Nothing Special   »   [go: up one dir, main page]

Dielectric Breakdown by Electric-Field Induced Phase Separation

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Journal of the Electrochemical

Society

ACCEPTED MANUSCRIPT

Dielectric breakdown by electric-field induced phase separation


To cite this article before publication: Dimitrios Fraggedakis et al 2020 J. Electrochem. Soc. in press https://doi.org/10.1149/1945-7111/aba552

Manuscript version: Accepted Manuscript


Accepted Manuscript is “the version of the article accepted for publication including all changes made as a result of the peer review process,
and which may also include the addition to the article by IOP Publishing of a header, an article ID, a cover sheet and/or an ‘Accepted
Manuscript’ watermark, but excluding any other editing, typesetting or other changes made by IOP Publishing and/or its licensors”

This Accepted Manuscript is © 2020 The Author(s). Published by IOP Publishing Ltd..

This article can be copied and redistributed on non commercial subject and institutional repositories.

Although reasonable endeavours have been taken to obtain all necessary permissions from third parties to include their copyrighted content
within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this
article, please refer to the Version of Record on IOPscience once published for full citation and copyright details, as permissions will likely be
required. All third party content is fully copyright protected, unless specifically stated otherwise in the figure caption in the Version of Record.

View the article online for updates and enhancements.

This content was downloaded from IP address 134.117.10.200 on 17/07/2020 at 02:22


Journal of The Electrochemical Society

pt
Dielectric breakdown by electric-field induced phase
separation

cri
Journal: Journal of The Electrochemical Society

Manuscript ID JES-101542.R1

Manuscript Type: Research Paper

Date Submitted by the


07-Jul-2020
Author:

us
Complete List of Authors: Fraggedakis, Dimitrios; Massachusetts Institute of Technology,
Mirzadeh, Mohammad; Massachusetts Institute of Technology
Zhou, Tingtao; Massachusetts Institute of Technology
Fo
Bazant, Martin; Massachusetts Institute of Technology

Intercalation, Phase Separation, Memristors, Energy Storage,


Keywords:

an
Information Storage, Dielectrics, Metal Insulator Transition
rR
dM ev
iew
On
ly
pte
ce
Ac

https://mc04.manuscriptcentral.com/jes-ecs
Page 1 of 27 Journal of The Electrochemical Society

1
2
3 Dielectric breakdown by electric-field induced phase separation
4

pt
5
6 Dimitrios Fraggedakis1 ,∗ Mohammad Mirzadeh1 , Tingtao Zhou2 , and Martin Z. Bazant1,3†
7
1
8 Department of Chemical Engineering,
9

cri
10 Massachusetts Institute of Technology, Cambridge, MA 02139 USA
11 2
12
Department of Physics, Massachusetts Institute
13 of Technology, Cambridge, MA 02139 USA and
14
15 3
Department of Mathematics, Massachusetts Institute
16

us
17 of Technology, Cambridge, MA 02139 USA
18
19 (Dated: July 7, 2020)
Fo
20
21 Abstract
22

an
The control of the dielectric and conductive properties of device-level systems is important for
rR
23
24
25 increasing the efficiency of energy- and information-related technologies. In some cases, such as
26
ev

neuromorphic computing, it is desirable to increase the conductivity of an initially insulating


27
28 medium by several orders of magnitude, resulting in effective dielectric breakdown. Here, we
dM
29
iew

30 show that by tuning the value of the applied electric field in systems with variable permittivity
31
32 and electric conductivity, e.g. ion intercalation materials, we can vary the device-level electrical
33
34 conductivity by orders of magnitude. We attribute this behavior to the formation of filament-
On

35
like conductive domains that percolate throughout the system, which form only when the electric
36
37 conductivity depends on the concentration. We conclude by discussing the applicability of our
38
ly
pte

39 results in neuromorphic computing devices and Li-ion batteries.


40
41
42
43
44
45
46
ce

47
48
49
50
51
52
Ac

53
54
55

56 email: dimfraged@gmail.com
57 †
email: bazant@mit.edu
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
1
Journal of The Electrochemical Society Page 2 of 27

1
2
3 I. INTRODUCTION
4

pt
5
6
7
Phase separating materials play a key role in several applications related to energy har-
8 vesting and storage, as well as information storage and processing. Some characteristic
9

cri
10 examples are Li-ion batteries [1], phase change [2] and redox [3] memristive devices [4], alloy
11
12 catalysts [5] and self-organized surface nanoreactors [6]. In some cases, phase separation is
13
desirable as one can increase the efficiency of the system (e.g. increase catalytic activity [7],
14
15 change of electric [8, 9] and/or thermal [10] conductivity). However, there are other cases
16

us
17 where phase separation degrades the performance of a device resulting in decreased life-
18
19 time (e.g. fracture of secondary electrode particles in Li-ion batteries [11, 12] that decreases
Fo
20
the available active material, delamination at electrode-electrolyte interface in all-solid-state
21
22

an
batteries [13] resulting in loss of contact of active material with the electrode). Hence, it
rR
23
24 is crucial to find ways that can actively control the occurrence and/or suppression of phase
25
26 separation, which can help us increase the efficiency of existing technologies, as well as open
ev

