Brought to you by:
Paper

Mathematical aspects of catalyst positioning in lithium/air batteries

, , and

Published 25 February 2020 © 2020 IOP Publishing Ltd
, , Citation Thuong-Huyen Nguyen et al 2020 Inverse Problems 36 044001 DOI 10.1088/1361-6420/ab47e6

0266-5611/36/4/044001

Abstract

In this paper we investigate a system of partial differential equations describing the clogging process of non-aqueous lithium/air batteries. The main aim is to optimize the positioning of catalysts on the surface of the pore of the battery in order to maximize the efficiency of the battery. To this end we analyze the parameter-to-state map which maps the initial catalyst positions cat0(x) to the diminishing radius of the pore until clogging. We obtain results on the continuity and Fréchet differentiability of the parameter-to-state map and we obtain the related sensitivity system. The paper also includes numerical experiments, which compare a greedy approach to catalyst positioning with results obtained by an analytic optimisation procedure.

Export citation and abstract BibTeX RIS

1. Introduction

Portable electronic devices play important roles in our daily modern life. It is imperative to investigate efficient and environmentally friendly ways of storing and delivering energy for such devices. Lithium/air batteries recently attract great attention of many engineers and scientists because of their potential for higher energy values in comparison with other metal/air batteries. Moreover, cathodes with a porous structure are good candidates for lighter power sources for aerial vehicles or camping equipments compared with other battery types [21]. There are basically two kinds of lithium/air batteries: aqueous and non-aqueous. The non-aqueous type differs from the aqueous lithium/air batteries in that the discharged products are insoluble in the aqueous electrolytes [1, 18, 19], which leads to the pore's passivation. However, the aqueous lithium/air batteries require an isolated protection layer between metallic lithium and the electrolyte, that will significantly limit the Li+ -conductivity. Hence, most of the literatures on lithium/air batteries is devoted to the non-aqueous electrolyte batteries, which is also the model under consideration for this paper. For a detailed survey on the capacities of aqueous, resp. non-aqueous, Li/air and Li-ion batteries we refer to Christensen et al [3].

Non-aqueous batteries are based on an oxygen reduction reaction (ORR) at the cathode surface during the discharge process. Here, oxygen is consumed by its reaction with lithium ions thus generating the insoluble lithium peroxide, see Laoire et al [13, 14]. There are a number of fundamental and practical problems requiring to be considered when designing batteries: the role of some specific catalysts in promoting the cathode electrode reaction, optimization of the electrode porosity, the structure of the porous cathode, the composition of the cathode electrode, the prevention of water or carbon dioxide from entering the cell when operated in air [4, 21] and many more. According to [1, 18, 19], a critical difficulty is that the discharged oxygen reduction product is insoluble in the organic electrolytes. The low oxygen solubility causes the deposition of the discharged products, see [4, 21] for a diffusion-limited model and [19] for a steady-state model. This eventually clogs the oxygen entrance to the pore and limits the capacity of the batteries by narrowing the active surface inside the pore.

In this paper, we aim at maximizing the capacity of non-aqueous batteries by optimal catalyst positioning. To this end we will analyse a mathematical-physical model describing the clogging process. The overall optimization problem asks to determine an initial catalyst distribution cat0 which allows to maximize the utilization of the pore volume during the oxygen reduction process. Our motivation is taken from the work of Dabrowskji et al [4], where the authors focus on optimizing and positioning of catalytically active sites within the pores to maximize the volume of discharge products before pore clogging. There, the authors propose a dynamic but discrete catalyst model for the discharge process, where the catalytic function ${\rm cat}(t,x)$ can take values 0 or 1 depending on the initial catalyst distribution cat0 and also on the radius of the deposited oxygen reduction products nearby.

This paper is organized as follows: our main interest in this work is to approach the catalyst positioning problem by proposing a continuous propagation process with an analytic catalytic model. This will lead to a system of partial differential equations (PDEs) which links the catalyst positions cat0 at time t  =  0 to the consumption of oxygen leading to a decreasing oxygen concentration $c(t,x)$ and to a diminishing pore radius $r(t,x)$ . The resulting mathematical models for a stationary catalytic function ${\rm cat}(t,x)={\rm cat}_0(x)$ as well as for a dynamic ${\rm cat}(t,x)$ are introduced in the next section.

The subsequent mathematical analysis in section 3 is restricted to the stationary case. We start by smoothing the PDEs system in order to avoid singularities in the coefficients of the PDE system. This modification allows us to prove existence and uniqueness of a solution as well as determining continuity and Fréchet differentiability of the parameter-to-state map, which links cat0 to the pore radius r. Section 4 then contains an analysis of the optimization problem including a derivation of the adjoint system for $r_{{\rm cat}_0}^*$ . The final section contains numerical experiments for comparing a standard greedy method for catalysts positioning with more advanced approaches. Here we focus on numerical examples for the dynamic case.

2. Mathematical setting

2.1. Mathematical setting of the diffusion-limited model for lithium/air batteries

The main factor limiting the efficiency of non-aqueous lithium/air batteries is the deposition of lithium-dioxide during the reduction process, which leads to a narrowing and finally clogging of the pores of the batteries, for some theoretical and experimental research on this topic see [19, 22]. In this paper we aim at optimizing the initial placement of catalysts at the pore surface in order to optimize the efficiency of such batteries. In this context efficiency is measured in terms of the remaining free pore volume after clogging.

Our starting point is the approach of [4, 21], where a system of non-linear differential equations has been introduced for describing the process of discharging and clogging in a single pore. The basic parameter in these models is the initial distribution cat0(x) of catalysts along the surface of the pore. These previous approaches assumed, that either cat0 is constant 1 [21], or that cat0 can take only values 1 or 0 indicating the presence of a catalyst at certain positions.

Our first target, see Problem I below, is to state a continuous model for cat0 and to analyse the properties of the resulting parameter-to-state mapping (existence, uniqueness, Fréchet derivative, sensitivity system, adjoint operator) as a prerequisite to attacking the optimization problem in section 4. Hence, following [4, 21] we define $\Omega=[0,1]$ and $Q_T = [0,T] \times \Omega$ as well as $c: Q_T \rightarrow [0,\infty)$ as the oxygen concentration at time t at position x of a normalized pore. By $r: Q_T \rightarrow [0,\infty)$ we denote the radius of the pore, which diminishes over time and finally leads to pore clogging. The first catalyst model assumes a catalyst distribution cat$_0: \Omega \rightarrow I\!R$ , which is constant over time (Problem I):

Equation (2.1a)

Equation (2.1b)

Equation (2.1c)

Equation (2.1d)

Equation (2.1e)

where $\beta=\beta(t,c,r)$ and $\gamma=\gamma(t,c,r)$ are given by equations (2.2) and (2.3). For the initial oxygen concentration c at time t  =  0, we assume the compatibility conditions $c(0,0)=h(0)=1$ and $c(0,1)=h(1)=0$ , which models that at t  =  0 oxygen is entering the pore at x  =  0 but has not yet reached the end of the pore at x  =  1.

Remark 2.1. Following [21], the mathematical model (2.1a)–(2.1e) can be justified by considering the lithium/air reduction process in a cylindrical pore of radius $r_p^0$ and length l, a porous cathode of volume $V_{\rm cath}$ , and given initial porosity $ \newcommand{\e}{{\rm e}} \epsilon_0$ . The authors introduce a coordinate system where tp, resp. xp, denotes time, resp. spatial position along the central axis of the pore. Furthermore, assume that the oxygen concentration $C_{\rm O_2}$ , which is a function of tp and xp, is angularly symmetric about the diffusion direction, which is identical with the radial direction along the pore. The initial radius of the pore is given by $r_p^0$ , it diminishes during the reduction process, its value at position xp at time tp is denoted by rp.

