krajden S, Kane J (1989) Potted plants in hospitals as reservoirs of pathogenic fungi. Mycopathologia 106: 13-22. |
[42] | Summerbell RC, Staib F, Dales R, et al. (1992) Ecology of fungi in human dwellings. J Med Vet Mycol 30: 279-285. |
[43] | Goebes MD, Boehm AB, Hildemann LM (2011) Contributions of foot traffic and outdoor concentrations to indoor airborne Aspergillus. Aerosol Sci Tech 45: 352-363. |
[44] | Kildesø J, Vinzents P, Schneider T, et al. (1999) A simple method for measuring the potential resuspension of dust from carpets in the indoor environment. Textile Res J 69: 169-175. |
[45] | Chen Q, Hildemann LM (2009) The effects of human activities on exposure to particulate matter and bioaerosols in residential homes. Environ Sci Technol 43: 4641-4646. |
[46] | Adams RI, Miletto M, Taylor JW, et al. (2013) Dispersal in microbes: fungi in indoor air are dominated by outdoor air and show dispersal limitation at short distances. The ISME journal 7: 1262-1273. |
[47] | Takahashi T (1997) Airborne fungal colony-forming units in outdoor and indoor environments in Yokohama, Japan. Mycopathologia 139: 23-33. |
[48] | Mitakakis TZ, Tovey ER, Xuan W, et al. (2000) Personal exposure to allergenic pollen and mould spores in inland New South Wales, Australia. Clin Exp Allergy 30: 1733-1739. |
[49] | Sessa R, Di PM, Schiavoni G, et al. (2002) Microbiological indoor air quality in healthy buildings. The New Microbiologica 25: 51-56. |
[50] | Madsen AM, Matthiesen CB, Frederiksen MW, et al. (2012) Sampling, extraction and measurement of bacteria, endotoxin, fungi and inflammatory potential of settling indoor dust. J EnvironMonitor 14: 3230-3239. |
[51] | Scheff PA, Paulius VK, Curtis L, et al. (2000) Indoor air quality in a middle school, Part II: Development of emission factors for particulate matter and bioaerosols. Applied Occupational and Environmental Hygiene 15: 835-842. |
[52] | Brandl H, von Däniken A, Hitz C, et al. (2008) Short-term dynamic patterns of bioaerosol generation and displacement in an indoor environment. Aerobiologia 24: 203-209. |
[53] | Qian J, Hospodsky D, Yamamoto N, et al. (2012) Size-resolved emission rates of airborne bacteria and fungi in an occupied classroom. Indoor Air 22: 339-351. |
[54] | Hospodsky D, Yamamoto N, Nazaroff WW, et al. (2014) Characterizing airborne fungal and bacterial concentrations and emission rates in six occupied children’s classrooms. Indoor Air doi: 10.1111/ina.12172. |
[55] | Yamamoto N, Hospodsky D, Dannemiller KC, et al. (2015) Indoor emissions as a primary source of airborne allergenic fungal particles in classrooms. Environ Sci Technol 49: 5098-5106. |
[56] | Bush RK, Portnoy JM (2001) The role and abatement of fungal allergens in allergic diseases. J Allergy Clin Immunol 107: S430-S440. |
[57] | Levetin E. Fungi. In: Burge HA, editor. Bioaerosols.Lewis Publishers; 1995, 87-120. |
[58] | Sharpe RA, Bearman N, Thornton CR, et al. (2015) Indoor fungal diversity and asthma: a meta-analysis and systematic review of risk factors. J Allergy Clin Immunol 135: 110-122. |
[59] | Reponen T, Lockey J, Bernstein DI, et al. (2012) Infant origins of childhood asthma associated with specific molds. J Allergy Clin Immunol 130: 639-644. |
[60] | Gravesen S (1985) Indoor airborne mould spores. Allergy 40: 21-23. |
[61] | Ren P, Jankun TM, Leaderer BP (1999) Comparisons of seasonal fungal prevalence in indoor and outdoor air and in house dusts of dwellings in one Northeast American county. J Expo Anal Environ Epidemiol 9: 560-568. |
[62] | Chew GL, Rogers C, Burge HA, et al. (2003) Dustborne and airborne fungal propagules represent a different spectrum of fungi with differing relations to home characteristics. Allergy 58: 13-20. |