27
new possibilities on exploiting physical phenomena for new applications.
28
dM
29
iew

Phase separation can be described as a form of instability. For example, a homogeneous


30
31 binary mixture is thermodynamically unstable when its average concentration lies inside the
32
33 spinodal region [14]. In this situation, any infinitesimal perturbation on the concentration
34
On

35 field would evolve in time, making the system to form domains of the two phases [15]. In
36
general, there have been several efforts to control or induce the formation of instabilities in
37
38 equilibrium and non-equilibrium systems. Some examples are: 1) control of viscous finger-
ly
pte

39
40 ing through the application of electric fields [16, 17], 2) stabilization of thermodynamically
41
42 unstable mixtures used in Li-ion batteries, such as LiFePO4 , using non-equilibrium driving
43
forces, galvanostatic conditions [1, 18–20] 3) destabilization of homogeneous polymeric, col-
44
45 loidal, electrolyte and glass mixtures under the application of electric field [21–28]. In the
46
ce

47 present work, we are interested in understanding the control of phase separation of mixtures
48
49 through electric fields and its impact on the transport properties of the system, e.g. change
50 of electric conductivity after phase separation occurs.
51
52
In many dielectric mixtures that phase separate under electric fields, e.g. colloids [29–32],
Ac

53
54 polymers [33], amorphous solids [23], electrolytes [27], the electric permittivity depends on
55
56 the corresponding species concentration [34]. The effect of this dependence can be under-
57
58 stood in the simple case of a binary mixture, where the applied electric field contributes
59
60 https://mc04.manuscriptcentral.com/jes-ecs
2
Page 3 of 27 Journal of The Electrochemical Society

1
2
3 to the total chemical potential µ, i.e. µE ∼ ∂c ε |E|2 [35]. Therefore, a combination of a
4

pt
5 nonlinear concentration-dependent permittivity combined with high electric fields can alter
6
7 the free energy to change the miscibility gap and spinodal region.
8
9 The effect of the electric field has mainly to do with the thermodynamics stability, how-

cri
10
ever, the formed phase morphologies, e.g. filament-like structures [36, 37], can greatly affect
11
12 the transport properties on the macroscopic level, e.g. from low to high electric conductance
13
14 and vice versa, [9], leading to phenomena that resemble dielectric breakdown [38, 39]. Pre-
15
16 vious studies have focused identifying the conditions for dielectric breakdown due to phase

us
17
separation solely based on thermodynamics [9, 23, 29–32], and few studies have discussed the
18
19 formation and dynamics of the conductive filaments [36], which is crucial for technological
Fo
20
21 applications [2]. There are several solid-state materials that form conductive and insulating
22

an
domains when they phase separate. For example, most of the commercial Li-ion interca-
rR
23
24
lation materials, i.e. Lix CoO2 [8, 40], Lix Ni1/3 Co1/3 Mn1/3 O2 [41], Li4+3x Ti5 O12 [42, 43],
25
26 share this property as they undergo a metal-to-insulator transition (MIT), along with an
ev

27
28 ion concentration-dependent permittivity. This combination of characteristics makes them
dM
29
iew

30 perfect candidates for studying the effect of electric fields on both the phase separation and
31 the electric conduction, which can lead to dielectric breakdown.
32
33 The goal of the present work is to develop a simple phenomenological theory that de-
34
On

35 scribes dielectric breakdown due to electric-field induced phase separation. Based on a


36
37 concentration-dependent electric permittivity, we show that a homogeneous stable solution
38
ly

can phase separate in two (or more) phases after a critical electric field is applied. Addi-
pte

39
40 tionally, we consider the case where one of the phases is a metal and the other an insulator,
41
42 which translates in concentration-dependent electric conductivity. When the initial con-
43
44 centration is such that the material is insulating, after phase separation occurs we find the
45
system to conduct current like a metal. The transition from insulating to metallic behav-
46
ce

47 ior after electric field is applied corresponds to an effective dielectric breakdown, which is
48
49 attributed to the formation of filament-like structures that span the entire domain. This
50
51 phenomenon is related to the pioneering works by Goldhammer [44] & Herzfeld [45], that
52
link the changes in the electric permittivity with the species concentration to the metal-to-
Ac

53
54 insulator transition. We relate our results to Li-ion intercalation materials, such as Lix CoO2
55
56 and Li4+3x Ti5 O12 , and we discuss the implications of our theory for resistive switching and
57
58 Li-ion battery applications.
59
60 https://mc04.manuscriptcentral.com/jes-ecs
3
Journal of The Electrochemical Society Page 4 of 27