Let $C_{\rm O_{2},0}= 2.10\cdot 10^{-6}~{\rm (mol~cm^{-3})}$ denote the initial oxygen concentration and $D_{\rm O_2}= 16.70\cdot 10^{-6}~{\rm (cm^2~s^{-1})}$ the oxygen mass diffusivity. This allows to introduce the dimensionless quantities $ \newcommand{\e}{{\rm e}} t=\frac{t_pD_{\rm O_2}}{\sqrt{\epsilon_0} \ l^2}$ , $r=\frac{r_p}{r_p^0}$ , $x=\frac{x_p}l$ , $c=\frac{C_{\rm O_2}}{C_{{\rm O_2},0}}$ , i.e. $0\leqslant r,c,x \leqslant 1$ , 0  <  t  <  T. The above stated model is obtained by introducing parameters $\gamma$ and $\beta$ given by

Equation (2.2)

with

Equation (2.3)

We note that $\beta$ and $\gamma$ depend on k, whose definition involves an integral over c and r, hence they are functions of time t. In order to highlight this non-linear dependence we use the full notation $\beta=\beta(t, c,r)$ and $\gamma = \gamma (t,c,r)$ whenever suitable.

The other parameters are set as follows (see [4, 21]): $F=9.648\,533\,65 \cdot 10^7$ (Faraday constant), $r_{p}^0 = 2.0 \times 10^{-7}$ cm; l  =  0.07 cm; $ \newcommand{\e}{{\rm e}} \epsilon_{0} = 0.73$ ; $M_{\rm Li_2O_2} = 45.8765~{\rm g~mol^{-1}}$ ; ${\rho}_{\rm Li_2O_2} = 2.3~{\rm g~cm^{-3}}$ ; ${i}_{\rm geom} = 0.1~{\rm mA~cm^{-2}}$ .

Note that [4] discusses different exponents in the reaction equation (2.1a), namely,

Equation (2.4)

In this case, the dimensionless time is $ \newcommand{\e}{{\rm e}} t=\frac{t_pD_{\rm O_2}\sqrt{\epsilon_0}}{l^2}$ and parameters $\gamma$ , $\beta$ are defined by

Equation (2.5)

with

Equation (2.6)

In the following, we will state the mathematical analysis for the model (2.1a); however we emphasize, that the same methodology can be used to analyse the model (2.4) as well. For more details and the parameters concerned, we refer the readers to [4, 21].

Problem I is our basic model, which will be analysed as a parameter identification problem in the following sections. Finally, we aim to optimize the initial catalyst positioning cat0, which we regard a major design decision for optimizing non-aqueous lithium/air batteries. In order to stress this dependence we use the notation $r({\rm cat}_0;t,x)$ as well as $c({\rm cat}_0;t,x)$ . Depending on the context we will sometimes eliminate this dependence and simply use e.g. $r({\rm cat}_0;t,x)=r(t,x)=r$ .

Numerical tests and simulations we will be based on an extended model by making ${\rm cat}={\rm cat}(t,x)$ a time dependent function with ${\rm cat}(0,\cdot)={\rm cat}_0$ . This is in agreement with experiments and yields a more realistic model. Such a dynamic approach has already been incorporated in [4], which used a simple discretized evolution based on thresholding: assume that the initial catalysts distribution cat0 takes value 1 at position x0 but is 0 at neighboring discretization points; the deposition will start at point x0 only; the value of the cat-function at neighboring points is changed to 1 whenever r(x0) reaches a critical value. In the present paper we use a different approach, where the system of differential equations for describing the evolution of the pore radius r and the oxygen concentration c is expanded by a simple diffusion model for the time evolution of ${\rm cat}(t,x)$ . Learning from growth kinetics, see e.g. [4] an the references there, the evolution of cat is based on incorporating a threshold model ($T_{R_0}$ ) in combination with a generic diffusion model (Gaussian). The threshold function models the fact, that once the radius has decreased below the level R0 then the catalytic action is increased and spreads out to neighboring positions. Hence Problem II is obtained from Problem I by adapting the time derivative for r

Equation (2.7)

and by adding the following evolution for cat

Equation (2.8)

and

Equation (2.9)

where $\alpha_0$ is some small positive constant, R0 is a positive constant, 0  <  R0  <  1. Here $ \newcommand{\e}{{\rm e}} G(x)=\frac{1}{2\pi}\exp(-\frac12\,x^2) $ denotes the Gaussian distribution and $T_{\chi,g,R_0}(r)$ is a C1-threshold function given by

Equation (2.10)

For convenience, we denote $T_{\chi,g,R_0}(s)$ by $T_{R_0}(s)$ . The constant $\chi >0$ should be large, resp. g  >  0 should be small, such that $-\frac{g}{2\chi}+\frac{1}{2\chi^2} < \frac{g}{2\chi}-\frac{1}{2\chi^2}$ . The choice of these constants influences the 'sloppy'-ness of $T_{\chi,g,R_0}(s)$ in $[-\frac{g}{2\chi}-\frac{1}{2\chi^2}+R_0,\frac{g}{2\chi}+\frac{1}{2\chi^2}+R_0]$ , see the figure below.

2.2. Optimization problems for catalyst positioning models

Our final aim is to minimize the free volume of the pore or to maximize the volume of deposited products before pore clogging. We do consider two optimization problems. In the first approach, we consider the minimization problem of the remaining pore volume at time $T, 0 < T < \infty $ , where T is the time of clogging or a fixed approximation of it.

2.2.1. The first optimization problem.

For Problems I and II, the free volume $V({\rm cat}_0)$ of the pore at time T is defined by:

Equation (2.11)

The first optimization problem aims at minimizing the free volume $V$ with respect to cat0 over some admissible set.

2.2.2. The second optimization problem.

The second optimization problem aims at minimizing the mean value of free volume in the pore in a small interval of time [T1,T] which is defined by

Equation (2.12)

for Problems I and II.

This problem avoids point evaluations in time and is somewhat easier to analyse.

2.2.3. Minimization by gradient descent.

One of the numerical schemes, which will be used in the section 5, will be a gradient descent method for the functional $\mathcal{V}$ , i.e.

This approach follows the classical strategy for parameter identification and optimization, as e.g. described in [6, 1012, 15, 16]. The computation of $\partial {\mathcal{V}}({\rm cat}_0^n)$ requires to determine the Fréchet derivative of the parameter-to-state mapping $r({\rm cat}_0; \cdot , \cdot)$ with respect to cat0 as well as the computation of its adjoint. This will be addressed in the following section.

3. Well-posedness of the smoothed mathematical models for continuous catalyst positioning in lithium/air batteries

In this section we study the parameter-to-state map of Problem I. In order to avoid singularities in the coefficient functions $\beta$ and $\gamma$ , we use a slightly modified model. The main reason for introducing this smoothing is to avoid singularities as the concentration c tends to zero. In the limit this changes the nature of the system of the differential equations in such a way that the theoretical results on the smoothness of its solutions no longer hold, i.e. applying the comparison principle for the weak solutions no longer works. At least we did not find applicable results in the available literature on PDE systems which cover the special singularity of our system.

From the engineering point of view, this is not a strong restriction. The change in concentration is a relative reduction, i.e. a non-zero concentration at the beginning yields non-zero concentrations for all times. Hence a complete stop of oxygen flow only occurs after pore blocking, which would yield c  =  0 at positions behind the blocking point. However, the efficiency of the battery drops significantly before that and batteries are typically replaced sometime before blocking. Hence, assuming, that there is always a minimal flow of oxygen is natural for any model, which covers the time span $[0,T]$ , where T is slightly smaller than the final time of pore blocking.

The effect of the singularity will also be reflected by our numerical simulations presented in section 5. We stop the simulations before complete blocking and hence the modification presented in the 'Modified system for Problem I' does not affect the numerical simulations, i.e. the introduction of the modified systems primarily serves analytical consistency.