[63] | Mandal J, Brandl H (2011) Bioaerosols in Indoor Environment-A Review with Special Reference to Residential and Occupational Locations. Open Environ Biol Monit J 4: 83-96. |
[64] | D’Amato G, Spieksma FThM (1995) Aerobiologic and clinical aspects of mould allergy in Europe. Allergy 50: 870-877. |
[65] | Rosas I, Calderón C, Martínez L, et al. (1997) Indoor and outdoor airborne fungal propagule concentrations in Mexico City. Aerobiologia 13: 23-30. |
[66] | Kozak PP, Gallup J, Cummins LH, et al. (1979) Factors of importance in determining the prevalence of indoor molds. Ann Allergy 43: 88-94. |
[67] | Amend AS, Seifert KA, Samson R, et al. (2010) Indoor fungal composition is geographically patterned and more diverse in temperate zones than in the tropics. Proc Natl Acad Sci 107: 13748-13753. |
[68] | Gots RE, Layton NJ, Pirages SW (2003) Indoor health: background levels of fungi. AIHA J 64: 427-438. |
[69] | Larsen L, Gravesen S (1991) Seasonal variation of outdoor airborne viable microfungi in Copenhagen, Denmark. Grana 30: 467-471. |
[70] | Frankel M, Beko G, Timm M, et al. (2012) Seasonal variation of indoor microbial exposures and their relations to temperature, relative humidity and air exchange rates. Appl Environ Microbiol 78: 8289-8297. |
[71] | Stetzenbach LD, Buttner MP, Cruz P (2004) Detection and enumeration of airborne biocontaminants. Curr Opin Biotechnol 15: 170-174. |
[72] | Madsen AM (2006) Exposure to airborne microbial components in autumn and spring during work at Danish biofuel plants. Ann Occup Hyg 50: 1-11. |
[73] | Licorish K, Novey HS, Kozak P, et al. (1985) Role of Alternaria and Penicillium Spores in the Pathogenesis of Asthma. J Allergy Clin Immunol 76: 819-825. |
[74] | Bagni N, Davies RR, Mallea M, et al. (1977) [Spore concentration in cities of the European Economic Community. II. Spores of Cladosporium and Alternaria]. Acta Allergologica 32: 118-138. |
[75] | Gravesen S (1979) Fungi as a Cause of Allergic Disease. Allergy 34: 135-154. |
[76] | Garrett MH, Rayment PR, Hooper MA, et al. (1998) Indoor airborne fungal spores, house dampness and associations with environmental factors and respiratory health in children. Clin Exp Allergy 28: 459-467. |
[77] | Adhikari A, Lewis JS, Reponen T, et al. (2010) Exposure matrices of endotoxin, (1-->3)-beta-d-glucan, fungi, and dust mite allergens in flood-affected homes of New Orleans. Sci Total Environ 408: 5489-5498. |
[78] | Reponen T, Nevalainen A, Jantunen M, et al. (1992) Normal range criteria for indoor air bacteria and fungal spores in a subarctic climate. Indoor Air 2: 26-31. |
[79] | Sharpe R, Thornton CR, Osborne NJ (2014) Modifiable factors governing indoor fungal diversity and risk of asthma. Clin Exp Allergy 44: 631-641. |
[80] | Micheluz A, Manente S, Tigini V, et al. (2015) The extreme environment of a library: Xerophilic fungi inhabiting indoor niches. Int Biodeterior Biodegradation 99: 1-7. |
1. | Jakob Schneider, Ksenia Korshunova, Zeineb Si Chaib, Alejandro Giorgetti, Mercedes Alfonso-Prieto, Paolo Carloni, Ligand Pose Predictions for Human G Protein-Coupled Receptors: Insights from the Amber-Based Hybrid Molecular Mechanics/Coarse-Grained Approach, 2020, 60, 1549-9596, 5103, 10.1021/acs.jcim.0c00661 | |
2. | Stefano Capaldi, Eda Suku, Martina Antolini, Mattia Di Giacobbe, Alejandro Giorgetti, Mario Buffelli, Allosteric sodium binding cavity in GPR3: a novel player in modulation of Aβ production, 2018, 8, 2045-2322, 10.1038/s41598-018-29475-7 | |
3. | Jinan Wang, Yinglong Miao, 2019, 116, 9780128155615, 397, 10.1016/bs.apcsb.2018.11.011 | |
4. | Filippo Baldessari, Riccardo Capelli, Paolo Carloni, Alejandro Giorgetti, Coevolutionary data-based interaction networks approach highlighting key residues across protein families: The case of the G-protein coupled receptors, 2020, 18, 20010370, 1153, 10.1016/j.csbj.2020.05.003 |