1
2
3 II. THEORY
4

pt
5
6 We assume a phase separating dielectric medium placed between two blocking electrodes,
7
8 as shown in Fig. 1. Purple and yellow show the two phases, respectively, that can be formed
9

cri
10 either because the mixture is thermodynamically unstable or due to the application of an
11
12 electric field E that demixes the solid solution mixture. In the present picture, the two
13 phases have different electric permittivities, ε1 and ε2 , and electrical conductivities, σ1 and
14
15 σ2 . For simplicity we assume ε2 > ε1 and σ2 > σ1 , respectively. Representative examples of
16

us
17 such system are ion intercalation materials [46–49], which are thermodynamically unstable
18
19 for a wide range of Li-ion fraction. Additionally, some of them undergo metal-to-insulator
Fo
20 transition, e.g. Lix CoO2 [8, 40] (LCO), Li4+3x Ti5 O12 [42, 43] (LTO), and Lix TiO2 [50, 51],
21
22 where one of the phases has much larger electrical conductivity compared to the other - the

an
rR
23
24 difference can be up to six orders of magnitude [38]. In the following sections, we continue
25
26
with the thermodynamics and transport theory that describes the electric-field induced phase
ev

27 separation and consequently the dielectric breakdown of the medium along the direction of
28
dM
29 the electric field.
iew

30
31
32
33
34 A. Thermodynamics
On

35
36
37 We are interested in modeling the dielectric phase separating medium shown in Fig. 1.
38
ly

The neutral species is described by the local fractional concentration c = n/nmax , where
pte

39
40
41
nmax is the maximum species concentration in the medium. Under constant temperature
42 and pressure, the Gibbs free energy of the system is [9, 15, 35, 52]
43
44 Z  
45 1 1
46 G [c, φ, ρ] = dx gh (c) + κ |∇c|2 − ε(c) |∇φ|2 + ρφ (1)
V 2 2
ce

47
48
49 The first term is the homogeneous free energy of the system gh and the second term cor-
50
51 responds to the penalty gradient term [15, 53–55], which is used to describe the phase
52
separation of the material. The phenomenological parameter κ controls the thickness of the
Ac

53
54 interface between the formed phases and is linked to their interfacial tension. The third and
55
56 fourth term describe the total electrostatic energy, where ε(c) is the electric permittivity of
57
58 the material as a function of the species concentration, ρ is the mobile charge density. For
59
60 https://mc04.manuscriptcentral.com/jes-ecs
4
Page 5 of 27 Journal of The Electrochemical Society

1
2
3 the homogeneous free energy term gh we chose the regular solution model [18, 56, 57]
4

pt
5 kB T
6 gh = Ωc(1 − c) + (c ln c + (1 − c) ln(1 − c)) (2)
v
7
8 where Ω controls the interaction between the species particles - positive (negative) Ω corre-
9

cri
10 sponds to attractive (repulsive) interactions between the species. In the absence of particle-
11 particle interactions, v is the particle volume. Assuming local equilibrium, we define the
12
13 chemical potential of the neutral species as the variational derivative of the Gibbs free en-
14
15 ergy [14, 58]
16 δG 1 ∂ε

us
17 µ= = µh (c) − κ∇2 c − |∇φ|2 (3)
δc 2 ∂c
18
19 where µh = ∂gh /∂c. Also, the chemical potential of the charged species that contribute to
Fo
20
the charge density ρ reads
21
22
δG

an
µρ = e= eφ (4)
rR
23 δρ
24 Regarding the dielectric model, we assume a simple monotonically increasing phenomeno-
25
26 logical form
ev

27
28 ε = εf ε0 eγc (5)
dM
29
iew

30 where γ controls both the change and the curvature of the permittivity. This form has been
31
32 previously used to model the electric-field induced phase separation of colloidal mixtures [32,
33
59].
34
On

35
36
37 B. Transport
38
ly
pte

39
40 We consider a phase separating material with both the permittivity and conductivity to
41
42 be functions of species concentration. The equations to model the process are [38, 60, 61]
43 ∂c
44 = −∇ · j (6a)
45 ∂t
46 ∂ρ
= −∇ · jρ (6b)
ce

47
∂t
48
49 δG
=ρ−∇·D=0 (6c)
50 δφ
51 where j and jρ are the species and electronic fluxes, respectively. Based on the assumptions
52
Ac

53 of local equilibrium [14, 62] and microscopic reversibility [63, 64], the constitutive relation
54
55 for the fluxes and the dielectric displacement D are
56
57
D(c)c δG D(c)c
j=− ∇ =− ∇µ (7a)
58 kB T δc kB T
59
60 https://mc04.manuscriptcentral.com/jes-ecs
5
Journal of The Electrochemical Society Page 6 of 27