3.1. Modified system for problem I

Let $c=c(t,x) : Q_T\rightarrow \mathbb{R}$ be the oxygen concentration inside the pore and $r= r(t,x) : Q_T\rightarrow \mathbb{R}$ the pore radius. Instead of considering Problem I from (2.1a)–(2.1e) with $\beta$ and $\gamma$ defined by (2.2), we study a smoothed version by considering the system of PDEs given in (2.1a)–(2.1e) with the following definitions of $\beta$ and $\gamma$ :

Equation (3.1)

with $ \newcommand{\e}{{\rm e}} \epsilon$ being a small positive number and $ \newcommand{\e}{{\rm e}} H_{\epsilon}(s)$ being the Huber function:

Equation (3.2)

where

Equation (3.3)

The definition of the Huber function depends on $ \newcommand{\e}{{\rm e}} \epsilon$ and $\chi$ , see figure 1, where $\chi$ should be chosen as large as possible. In fact, the numerical computations in section 5 were done with $\chi= \infty$ . Hence, for ease of notation we just call it $ \newcommand{\e}{{\rm e}} H_{\epsilon}$ .

Figure 1.

Figure 1. Left: graph of the C1-threshold function $T_{\chi,g,R_0}(x)$ with $g=1/2,\,\chi=50,\, $ $ {\rm and }\, R_0=0.94.$ Right: graph of $ \newcommand{\e}{{\rm e}} H_\epsilon(s)$ defined by (3.2) with $ \newcommand{\e}{{\rm e}} \epsilon = 0.5$ .

Standard image High-resolution image

We see that $\beta(t,r,c)$ and $\gamma(t,r,c)$ are smoothed versions of (2.2). The constants $D_1, D_2$ can be made explicit by using (2.2) and (2.3).

Remark 3.1. The Huber function $ \newcommand{\e}{{\rm e}} H_{\epsilon}$ is monotone, twice continuously differentiable, and $ \newcommand{\bbR}{\mathbb{R}} \newcommand{\e}{{\rm e}} H_{\epsilon}(s)\geqslant \epsilon, \forall s \in \bbR$ . The Huber function $ \newcommand{\e}{{\rm e}} H_\epsilon$ is introduced to avoid singularities if c approaches 0. This process of Huberization has been used e.g. in [5] for smoothing step functions.

A mathematical analysis of this system of equations is more accessible after expanding $\frac{\partial}{\partial t}(r^2c)$ and after using a standard substitution for obtaining a homogenous system.

We suppose that r and c are sufficiently smooth and $0<r_\alpha\leqslant r, \forall (t,x)\in Q_T$ , where $r_\alpha$ is some small positive constant. Then, expanding the derivatives in (2.1a), and dividing both sides by r2, we get

Equation (3.4)

To remove the non-homogeneous boundary condition (2.1e) we set

Equation (3.5)

Then we have u(t,0)  =  ux(t,1)  =  0 and equation (3.4) has the form

Equation (3.6)

Inserting modified rt from (2.1b) by $ \newcommand{\e}{{\rm e}} r_t=-\frac{D_2}{\int_0^1{r H_{\epsilon}(u+1)}{\rm d}x}{\rm cat}_0(x)(u+1)$ into the right hand side of (3.6), and denoting u0(x)  =  h(x)  −  1, we transform modified system for Problem I into its final form

Equation (3.7)

Equation (3.8)

Equation (3.9)

Equation (3.10)

Equation (3.11)

with

and

Similarly, we obtain a dynamic version of Problem II by replacing cat0(x) with ${\rm cat}(t,x)$ and by adding a differential equation for cat as in (2.7) and (2.8).

3.1.1. Uniqueness and existence.

We use the standard Sobolev spaces $H^{\,p}(\Omega)$ , $1 \leqslant p \leqslant \infty$ [7]. For a Banach space X with norm $\Vert \cdot\Vert_{X}$ we denote by Lp(0,T;X) the space consisting of all measurable functions $u: [0,T]\longrightarrow X$ with $\left\|u\right\|_{L^{\,p}{(0,T;X)}}:=\left(\int_0^T{\left\|u(t)\right\|^{\,p}_X {\rm d}t}\right){}^{\frac1p}<\infty$ for $1\leqslant p \leqslant \infty$ , and $\left\|u\right\|_{L^{\infty}(0,T;X)}:={{\rm~esssup}}_{0\leqslant t \leqslant T}{\left\|u(t)\right\|_X}< \infty.$ We define

We sometimes also use the notation $W(0,T)$ for $W(0,T;H^1(\Omega))$ .

Definition 3.2. A pair $(r,u)\in W^{1,\infty}(Q_T)\times W(0,T;H^1(\Omega)) $ is said to be a weak solution of (3.7)–(3.11) if

the following equations are satisfied:

Equation (3.12)

Equation (3.13)

Equation (3.14)

To prove that there exists a weak solution to the system (3.7)–(3.11), we will apply the Banach fixed point theorem. The proof is long and tedious but follows the classical path of arguments for proving such results for non-linear partial differential equations, hence, we present only its main steps. We define the following spaces [17]

with norm $ \newcommand{\bfX}{\mathbf{X}} \left\|r\right\|_{\bfX_T}=\left\|r\right\|_{L^{\infty}(Q_T)}+\left\|r_t\right\|_{L^2(0,T;L^\infty(\Omega))} + \left\|r_x\right\|_{L^2(0,T;L^\infty(\Omega))}, $ and

with norm $ \newcommand{\bfY}{\mathbf{Y}} \left\|u\right\|_{\bfY_T}=\left\|u\right\|_{L^{2}(0,T;H^2(\Omega))} + \left\|u\right\|_{L^{\infty}(0,T;H^1(\Omega))}+ \left\|u_t\right\|_{L^{2}(0,T;L^2(\Omega))}. $ Let $M, r_\alpha, R$ denote some fixed positive constants, we define the set

Equation (3.15)

where $0 < r_\alpha < 1 < R.$ We see that $ \newcommand{\bM}{\mathbf{M}} \bM_T$ is closed, convex and bounded in $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T$ .

First, we prove the existence of a solution $ \newcommand{\bM}{\mathbf{M}} (r,u) \in \bM_T$ for sufficiently small T. Before proving the contraction property need for applying Banach's theorem, we require a general result concerning the existence of solutions of partial differential equations. Relying on [7], we have the following result.

Theorem 3.3. Consider the initial/boundary-value problem

Equation (3.16)

where $ \newcommand{\bbR}{\mathbb{R}} f: Q \rightarrow \bbR $ and $ \newcommand{\bbR}{\mathbb{R}} u_0: \Omega \rightarrow \bbR $ are given, and $u=u(t,x)$ is unknown, the operator $ \newcommand{\calL}{\mathcal{L}} \calL$ denotes for each $t\in [0,T]$ the second-order non-divergence partial differential operator

Equation (3.17)

for given coefficients $a(t,x), b(t,x), d(t,x)$ satisfying

  • 1.  
    Equation (3.18)
  • 2.  
    $a(t,x), a_t(t,x), b(t,x), d(t,x) \in L^{\infty}(Q_T),$
  • 3.  
    $f\in L^2(0,T; L^2(\Omega))$ .

Then, there exists unique weak solution u in $W(0,T;H^{1}(\Omega))$ , which satisfies

Equation (3.19)

and we have the estimate

Equation (3.20)

For the proof of this theorem we refer the readers to section 1 in [17].

Due to the required regularity (3.19) this result is non-standard, at least we could not find a suitable result in the literature. The proof follows by an adaptation of the methods described in [7], for details see [17]. We now turn to prove the existence of a solution to the non-linear system of PDE's describing Problem I.

The functions u0(x) and cat0(x) are supposed to satisfy

Equation (3.21)

Theorem 3.4. For T  >  0 sufficiently small and under condition (3.21), there exists a unique weak solution $(u,r)$ on $ \newcommand{\bM}{\mathbf{M}} \bM_T$ of (3.7)–(3.11).

Sketch of proof. As usual the existence proof for a non-linear PDE is turned into a fixed point problem involving a linear differential equation. We define the operator $\mathcal{T}$

which maps $ \newcommand{\bM}{\mathbf{M}} (\bar r, \bar u) \in \bM_T$ onto $ \newcommand{\calT}{\mathcal{T}} (r,u)=\calT(\bar{r}, \bar{u}) $ with

Equation (3.22)

and u is the solution to the linear parabolic equation

Equation (3.23)

with initial and boundary conditions given by (3.10) and (3.11).