1
2
3 δG
4 jρ = −σ(c)∇ = −σ(c)∇φ (7b)

pt
δρ
5
6 D = ε(c)E = −ε(c)∇φ (7c)
7
8
9 Eqs. 6(a)-(b) are the conservative descent [65] of the free energy G with respect to c and ρ,

cri
10
11
while eq. 6(c) is an equilibrium point with respect to φ.
12 In eq. 7(b), we consider the electric flux to be described by Ohm’s law [38] and we do
13
14 not specify neither the mechanism of electric conduction nor the identity of charge carriers.
15
16 This approach is similar to the model of leaky dielectrics that has been widely used in

us
17
18
electrohydrodynamics [66–68], where the conductivity is assumed to originate from dissolved
19 ions. The model of eq. 7(b) is known to be a special limit of Poisson-Nerst-Planck type
Fo
20
21 models under large applied electric fields and low carrier densities [69, 70].
22

an
rR
23 The functional form of the diffusivity D (c) is taken that of a lattice gas model, where the
24
25 transition state is influenced by excluded volume effects leading to D (c) = D0 v (1 − c) [35].
26
ev

The permittivity is given in eq. 5. The conductivity is assumed to be linear in concentration


27
28 σ (c) = σ0 + σ1 c, where for σ1 > 0 increasing charge carriers correspond to increasing
dM
29
iew

30 conduction [38]. While being simple, this specific form of σ does not alter our conclusions.
31
32
When phase separation occurs, we assume the two formed phases to have conductivities
33 which differ by several orders of magnitude (metal-insulator/semi-conductor contact and
34
On

35 vice versa [8]). This large difference in combination with the phase separation due to electric
36
37 fields leads to dielectric breakdown in our system. Finally, we close the description of the
38
ly

system using the following boundary conditions: 1) for the species number density we assume
pte

39
40 blocking electrodes, which translates into j · n = 0 along all domain boundaries; 2) for the
41
42 potential φ, we consider φ (0) = V and φ (L) = 0 to simulate the voltage drop across the
43
44 cell, while on all the other boundaries we assume n · ∇φ = 0; 3) we assume a contact angle
45
46
of π/2 at any triple contact point between the formed phases and the boundaries of the
ce

47 system, n · ∇c = 0. In all cases, n is the outward normal vector.


48
49
50
51 C. Characteristic Scales and Material Parameters
52
Ac

53
54 The characteristic scales we consider herein are the following: (i) time tch = L2 /Dmax ,
55
56 (ii) length Lch = L, (iii) voltage φch = kB T /e, (iv) conductivity σch = σmax , (v) charge
57
58 density ρch = enmax NA , (vi) volumetric energy kB T /v, where NA is the Avogadro number.
59
60 https://mc04.manuscriptcentral.com/jes-ecs
6
Page 7 of 27 Journal of The Electrochemical Society

1
2
3 Substituting these scales in the transport equations we arrive at the following dimensionless
4

pt
5 forms  
6 ∂c D(c)
7 =∇· ∇µ (8a)
∂τ Dmax
8  
9 ∂q σ(c)

cri
10 =∇· Φ ∇φ (8b)
∂τ σmax
11
12 1
13 − ∇ · [ε(c)∇φ] = q (8c)
λ2D
14
15 where Φ = σmax kB T
, λ−2
εf ε0 kB T
Dmax nmax NA e2 D = nmax NA L2 e2
. The dimensionless number Φ quantifies the
16

us
17 ratio between the electronic and the species mobilities. An interesting limit is that of
18
19 Φ  1, where the electron/hole redistribution in the domain occurs much faster compared to
Fo
20
species diffusion, and can be thought as the continuum equivalent of the Born-Oppenheimer
21
22

an
approximation [71]. In that case, eq. 8(b) is always at quasi-equilibrium. This is not the
rR
23
24 case though when the material enters the insulating regime. Finally, the homogeneous free
25
26 energy combined with the electric energy reads
ev

27
28 gv Ωv
c (1 − c) + c ln c + (1 − c) ln(1 − c)
dM
=
29 kB T kB T
iew

(9)
30 εf ε0 vkB T γc
31 − 2 2
e |∇φ|2
32
2e L
33 Before we dive further into analyzing the implications of the applied electric field on the
34
On

35 thermodynamics stability of mixing and the dynamics of phase separation, it is instructive


36
37 to consider a specific material for the parameters of our model: εr , γ, κ, Ω, v, nmax , σmax ,
38
ly

Dmax and L. An interesting example with practical implications in Li-ion batteries and
pte

39
40 neuromorphic computing is Li4+3x Ti5 O12 [9, 72]. Based on the density and the molecular
41
42 weight of Li7 Ti5 O12 , we know that nmax ' 0.72 × 104 mol/m3 as well as v ' 1 nm3 .
43
44 Additionally, previous phase field modeling [72] has reported NA Ωv ' 8.612 kJ/mol and
45
NA κv ' 8.61×10−15 J m2 /mol. The electric permittivity is known to increasing as a function
46
ce