For proving the above stated theorem, we need the following two properties:

  • 1.  
    $ \newcommand{\calT}{\mathcal{T}} \calT$ is defined on $ \newcommand{\bM}{\mathbf{M}} \bM_T$ and maps $ \newcommand{\bM}{\mathbf{M}} \bM_T$ into $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T$ and there exists T small enough such that $ \newcommand{\calT}{\mathcal{T}} \calT$ maps $ \newcommand{\bM}{\mathbf{M}} \bM_T$ onto $ \newcommand{\bM}{\mathbf{M}} \bM_T$ .
  • 2.  
    There exists T small enough such that $ \newcommand{\calT}{\mathcal{T}} \calT$ is s-contractive in $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T\times \bfY_T$ , i.e.
    for all $ \newcommand{\bM}{\mathbf{M}} (u_1,r_1),\,(u_2,r_2) \in \bM_T$ , with $s \in [0,1).$

This approach, as well as the proofs, follow from the classical outline for non-linear PDEs given in [7].

Proof of (1). The tricky part is an appropriate definition of $ \newcommand{\bM}{\mathbf{M}} \bM_T$ and of the function spaces used, the results then follow directly or by applying Hölder's inequality several times. For readers interested in the details of the different steps, we refer to [17]. We directly obtain, that $ \newcommand{\bM}{\mathbf{M}} (\bar{r}, \bar{u})\in \bM_T$ and assuming that cat0 is fixed and satisfies (3.21), implies

  • (i)  
    there exists a constant $C(M,T,R,r_\alpha,{\rm cat}_0)$ depending on $M, T, R, r_\alpha, {\rm cat}_0$ such that $ \left\|r_t\right\|_{L^\infty(Q_T} \leqslant C(M,T,R,r_\alpha,{\rm cat}_0)$ and $\left\|r_x\right\|_{L^\infty(Q_T)}\leqslant C(M,T,R,r_\alpha,{\rm cat}_0)$
  • (ii)  
    There exists a positive constant T1 such that $ r>r_\alpha, \forall (t,x) \in [0,T_1]\times [0,1] := Q_{T_1} $ and the coefficients of the operator $ \newcommand{\calL}{\mathcal{L}} \calL$ in (3.23) belong to $L^\infty(Q_{T_1})$ . Based on theorem 3.3, there exists a unique solution u of (3.23) satisfying $u\in W(0,T_1)$ and by using Hölder's inequality we obtain the following improved regularity result
    and
    Equation (3.24)
    For convenience, we suppose that $T_1\leqslant 1$ , which can be achieved by rescaling the time. From the above results we get the following lemma.

Lemma 3.5 Assume that $ \newcommand{\bM}{\mathbf{M}} (\bar{r}, \bar{u})\in \bM_{T_1}$ . Then, the weak solution u of the problem (3.23), (3.10) and (3.11) satisfies the estimate

Equation (3.25)

It follows from this lemma and the continuous embedding $H^1(\Omega)\subset L^\infty(\Omega)$ that the weak solution u of problem (3.23), (3.10) and (3.11) satisfies the estimates

Equation (3.26)

Equation (3.27)

Equation (3.28)

All constants depend continuously on T and tend to zero for $T \rightarrow 0$ , hence for T sufficiently small we have that $ \newcommand{\bM}{\mathbf{M}} \newcommand{\calT}{\mathcal{T}} \calT: \bM_T \rightarrow \bM_T$ .

Proof of (2). Now we show that $\mathcal{T}$ is contractive in $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T$ norm with T small enough. Let $(\bar{u_1},\bar{r_1})$ , $ \newcommand{\bM}{\mathbf{M}} (\bar{u_2},\bar{r_2})\in \bM_{T^*}$ , where T*  >  0 is such that $ \newcommand{\bM}{\mathbf{M}} \newcommand{\calT}{\mathcal{T}} \calT: \bM_{T^*} \rightarrow \bM_{T^*}$ . is valid. Suppose that

Equation (3.29)

Then $\bar{u}_1,\bar{r}_1,u_1,r_1$ and $\bar{u}_2,\bar{r}_2,u_2,r_2$ satisfy equations (3.7)–(3.11). Denote $\tilde{u}=u_1-u_2, \tilde{r}=r_1-r_2$ . Subtracting the two parabolic equations and the two differential equations, we get the system:

Equation (3.30)

with the initial and boundary conditions

Equation (3.31)

Equation (3.32)

coupled with the equation

Equation (3.33)

Using the same estimates as in the proof of (1.) we obtain the following estimate for the problem (3.30)–(3.32) on $Q_{T^*}=[0,T^*]\times [0,1]$ with T* sufficiently small

Equation (3.34)

and the functions $r_1, r_2$ defined in (3.29) satisfy the estimate:

Equation (3.35)

Combining these two estimates, we see that there exists a positive number T** such that $ \newcommand{\calT}{\mathcal{T}} \calT$ is s-contractive in $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_{T} \times \bfY_{T}$ if $T\leqslant T^{**}$ , where s is a positive constant less than 1.

This ends the sketch of the proof of theorem 3.4.

From now on, we tacitly assume that T is small enough so that there exists a unique solution to the system (3.7)–(3.11).

From lemma 3.5 we also obtain the continuous dependency of the weak solution in $ \newcommand{\bM}{\mathbf{M}} \bM_T$ on the initial data [17].

Theorem 3.6. Let $(r,u)$ be a weak solution of the system (3.7)–(3.11). Then it continuously depends on the data

Equation (3.36)

3.1.2. Continuity and differentiability of the parameter-to-state mapping.

Now we analyse the parameter-to-state map, i.e. we consider $r=r({\rm cat}_0)$ as well as $u=u({\rm cat}_0)$ and show the continuity of $(r,u)$ in $ \newcommand{\bfY}{\mathbf{Y}} W^{1,\infty}(Q_T)\times \bfY_T $ with respect to cat$_0(x) \in W^{1,\infty}(\Omega)$ for Problem I.

Let cat$_{01}(x), {\rm cat}_{02}(x)\in W^{1,\infty}(\Omega)$ , and assume that there exists a positive constant $C_{{\rm cat}}$ such that

Equation (3.37)

Suppose that $ \newcommand{\bM}{\mathbf{M}} (u_1,r_1), (u_2,r_2) \in \bM_T$ , where $T\leqslant T^{**}$ , are solutions to (3.7)–(3.11). Denoting $ \newcommand{\calU}{\mathcal{U}} \newcommand{\calR}{\mathcal{R}} \calU=u_1-u_2,\, \calR=r_1-r_2 $ , we have the following result.

Theorem 3.7. There exists T small enough and a positive constant $C(M,T,R,r_\alpha, u_0, C_{{\rm cat}_0})$ such that

Equation (3.38)

where

Due to the continuous embedding of $W^{1,\infty}(Q_T)$ in $ \newcommand{\bfX}{\mathbf{X}} \bfX_T$ , from theorem 3.7, we have

Corollary 3.8. The pair $(r,c)$ is continuous on $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T $ with respect to ${\rm cat}(x)$ in $W^{1,\infty}(\Omega)$ for problem I with T small enough.

Now we investigate the Fréchet differentiability of r and u with respect to cat0(x) for Problem I. Under the assumptions of theorem 3.4 we define the state mapping $ \newcommand{\calP}{\mathcal{P}} \calP$ :

Equation (3.39)

For $\delta \in W^{1,\infty}(\Omega)$ we denote $\frac{\partial r}{\partial {\rm cat}_0}({\rm cat}_0)\delta$ by $ \newcommand{\e}{{\rm e}} \eta(t,x)$ and $\frac{\partial c}{\partial {\rm cat}_0}({\rm cat}_0)\delta$ by $\varphi(t,x)$ .