47 of average Li-ion fraction x. In particular, for x → 0, ε ' 1.5ε0 , while for the fully lithiated
48
49 state, x → 1, the permittivity becomes ε ' 50ε0 [73]. This behavior can be approximated by
50
51 choosing εr ' 1.5 and γ ' 3.5. In terms of tracer diffusivity, NMR studies [74–76] measured
52
D0 ' 4 × 10−16 m2 /s, while electrochemical measurements found that the conductivity
Ac

53
54 changes from σ(x → 0) ' 10−5 S/m (insulating) to σ(x → 1) ' 102 S/m (metallic) [42].
55
56 Finally, we consider a thin-film device with dimensions around L ' 100 nm, which is kept
57
58 at constant temperature T = 298 K.
59
60 https://mc04.manuscriptcentral.com/jes-ecs
7
Journal of The Electrochemical Society Page 8 of 27

1
2
3 The discussed material and system parameters result in Φ ' 107 and λ−1 −4
D ' 10 . Given
4

pt
5 this value for Φ, it is clear that when the material is metallic, we can assume eq. 8(b) to
6
7 be in quasi-equilibrium. The small value of λ−1
D ' 10
−4
corresponds to double layers on the
8
9
scale of 10−1 Å, which is a reasonable value for a perfect metal. The RC timescale τC =
q

cri
10 e2 nmax NA εr ε0 L2
for charging the formed double layers after the electric field is applied [77]
σ 2 kB T
11
12 is of the order of 5 × 10−10 s for the conductive phase and 10−2 s for the insulating one,
13
14 values much lower than the diffusive timescale of the neutral species τD = L2 /Dmax ∼ 102 s.
15
Due to numerical stability, however, when we solve eqs. 8 we use a re-scaled value for λ−1
16 D

us
17 that matches the interface thickness.
18
19 Finally, we discretize the set of eqs. 8 using finite elements [78], and more specifically, we
Fo
20
21 use continuous linear basis functions for approximating all unknowns. Additionally, we solve
22

an
the system of equations in a monolithic fashion, while for the time integration a second order
rR
23
24 scheme is used (BDF2) [79]. The non-linear system of equation is solved using Newton’s
25
26 method and for the inversion of the resulting linear system we use LU decomposition.
ev

27
28
dM
29
iew

30 III. INSTABILITY OF A HOMOGENEOUS STATE


31
32
33 A. Thermodynamics Stability
34
On

35
36 According to phase equilibria, cs,1 and cs,2 are the two spinodal points that indicate the
37 change in the curvature of the homogeneous free energy. Both of these values are solutions
38
ly
pte

39 of [14, 52]
40 ∂ 2 gh ∂µh
41 2
= =0 (10)
42
∂c ∂c
43 In the spinodal region, the homogeneous solution is unstable (gh00 < 0) and tends to phase
44
45 separate in two immiscible phases. The concentration in each phase is determined by the
46
common tangent construction
ce

47
48
49 µ(ceq,1 , |E| = 0) =µ(ceq,2 , |E| = 0) =
50
gh (ceq,1 ) − gh (ceq,2 ) (11)
51
52 ceq,1 − ceq,2
Ac

53
54 Here, we are interested in studying the electric-field induced phase separation. When the
55
56 electric field is uniform across medium, E ' −∆V ex (∆V is the dimensionless voltage drop,
57
58 scaled with the thermal voltage kB T /e), we can see from eq. 3 that the value of the chemical
59
60 https://mc04.manuscriptcentral.com/jes-ecs
8
Page 9 of 27 Journal of The Electrochemical Society

1
2
3 potential will change. The effect of the electric field on the thermodynamics stability of the
4

pt
5 homogeneous mixture, however, depends on the functional form of ε. For example, when
6
7 ε is a linear function of c, then the spinodal region is not affected. Therefore, for having
8
electric field-induced phase separation we require that ε00 (c) 6= 0. The equation for finding
9

cri
10 the spinodal points reads [9, 23, 34]
11
12
13
14 ∂ 2 gh (c) 1 ∂ 2ε
15 = |∆V |2 (12)
16 ∂c2 2 ∂c2

us
17
18
19
Fo
20 Fig. 2(a) shows the dimensionless free energy as a function of the species fraction c for
21
22 four different values of the applied electric field |∆V |. In this example we set Ω = 0, as

an
rR
23
24 we are interested to understand the implications of the electric field on the de-mixing of a
25
26 homogeneous solution. For |∆V | = 0, the energy is convex, and the mixture remains in
ev

27 the solid solution regime. With increasing |∆V |, however, the energy landscape becomes
28
dM
29 distorted, shifting the minimum energy toward c ' 0.8. For |∆V | = 400, the electric
iew

30
31 field is strong enough to change the convexity of g (dashed region), making phase separation
32
33 thermodynamically favorable. At this point, the electrostatic energy becomes comparable to
34 the entropy of mixing due to alignment of the microscopic dipoles in the medium, Figs. 2(a) &
On

35
36 (b). As it will be shown in more detail in later sections, after phase separation is completed
37
38 the system will consist of domains with low and high permittivity, respectively. Further
ly
pte

39
40 increase of the E-field increases the distance between the binodal points (red dots) which
41 leads to further increase in the dielectric mismatch between the phase separated domains,
42
43 Fig. 2(a).
44
45
These observations can be summarized in the phase diagram of Fig. 2(c), which is con-
46
ce

47 structed in terms of the applied electric field |∆V | and the system fractional concentration
48
49 c. The binodal region, which is thermodynamically metastable, is shown with blue, while
50
51 the spinodal region with light brown. It is clear that there exist a critical electric field
52
(around |∆V |c ∼ 380 for the parameters used herein) above which the convexity of the free
Ac

53
54 energy changes and the homogeneous state of the material becomes thermodynamically un-
55
56 stable. The implications of this phase diagram on the dielectric breakdown of the material
57
58 are discussed in Section IV.
59
60 https://mc04.manuscriptcentral.com/jes-ecs
9
Journal of The Electrochemical Society Page 10 of 27

1
2
3 B. Linear Stability
4

pt
5
6 The standard way to understand the dynamics during the onset of phase separation is
7
8 through linear stability analysis. In particular, given the physical parameters of our model,
9

cri
10 we can identify the critical wavelength that can be induced by random fluctuations that lead
11
to de-mixing of a homogeneous state. To do so, we assume an infinitesimal perturbation of
12
13 the form δy = δeik·x+ωt , where yT = (c, q, φ), δ is an infinitesimal vector, kT = (kx , ky , kz )
14
15 is the wavenumber, and ω is the growth rate of the instability. The base state around which
16

us
17 we linearize eqs. 8 is y0T = (c0 , 0, −∆V (1 − x)). The dispersion relation is found by solving
18 the secular equation [80]
19
Fo
20 det |J − ωe1 e1 | = 0 (13)
21
22

an
where J is the Jacobian matrix defined as J = δy f - the components of J are given in the
rR
23
24 appendix - and e1 corresponds to the unit vector along the ‘concentration’ axis. Assuming
25
26 Φ  1, the growth rate becomes
ev

27 
28 ω = − ε |∆V |2 (∂c ln σ) kx2 − D kx2 + kr2 ×
dM
29   (14)
iew

30 − (1/2) |∆V |2 ∂c2 ε + ∂c2 gh + κ kx2 + kr2


31
32 where kr2 = ky2 + kz2 , and D, σ and κ are in their dimensionless form. From eq. 14, it is clear
33
34
that the direction of the E-field affects critical wavelength in the x-direction. We can further
On

35 analyze the result by determining the set of (kx , kr ) that maximize ω. Solving ∂ω/∂k = 0,
36
37 we find r !
38 Q
ly

(kx , kr ) = 0, (15a)
pte

39
40

41  s 
2
42 DQ − 2ε |∆V | ∂c ln σ 
43 (kx , kr ) = ± ,0 (15b)
4κD
44
45
where Q = |∆V |2 ∂c2 ε − 2∂c2 gh . From the first set of solution, it is clear that for kr to be
46
ce

47 physical it has to be positive. This is true for |∆V |2 ∂c2 ε − 2∂c2 gh > 0, which is equivalent to
48
49 the thermodynamics stability condition we discuss in Sec. IIIA. On the contrary, the second
50 
51 locus of solutions is physical for D |∆V |2 ∂c2 ε − 2∂c2 gh > 2ε |∆V |2 ∂c ln σ, which shows that
52
conductivity variations can suppress phase separation in the direction of electric field when
Ac

53
54 ∂c ln σ > 0.
55
56 Fig. 3(a) demonstrates the stability diagram for a mixture with average concentration
57
58 c = 0.8 in terms of the wavenumber set (kx , kr ). The lines denote the isocontour ω = 0
59
60 https://mc04.manuscriptcentral.com/jes-ecs
10
Page 11 of 27 Journal of The Electrochemical Society

1
2
3 for different applied voltages across the domain. Inside the formed envelopes lies the region
4

pt
5 where ω > 0, which corresponds to the long-wave modes for which de-mixing occurs, while
6
7 short-waves are damped by the action of surface tension. As shown by eq. 15(b), when
8
kr = 0 there is a critical applied voltage - |∆V | . 700 below which perturbation in the x
9

cri
10 direction are suppressed.
11
12
13
14
C. Phase Separation Dynamics
15
16

us
17
In order to test the predictions of the theory and understand the time evolution of de-
18
19 mixing, we perform two-dimensional simulations of an initially homogeneous mixture with
Fo
20
21 concentration c = 0.8. As a representative example, we consider the case where the value
22

an
of the applied electric field across the domain is |∆V | = 600. According to the phase
rR
23
24
diagram of fig. 2(c), we expect the two formed phases to have concentrations cb,1 ' 0.12 and
25
26 cb,2 ' 0.998, respectively, where cb is the binodal point concentration.
ev