By the usual approach, we use the differential equations for $r({\rm cat}_0),u({\rm cat}_0)$ and for $r({\rm cat}_0+\delta),u({\rm cat}_0 +\delta)$ and obtain after expanding all terms and deleting higher order terms the so-called sensitivity system:

Equation (3.40)

Equation (3.41)

Equation (3.42)

Equation (3.43)

Equation (3.44)

We then have to show that a solution to this sensitivity system exists and that its weak solution is indeed the desired Fréchet derivative.

First of all, we see that the sensitivity system (3.40)–(3.44) is linear in $ \newcommand{\e}{{\rm e}} (\eta, \varphi)$ and the right-hand side if of order ${{\mathcal O}}(\delta)$ . Hence standard arguments, e.g. adapting the fix-point techniques of the previous subsection to this system on linear equations, yield the existence of a weak solution of this sensitivity system.

More precisely, for some given positive constant N, we define the set

Equation (3.45)

Lemma 3.9. For T small enough, there exists a unique solution $ \newcommand{\bN}{\mathbf{N}} \newcommand{\e}{{\rm e}} (\eta, \varphi) \in \bN_T$ of the sensitivity system (3.40)–(3.44) and $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \newcommand{\e}{{\rm e}} \|(\eta,\varphi) \|_{L_\infty(\bfX_T \times \bfY_T)}={{\mathcal O}}(\delta) $ .

We now prove, that the parameter-to-state map is indeed Fréchet-differentiable with respect to cat0.

Theorem 3.10. There exists T small enough such that $ \newcommand{\bfX}{\mathbf{X}} r({\rm cat})\in\bfX_T$ and $ \newcommand{\bfY}{\mathbf{Y}} u({\rm cat}_0)\in\bfY_T$ are Fréchet differentiable with respect to cat$_0\in W^{1,\infty}(\Omega)$ in the sense that

To prove the theorem we need the following result.

Lemma 3.11. Denote

Then, $\xi$ and $\psi$ satisfy the problem

Equation (3.46a)

Equation (3.46b)

Equation (3.46c)

Equation (3.46d)

Equation (3.46e)

Equation (3.46f )

where $ \newcommand{\e}{{\rm e}} F_2(\eta,\varphi)$ , $ \newcommand{\e}{{\rm e}} F_4(\eta,\varphi)$ contain only the functions in two variables $ \newcommand{\e}{{\rm e}} \eta$ and $\varphi$ , in which the lowest-degree term is of the second degree and $F_1(\xi,\psi)$ , $F_3(\xi,\psi)$ are linear with respect to $\xi$ and $\psi$ , and

Equation (3.47)

Proof. For proving the lemma we use the Taylor expansion of $ \newcommand{\e}{{\rm e}} H_\epsilon$ and theorem 3.7 to obtain

Equation (3.48)

where $\omega_\delta$ satisfies (3.47).

We need to expand all expression appearing in the differential equations of $\xi$ and $\psi$ . A somewhat longer calculation and ordering terms yields

Equation (3.49)

where

Equation (3.50)

and

Equation (3.51)

We see that $I_1(\xi,\psi)$ is linear with respect to $\xi$ and $\psi$ , and $ \newcommand{\e}{{\rm e}} J_1(\eta,\varphi)$ contains only expressions of second degree in $ \newcommand{\e}{{\rm e}} \eta$ and $\varphi$ .

Similarly, we expand

Equation (3.52)

where $I_2(\xi,\psi) $ is linear in $\xi$ and $\psi$ , $ \newcommand{\e}{{\rm e}} J_2(\eta,\varphi)$ is a second order polynomial in $ \newcommand{\e}{{\rm e}} \eta$ and $\varphi$ , and $\Vert\overline{\omega}_\delta\Vert_{L^\infty(Q_T)}=o(\Vert \delta\Vert_{W^{1,\infty}(\Omega)}).$ We use the differential equations for $r({\rm cat}_0+\delta)$ , $r({\rm cat}_0)$ and $ \newcommand{\e}{{\rm e}} \eta$ and obtain after some algebraic manipulations a differential equation for $\xi_t$ of the form

where $F_3(\xi,\psi) $ is linear in $\xi$ and $\psi$ ,and $ \newcommand{\e}{{\rm e}} F_4(\eta,\varphi)$ is a second order polynomial in $ \newcommand{\e}{{\rm e}} \eta$ and $\varphi$ , and $\Vert\tilde{\omega}_\delta\Vert_{L^\infty(Q_T)}=o(\Vert \delta\Vert_{W^{1,\infty}(\Omega)}).$ Similarly, we use the differential equations for $u({\rm cat}_0+\delta)$ , $u({\rm cat}_0)$ and $\varphi$ and obtain—after expanding and sorting all terms—equation (3.46a). For the details of the expansion see p 54 in [17]. □

Hence, $\xi$ and $\psi$ satisfy linear differential equations and by using theorem 3.7 we observe that the right hand side of the system of equations is of order $o(\Vert \delta\Vert_{W^{1,\infty}(\Omega)})$ . Hence, by standard arguments [7], the weak solution to this problem is unique and the limits in theorem 3.10 hold.

4. Optimization problems

In this section we prepare for the numerical experiments to be performed in the next section. The numerical results will use the original formulation of Problem I and Problem II., i.e. we will use the version without smoothing the coefficients $\beta$ and $\gamma$ . Hence, there is a gap in theory as we suppose that r and c, solutions to Problem I (or Problem II), exist and that equivalent results concerning derivatives and sensitivity systems hold.

4.1. The first optimization problem

The solution r (the pore radius) depends on the catalyst position cat0(x), time t and space variable x: $r := r({\rm cat}_0;t,x)$ . Our aim is to minimize the free volume (2.11) of the pore after some time T, with respect to cat0 over some admissible set $ \newcommand{\bA}{\boldsymbol{A}} \bA$ .

As $ \newcommand{\bfX}{\mathbf{X}} r\in \bfX_T \subset C([0,T]; L^2(\Omega))$ , we can define the linear bounded projection operator

Equation (4.1)

Hence

Equation (4.2)

In section 3 we have shown that $r({\rm cat}_0;t,x)$ is Fréchet differentiable with respect to cat$_0\in W^{1,\infty}(\Omega)$ for the smoothed version of Problem I, similarly we assume that there exists a linear bounded operator

Equation (4.3)

such that, for $\delta \in W^{1,\infty}(\Omega)$ ,

Equation (4.4)

which implies that

and from that we directly obtain that the functional $V({\rm cat}_0)$ is Fréchet differentiable with respect to cat$_0\in W^{1,\infty}(\Omega)$ .

Remark 4.1. For convenience, we will use the notation $r_{{\rm cat}_0}(\cdot)$ for $r_{{\rm cat}_0}({\rm cat}_0)$ .

Although $r({\rm cat}_0)$ and $c({\rm cat}_0)$ are Fréchet differentiable with respect to cat0 in $L^\infty$ -norm, however, for computing their adjoint operators we have to suppose that they are Fréchet differentiable with respect to cat0 in L2-norm, that means $r_{{\rm cat}_0}(\cdot)$ and $c_{{\rm cat}_0}(\cdot)$ are linear bounded operators

Equation (4.5)

Equation (4.6)

Further, we suppose that the projection operator $\mathbf{P}_{T}$ is bounded from $L^2(Q_T)$ onto $L^2(\Omega)$

Equation (4.7)

From these assumptions, by the Riesz representation theorem, we see that $V_{{\rm cat}_0}(\cdot)$ is identified with

Equation (4.8)

A similar result can be obtained for Problem II. In the next paragraph we will present a scheme how to calculate $(r_{{\rm cat}_0}(\cdot)){}^*$ .

To prove the existence of a minimizer of the functional (4.2) we introduce the admissible set $ \newcommand{\bA}{\boldsymbol{A}} \bA$ (see [9, 20]) as follows:

Equation (4.9)

where $M_1, M_2$ are some given positive numbers and $0<\lambda<1$ . The set $ \newcommand{\bA}{\boldsymbol{A}} \bA$ is compact in $C^1(\Omega)$ [20]. Since $V({\rm cat}_0)$ is continuous in $W^{1,\infty}(\Omega)$ , we have the following result.