27
28 Fig. 3(b) shows the temporal evolution of the concentration field after the electric field
dM
29
iew

30 is applied. The light yellow color represents the rich phase, while the dark purple the low
31
32
concentration one. At time t = 0, the homogeneous profile is perturbed with zero-mean white
33 noise. After some time, these perturbations grow exponentially in time as predicted by the
34
On

35 linear stability analysis of Sec. III B. The exponential evolution of the instability stops right
36
37 after the two phases begin to form, i.e. for t = 0.1. At this moment, filament-like patterns,
38
ly

which consist of the poor and rich phase, span the entire domain. It is noticeable that the
pte

39
40 formed filaments align with the direction of the applied electric field, an observation that
41
42 connects with the phenomenon of dielectric breakdown, which is discussed in more detail in
43
44 the next section (Sec. IV). At later times, t = 1.0, the initially formed filaments with the
45
lowest concentration break into smaller islands. Due to the existence of multiple interfaces
46
ce

47 this state is not energetically favorable. As a result, the system undergoes further coarsening
48
49 (Ostwald ripening [81]) making the islands to merge into larger filamentary domains, leading
50
51 to the non-equilibrium steady state shown for t = 20.0. At this time, three large domains
52
that consist of the high concentration phase are formed, while the low concentration ones
Ac

53
54 occupy smaller fraction of the total volume. This is in agreement with the theoretically
55
56 predicted phase diagram, fig. 2(c), where the lever rule predicts that the phase with the
57
58 lowest concentration occupies ∼ 22% of the total system. Finally, when the electric field
59
60 https://mc04.manuscriptcentral.com/jes-ecs
11
Journal of The Electrochemical Society Page 12 of 27

1
2
3 is removed, the homogeneous free energy becomes convex again, leading to mixing of the
4

pt
5 two formed phases. Therefore, the recovery of a solid solution after the electric field bias is
6
7 removed corresponds to volatile behavior [2].
8
9

cri
10
11 IV. DIELECTRIC BREAKDOWN DUE TO FILAMENT FORMATION
12
13
14 The de-mixing of an initially solid-solution system demonstrates the basic principle of the
15
16 electric-field induced phase separation. Although studying these materials is informative,

us
17 many materials of practical relevance phase separate at room temperature, even in the
18
19 absence of E-field. Such examples are Lix CoO2 , Li4+3x Ti5 O12 , and Lix TiO2 where during
Fo
20
21 phase separation undergo metal-to-insulator (and vice versa) transition. Here, we focus
22

an
on the case of an initially phase separated material and show that, by applying electric
rR
23
24 field, it is possible to control the orientation of the phase boundaries and, consequently, the
25
26 current-voltage response of the material.
ev

27
28 For the ease of computations and without altering the final conclusions, we choose a
dM
29
iew

30 system with NA Ωv ' 5.601 kJ/mol, and electrical conductivity of the form σ = σ0 eσ1 c ,
31 where σ1 = 10−7 and σ1 = 16.11. All the other properties are the same as discussed in
32
33 Sec. II C. The reason for changing the functional form of σ is to replicate the abrupt change
34
On

35 in the electrical conductivity during the insulator-to-metal transition that take place in
36
37
materials like Li4+3x Ti5 O12 [42, 74].
38 Fig. 4(a) demonstrates the temporal evolution of the resulting current for three different
ly
pte

39
40 applied ∆V . The current is defined as the surface integral of the electric current density
41 R
42 across one of the electrodes, I = − J · ndA. For all the applied voltages, the initial phase
43
44 morphology corresponds to the earliest time instant shown in the inset of Fig. 4(a), while
45 the average concentration is set to c = 0.5.
46
ce

47 For ∆V = 100, the resulting current is always of the order of 10−3 , which can be under-
48
49 stood in terms of an equivalent circuit. Given the functional form for σ, we know that one
50
51 of the phases is insulating. Also, the phase morphology does not change after the voltage is
52 applied. Therefore, the equivalent circuit consists of two resistances in series, one of which
Ac

53
54 corresponds to an insulator with resistance R = ∆V /I ∼ O (105 ).
55
56 For larger applied voltages, ∆V & 200, the temporal evolution of the current is quali-
57
58 tatively different. More specifically, we focus on the phase evolution for ∆V = 200. It is
59
60 https://mc04.manuscriptcentral.com/jes-ecs
12
Page 13 of 27 Journal of The Electrochemical Society