Theorem 4.2. The problem of minimizing $V({\rm cat}_0)$ subject to Problem I over $ \newcommand{\bA}{\boldsymbol{A}} \bA$ admits at least one solution.

4.2. The second optimization problem

We consider the second optimization problem

Equation (4.10)

subject to Problem I, and

Equation (4.11)

subject to Problem II.

Due to the continuous embedding $ \newcommand{\bfX}{\mathbf{X}} L^2(Q_T) \subset \bfX_T$ , the projection operator:

Equation (4.12)

is linear and bounded. Denoting by $<\cdot,\cdot>$ the inner product in the Hilbert space $L^2(Q_T)$ and $<\cdot,\cdot>_{L^2([T_1,T]\times \Omega)}$ the inner product in the Hilbert space $L^2([T_1,T]\times \Omega)$ , we see that the functional $\mathcal{V}$ in (4.10) is Fréchet differentiable with respect to cat$_0\in W^{1,\infty}(\Omega)$ . Indeed, taking $\delta\in W^{1,\infty}(\Omega) $ , we have

Equation (4.13)

As we will see, the derivative of ${\mathcal V}$ with respect to cat0 will involve the adjoint of the derivative $r_{{\rm cat}_0}(\cdot)$ . In order to compute this adjoint, we consider $r_{{\rm cat}_0}(\cdot)$ and $c_{{\rm cat}_0}(\cdot)$ as mappings between L2-spaces. This is a common approach, this technique has been used in several papers, e.g. [8]. In doing so, we suppose that there exist the linear and bounded operators:

Equation (4.14)

Equation (4.15)

Since these operators are linear and bounded, their adjoint operators $(r_{{\rm cat}_0}(\cdot)){}^*: L^2(Q_T)\longrightarrow L^2(\Omega)$ , and $(c_{{\rm cat}_0}(\cdot)){}^*: L^2(Q_T)\longrightarrow L^2(\Omega)$ are also so. From this, we immediately get

Equation (4.16)

Hence, $\mathcal{V}$ is Fréchet differentiable with respect to cat$_0\in {W^{1,\infty}(\Omega)}$ .

We also see that the problem of minimizing $\mathcal{V}({\rm cat}_0)$ with respect to cat$ \newcommand{\bA}{\boldsymbol{A}} _0(x) \in \bA$ defined by (4.9) admits a solution.

Denote the function z by

Equation (4.17)

By the Riesz representation theorem we see that the Fréchet derivative $\mathcal{V}_{{\rm cat}_0}(\cdot)$ can be identified with $ 2 (r_{{\rm cat}_0}(\cdot)){}^{*} z({\rm cat}_0)$ . Hence

Now we present the scheme for calculating the adjoint operator $(r_{{\rm cat}_0}(\cdot)){}^*$ for Problem I. With the assumptions of smoothness on $r,c$ and positivity of r we can transform Problem I: (2.1a)–(2.1e) into the problem:

Equation (4.18a)

Equation (4.18b)

Equation (4.18c)

Equation (4.18d)

Equation (4.18e)

Equation (4.18f )

where $\gamma$ and $\beta$ are defined in (3.1). Now, as in the previous section, supposing that there exists a solution of the problem for the original model in the space $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T$ , and $ \newcommand{\e}{{\rm e}} \eta$ and $\varphi$ are defined by $ \newcommand{\e}{{\rm e}} {\eta}:=r_{{\rm cat}_0}(\cdot)\delta,\, {\varphi}:=c_{{\rm cat}_0}(\cdot)\delta$ , for $\delta\in W^{1,\infty}(\Omega)$ , we introduce the sensitivity system:

Equation (4.19)

with the initial condition

Equation (4.20)

and

Equation (4.21)

with initial and boundary conditions

Equation (4.22)

We will apply the adjoint technique by introducing an adjoint system corresponding to the above sensitivity system in order to calculate $(r_{{\rm cat}_0}(\cdot)){}^{*} z({\rm cat})$ .

The structure of the sensitivity system is given by

Equation (4.23)

Using by integration by parts and summing up the two equations, we obtain

Equation (4.24)

This then yields the following theorem.

Adjoint system for Problem I. Suppose that the system

Equation (4.25)

Equation (4.26)

Equation (4.27)

Equation (4.28)

has a unique solution $(\nu,w)\in W^{1,\infty}(Q_T)\times W(0,T)$ for some positive T. Suppose further that

Equation (4.29)

Then, the operator $(r_{{\rm cat}_0}(\cdot)){}^{*}: L^2(Q_T) \mapsto L^2(\Omega)$ , $z \mapsto (r_{{\rm cat}_0}(\cdot)){}^{*} z$ has the form

Now present the scheme of calculating the adjoint operator $(r_{{\rm cat}_0}(\cdot)){}^*$ for Problem II.

Problem II leads to system

Equation (4.30a)

Equation (4.30b)

Equation (4.30c)

Equation (4.30d)

Equation (4.30e)

Equation (4.30f )

where ${\rm cat}(t,x)$ is defined by (2.8) and the initial condition (2.9) for given $\alpha_0$ and R0.

Assume that $ \newcommand{\bfX}{\mathbf{X}} r\in\bfX_T$ and $ \newcommand{\bfY}{\mathbf{Y}} c\in \bfY_T$ are Fréchet differentiable with respect to cat$_0\in W^{1,\infty}(\Omega)$ for T small enough. Denote $ \newcommand{\e}{{\rm e}} \bar{\eta}:=r_{{\rm cat}_0}(\cdot)\Delta,\, \bar{\varphi}:=c_{{\rm cat}_0}(\cdot)\Delta,\, {\psi}:={\rm cat}_{{\rm cat}_0}(\cdot)\Delta$ , for $\delta\in W^{1,\infty}(\Omega)$ . Similarly to the previous one, suppose that there exists a solution of Problem II in the space $ \newcommand{\bfX}{\mathbf{X}} \newcommand{\bfY}{\mathbf{Y}} \bfX_T \times \bfY_T$ . Then, we get the sensitivity system:

Equation (4.31)

with the initial condition

Equation (4.32)

and

Equation (4.33)

with initial and boundary conditions

Equation (4.34)

and

Equation (4.35)

with initial condition

Equation (4.36)

Denote the function $\mathbf{z}$ by

Equation (4.37)

Under assumptions about operators $(r_{{\rm cat}_0}(\cdot)){}^*$ and $(c_{{\rm cat}_0}(\cdot)){}^*$ in (4.14) and (4.15), the functional $\tilde{\mathcal{V}}$ defined in (4.11) is Fréchet differentiable with respect to cat$_0\in L^2(\Omega)$ . Moreover, according to Riesz representation theorem the Fréchet derivative $\tilde{\mathcal{V}}_{{\rm cat}_0}(\cdot)$ is identified by

$2(r_{{\rm cat}_0}(\cdot)){}^*(\mathbf{P}_{T_1,T}){}^* \mathbf{P}_{T_1,T}r({\rm cat}_0)$ . From which we see that $\tilde{\mathcal{V}}_{{\rm cat}_0}(\cdot)$ can also be identified by

$2(r_{{\rm cat}_0}(\cdot)){}^{*}\mathbf{z}({\rm cat})$ , where $\mathbf{z}({\rm cat}_0)$ is defined in (4.37). Hence

By integration by parts, we can calculate $(r_{{\rm cat}_0}(\cdot)){}^{*} \mathbf{z}({\rm cat}_0)$ by the adjoint system as follows:

Adjoint system for problem II: Suppose that the system

Equation (4.38)

Equation (4.39)

with the conditions

Equation (4.40)

Equation (4.41)

and

Equation (4.42)

with the condition

Equation (4.43)

has a unique solution $(\nu,w,{\rm cat})\in W^{1,\infty}(Q_T)\times W(0,T)\times W^{1,\infty}(Q_T)$ . Additionally, suppose that

Equation (4.44)

Then, the operator $(r_{{\rm cat}_0}(\cdot)){}^*: L^2(Q_T) \mapsto L^2(\Omega)$ ; $z \mapsto -\zeta(0,\cdot)$ has the form

Equation (4.45)

We note that we have applied the adjoint and sensibility systems to gradient methods but these gave good numerical results only for short time processes [17], therefore in the next section we use directional gradients for iterative methods.

5. Numerical simulation

We now address the problem of optimizing the initial catalysts cat0 numerically. However, we would like to stress that the main purpose of this paper is the analytical investigations of the catalyst models as presented above. The following numerical examples are highlighting some of the analytical features proved above and they serve illustration purposes. In addition, the numerical results presented are somewhat incomplete. The implementation presented below is based on an explicit discretization scheme resulting in a very large number of very small time steps. However, applying the analytical optimization procedure using the adjoint system, requires to store the forward solution for every time step. The complexity of the computations is beyond the capacity of the computers used in this simulation. Hence updating the numerics scheme is left for future work and poses a challenge by itself. It will require a complete reimplementation of the code and—most likely- novel concepts for storing compressed version of the forward solution together with a decompression and interpolation scheme.

We admit that the following numerical experiments therefore constitute only a first step towards an implementation useful for real life applications. We could either use Problem I or Problem II and also one of the two different optimization criteria listed in section 4, which gives 4 different optimization scenarios. Note that we refer again to the models for $(r,c)$ introduced in section 2. For numerical comparison we will treat three different algorithms (standard greedy, incremental greedy, adjoint method), which will be explained in this section. A detailed analysis of all these cases is beyond the scope of this paper. Hence, we only list results for what we believe to be the most important case for applications, namely Problem II with the first optimization criterion. Accordingly, in this paper we only present the results for Problem II, for a more extensive description of all numerical tests including tests of finite difference schemes versus finite element schemes as well as experiments for Problem I see [17].

As already mentioned, we use three different numerical approaches. The first one is a greedy approach following the experiments of [4], where cat0 can take only values 0 or 1. This illustrates the case of discrete catalysts positioning and the reaction efficiency of the catalysts is equal to 1. I.e. once a position x has been assigned the value cat0(x)  =  1 it will remain with this value. We apply this method for comparison with the other two numerical schemes.

The second scheme is a greedy algorithm with incremental updates, i.e. we fix a value $\Delta {\rm cat}$ , e.g. $\Delta {\rm cat} =0.1$ , and allow to update every position several times until an optimum is obtained. Further details are given in subsection 'Incremental greedy'.

As a third scheme we use the analytic results of the previous sections and update cat0 iteratively by a gradient descent algorithms

Equation (5.1)

Throughout this section we use an equidistant discretization for the spatial variable xj  =  j *h,j   =  0,..,(J  −  1) with $h=1/(J-1)$ , a typical choice for simulation is J  =  100. An equidistant time discretisation $t_n= n \Delta t $ with $\Delta t = 1/(N-1)$ is chosen, where the number of time steps depends on the clogging of the pore. The time discretisation uses $n=1,..,\tilde N$ with $\tilde N$ typically being much larger than N.

Hence cat$_j^n={\rm cat}(t_n,x_j)$ , similarly, we discretise r and c. The update for r is done by an explicit time stepping

Equation (5.2)

For the discretization of the differential equation for c we use a standard finite difference scheme. The integral equation for updating ${\rm cat}(t,x)$ is discretized by a Riemann sum using the same discretization points xj.

5.1. Greedy algorithm

As a starting point we implement the optimization scheme of [4] for comparison with our more involved schemes in the following subsections. That is, we consider the case that the catalytic function cat0 takes values 0 or 1 at positions along the surface of the pore. In discretized form we have cat$_j^0, j=0,\cdots,(J-1)$ , where cat$_j^0={\rm cat}_0(x_j)$ . We emphasize, that the update for catt is done a simpler way in this approach: whenever $r^n_j$ reaches a value smaller than R0, then both neighboring catalyst positions are set to 1, i.e. ${\rm cat}^{n+1}_{j-1}={\rm cat}^{n+1}_{j+1}=1$ .

Before starting the algorithm all values are set to zero. In the first round of the greedy algorithm all potential catalyst positions are tested separately, i.e. for $ \newcommand{\e}{{\rm e}} \ell = 0,.., (J-1)$ we set cat$ \newcommand{\e}{{\rm e}} _j^0=\partial_{j,\ell}$ and run the algorithm until clogging. We determine the index $ \newcommand{\e}{{\rm e}} \ell$ , which returns the minimal free volume. At this position cat$ \newcommand{\e}{{\rm e}} _\ell $ is fixed as one and the greedy algorithms proceeds with the remaining indices. Hence, after every full round of the greedy algorithm the number of non-zero cat$_j^0$ values increases by one. The optimal value for the remaining free volume was obtained with 52 non-zero catalyst positions, see figure 3 right. For the experiments presented in this section, we actually stopped the process prior to clogging, more precisely we ran the algorithm until $\min_j r_j^n \leqslant r_0$ with r0  =  0.1.

Figures 2 and 3 show the evolution of pore radii and concentration of oxygen along the pore in 3D plotting and in the line plotting over time with the optimal catalyst number of 52, which is illustrated in the figure 4. The optimal volume we deduce is $22.8\%$ . In this experiment, we take $J=100, N=4000$ , threshold of R0  =  0.96. Figure 4 demonstrates the dependency of free volume $V_{{{\rm free}}}$ before pore clogging on the number of catalyst positioned along the pore.

Figure 2.

Figure 2. Results obtained by classical greed algorithm. Left: Li2O growth profile or the radii of pore over time. Right: concentration of oxygen inside the pore over time.

Standard image High-resolution image
Figure 3.

Figure 3. Results obtained by classical greedy algorithm. Left: Li2O growth profile or the radii of pore over time, computations are shown with 52 non-zero cat$_j^0$ positions, i.e. cat$_j^0=1$ at 52 values of j . Right: concentration of oxygen inside the pore over time.

Standard image High-resolution image
Figure 4.

Figure 4. Results obtained by classical greedy algorithm. Left: distribution of optimal number of catalysts of 52 catalyst positions in Problem II. Right: minimum free volumes corresponding to number of catalyst along the pore via discrete greedy method.

Standard image High-resolution image

Figure 4 shows on the left that the greedy algorithm places most catalysts at the end of the pore. This is to be expected, since placing catalysts at the beginning of the pore, where the oxygen concentration is still high will lead to a rapid clogging of the pore at a position near the beginning of the pore and a large free volume in the later part. This is confirmed by figure 3 on the right, which indicates, that as time evolves more and more oxygen is consumed at the beginning of the pore. Hence rather little oxygen is transported towards the later parts of the pore. Hence, no catalysts are placed at the beginning of the pore before position 0.4 but after the growing process clogging occurs at position 0.34, see figure 3 on the left. Figure 4 on the right shows the evolution of the free volume as more and more catalysts are added by the greedy algorithm. Some interesting local minima occur, i.e. stopping the classical greedy algorithm in the local minimum after 24 iterations yields a suboptimal result.

Also, figure 3 on the left shows some drawbacks of the classical greedy algorithm, which put the initial catalysts activity either 0 or 1. This does not allow for a smooth development in the catalyst reaction process and results in the spike just before position 0.6

A single run of the forward pass requires approx. 12 s. The running time for testing all profiles is 142h with Intel Xeon Processor E5-2687W v4 with 12 cores and 1.54 TB of memory. We apply OpenMP application program interface to carry out parallel computations for optimizing the running time.

5.2. Incremental greedy algorithm for problem II

The results of the previous subsection are obtained with the version of the greedy algorithm for catalyst optimisation which was used in [4], hence the optimized value of approx. $20 \%$ free volume at the time of clogging is the reference value for the more involved algorithms of this and the next subsection.

In this section we simply improve the greedy algorithm by choosing smaller increments, i.e. 0.1 instead of 1, and by allowing to update to the same position of catalyst for several times. Hence, all J positions have to be tested in every iteration of this incremental greedy algorithm. The resulting cat0-function has a more geometric structure which also reflects engineering intuition about optimal catalyst positioning.

There is no limit to the maximal value of cat$_j^0$ , hence, we cannot test all possible cat0-profiles. Instead we use the first local minimum of $V_{\rm free}$ as stopping criterion. This minimum occurred after the 596th iteration, the optimal $V_{{{\rm free}}}$ is $7.8\%$ . The parameters we used for this experiment are: J  =  100 and $\Delta t=\frac1{100-1}$ and $\alpha_0=0.01$ . Figure 5 shows radii and concentration inside the pore before clogging corresponding to the optimal catalytic function cat0 in 3D plotting by incremental Greedy in Problem II, while the in line plots of radii and concentration inside the pore are shown in figure 6. The optimal catalytic function cat0 is plotted in figure 7 (left) and the dependence of free volume before pore clogging $V_{{{\rm free}}}$ with respect to numbers of increments is demonstrated in figure 7 (right).

Figure 5.

Figure 5. Results obtained by incremental greed algorithm. Left: Li2O growth profile or the radii of pore over time. Right: concentration of oxygen inside the pore over time.

Standard image High-resolution image
Figure 6.

Figure 6. Results obtained by incremental greed algorithm. Left: Li2O growth profile or the radii of pore over time. Right: concentration of oxygen inside the pore over time.

Standard image High-resolution image
Figure 7.

Figure 7. Left: distribution of optimal number of catalysts obtained by incremental greedy algorithm with steps of 0.1 in Problem II. Right: minimum free volumes corresponding to number of increments 0.1 along the pore via incremental greedy method in Problem II.

Standard image High-resolution image

Compared with the results obtained by the classical greedy algorithm in the previous subsection figure 5 shows a smoother transition of the catalyst growth and oxygen concentration along the pore. Most interesting is the final catalyst distribution after the incremental greedy stopped after 600 iterations, see figure 7. It matches the qualitative expectations, that catalysts should be placed predominantly towards the end of the pore in order to minimize the free volume after clogging. However, these simulation add a quantitative estimate to this behavior and yields an almost optimal distribution in terms of the free volume of $\% 7.8$ after clogging, see figure 7 on the right. This can only minimally improved by further analytic iterations see next subsection. Again, the convergence towards a minimum is rather discontinuous as the number of incremental greedy iterations increases, see figure 7 on the right. Figures 5 and 6 both display the same data, namely the development of the pore radius and the of the oxygen concentration over time. Figure 5 allows a more general visualization of this process. However, figure 6 shows a rather dynamic local behavior, which might be overcome by an even finer incremental step size. However, using OpenMP application program interface for implementing parallel computations on an Intel Xeon Processor E5  −  2687W $v4$ still results in a running time of 163h. Hence, a refined discretisations with smaller increments is beyond any feasible computation time for this comparative study.

5.3. Gradient descent method

For Problem II, we consider the first optimization Problem: find cat0 such that the free volume $V({\rm cat}_0)$ of the pore after some time T defined by (2.11) reaches to a minimum.

The optimization problem is solved by the gradient descent method

Equation (5.3)

The iteration stops when the free volume $V({\rm cat}_0^{\rm new})$ is greater than the previous one $V({\rm cat}_0^{\rm old})$ . In the context of inverse problems such parameter identification problems for partial differential equations have been studied intensively, see e.g. [2] for an application in scattering or [10] for a recent survey.

The update of cat0 uses the derivative of $V$ with respect to cat0. According to section 4,

This can be computed by solving the adjoint system for $r^*_{{\rm cat}_0}$ with $z= \mathbf{P}_T^* \mathbf{P}_T r({\rm cat}_0)$ , see (4.8). However, this is a parabolic equation backward in time and requires to store the solution of the forward problem at every time step. This is beyond the capacity of the computer used in these experiment.

Hence, suppose that $\mathbf{P}^{(k)}_T r^{(k)}_{{\rm cat}_0}$ is approximated for $\mathbf{P}_T r_{{\rm cat}_0}$ at the iteration k of the optimization algorithm, then $\mathbf{P}^{(k)}_T r^{(k)}_{{\rm cat}_0}$ is a $J\times J$ matrix, therefore the adjoint operator is defined by its transpose matrix $(\mathbf{P}^{(k)}_T r^{(k)}_{{\rm cat}_0}){}^*=(\mathbf{P}^{(k)}_T r^{(k)}_{{\rm cat}_0}){}^{\rm T}$ . In order to determine the derivative $r^{(k)}_{{\rm cat}_0}$ of pore radius r with respect to cat0, we calculate the directional derivatives by

Equation (5.4)

where h is small enough positive number and $e_j=(0,\cdot, 1, \cdots, 0)$ is a $(1\times J)$ vector, taking value of 1 at the j th coordinate.

For the experiment of the gradient descent, we take the initial catalytic function from the 596th step of the incremental greedy algorithm. The resulting catalytic function is plotted in figure 8 (left). The parameters for gradient descent method are

and the regularization parameter is chosen from a decreasing sequence

Figure 8.

Figure 8. Left: initial catalytic function used for gradient descent method in Problem II. Right: difference of optimal catalytic function compared to initial catalytic function, which was obtained by the incremental greedy algorithm for Problem II.

Standard image High-resolution image

The numerical iteration for the incremental greedy was stopped as soon as $r(t,x) < 0.1$ at some position x for the first time. In the presented experiment this point was reached after $T=1571\,393$ time steps.

In the following experiments using the gradient descent method, we kept $T=1571\,393$ for simplicity. After 48 iterations of the gradient descent, we get the local minima which refers us to a free volume $V{{{\rm free}}}=7,74\%$ . Hence, we get a minor improvement for the free volume $V_{{{\rm free}}}$ : $0.115\%$ . Figure 8 (right) shows the changing amount(increasing over time) of catalytic function by gradient descent method after 48 iterations. The running time is 14 h with Intel processor core i5-5200U and memory size is 12 GB.

6. Conclusions

We have introduced and analysed two systems of PDEs describing the clogging process of non-aqueous lithium/air batteries. The main parameter in the optimisation considered in this paper is the initial catalyst distribution cat0 along the surface of the pores of the battery. We have assume a simple symmetric pore model and the clogging is described by the diminishing pore radius $r(t,x)$ . For a static catalyst model, i.e. ${\rm cat}(t,x)={\rm cat}_0(x)$ we introduced a smoothed version for which we proved existence and uniqueness as well as Fréchet-differentiability of the parameter-to-state map which links cat0 to r. Furthermore, we have considered two optimisation scenarios. Here we derived the derivatives of the optimization functionals with respect to cat0. Finally, we presented numerical results for an extended model including a dynamic behaviour ${\rm cat}(t,x)$ of the catalyst distribution. These simulations result in a proposal for initial catalyst doting which reflects engineering intuition.

Future work will focus on extending the analytic results, in particular Fréchet differentiability of the parameter-to-state map, to the dynamic case. The final aim to also include shape optimization of the pores, i.e. we regard the present model of a cylindrical pore as a first step towards a more complex geometry.

Acknowledgments

The research by Dinh Nho Hào was supported by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under Grant No. 101.02-2017.318. The work of Peter Maass was supported by the Deutsche Forschungsgemeinschaft (DFG) within the research training network GRK 2224/1 $\pi^3$ : Parameter Identification—Analysis, Algorithms, Applications. The research by Thuong-Huyen Nguyen was supported by Project 911 VIED fellowships and Center for Industrial Mathematics, University of Bremen (ZeTeM). The authors would like to thank Dr Tatjana Dabrowski for introducing them to this interesting problem.

Please wait… references are loading.
10.1088/1361-6420/ab47e6