1
2
3 apparent from the inset images that the applied electric field across the domain is able to
4

pt
5 change the morphology completely. At early times, tc < 0.27, the electric field forces the
6
7 binodal concentration to change, as shown in the phase diagram of fig. 4(b). Due to this
8
change, the system is perturbed and the interface between the two phases becomes unsta-
9

cri
10 ble forming tips in the direction of the electric field. At around tc ' 0.27, the instability
11
12 grows abruptly leading to the formation of highly conductive filaments. By the time these
13
14 filaments ‘touch’ the second electrode, the electric current increases by three orders of mag-
15
nitude, from I ' 0.01 to I ' 7. After this critical event, a dendrite-like structure is formed,
16

us
17 t ' 1, which evolves in four filaments for t & 5 - two consisting of the highly conductive
18
19 phase and two of the insulating one. The insulating domains within the dendrite-like struc-
Fo
20
21 ture demonstrate a specific angle that can be related to the Taylor cone behavior [82]. The
22

an
final configuration yields a steady state current around I ' 19, which exceeds by almost four
rR
23
24 orders of magnitude the value obtained when ∆V = 100 is applied. Therefore, we conclude
25
26 that the formation of highly electric conductive filaments after a critical voltage is applied
ev

27
28 leads to the phenomenon of dielectric breakdown.
dM
29
iew

30
The question that arises, though, is why voltages below ∆V ' 200 do not alter the
31 phase morphology into filamentary ones. The answer becomes clear when analyzing the
32
33 phase diagram of fig. 4(b). When no E-field is applied, the system is separated in two
34
On

35 immiscible phases with concentration cb,1 ' 0.21 and cb,2 ' 0.78, respectively. When a
36
voltage drop of ∆V = 100 is applied, these values lie the binodal region, which is known to
37
38 be metastable. As a result, the species are redistributed between the two phases and a new
ly
pte

39
40 steady state is reached without altering the morphology of the existing interface. However,
41
42 when ∆V = 200 is applied, the initially stable state becomes thermodynamically unstable,
43
as the concentration of the rich phase lies inside the spinodal region - the spinodal points
44
45 are cs,1 ' 0.3 and cs,2 ' 0.8. Hence, any infinitesimal perturbation tends to destabilize the
46
ce

47 system from its initial state, which, as analyzed in fig. 4(a) leads to the formation of highly
48
49 conductive filaments that align with the direction of the applied electric field.
50
51
52
Ac

53 V. DISCUSSION
54
55
56 Our findings on dielectric breakdown, as a result of filament formation due to the applied
57
58 electric field, are relevant to resistive switching and memristor devices [2, 4, 83–88]. One of
59
60 https://mc04.manuscriptcentral.com/jes-ecs
13
Journal of The Electrochemical Society Page 14 of 27

1
2
3 the desired properties of memristors is the ability to change the electrical conductivity of
4

pt
5 the device by orders of magnitude under an external stimulus, e.g. voltage or temperature.
6
7 Here, we show that through de-mixing of a solid solution mixture or change in the phase
8
morphology of an already phase separated material, we can tune the device-level resistance
9

cri
10 by orders of magnitude, fig. 4(a). This idea has been recently demonstrated experimentally
11
12 in Li4+3x Ti5 O12 , for x → 0 and x → 1, a material proposed as a potential candidate
13
14 for memristive devices [9]. It is known that Li4+3x Ti5 O12 undergoes an insulator-to-metal
15
transition [42], resulting in a change in the electric resistance of the memristive device by
16

us
17 orders of magnitude. Here, by using the properties of LTO, we show that a dimensionless
18
19 voltage of ∼ O (102 ) is required to induce de-mixing, and therefore change in the conductivity
Fo
20
21 of the device. This value corresponds to approximately 3 − 4 V, which is on the same order
22

an
of magnitude with the experimentally observed value for dielectric breakdown.
rR
23
24 Fig. 5 shows a schematic on the device current-voltage response related to the phenomena
25
26 presented in Secs. III-IV. When the system is initially prepared as solid solution, de-mixing
ev

27
28 occurs at a critical voltage ∆Vc , figs. 2(b) & 4(b). Two scenarios are possible: i) the mixture
dM
29
iew

30 shows solid solution behavior for all values of c, ii) the mixture is phase separating for a
31 wide range of concentrations. In the former situation, when the applied voltage is close to
32
33 critical value the two phases have very similar concentration, and thus, similar conductivi-
34
On

35 ties. Therefore, for resistive switching applications a voltage much larger than the critical
36
37
value needs to be applied in order to establish phases with very large mismatch in their con-
38 ductivity. For the latter case, when the critical voltage is applied and de-mixing occurs, the
ly
pte

39
40 current changes abruptly due to the large difference on the electric resistance between the
41
42 formed phases, fig. 4(a). Although dielectric breakdown is necessary for resistive switching,
43
44
memristors have the additional requirement of non-volatile operation [2]. When the electric
45 field bias is lifted, the de-mixed mixture is going to return to its initial homogeneous state
46
ce

47 under diffusive timescales. In particular, for a device with L = 100 nm [9], and a material
48
49 with maximum diffusion of ∼ 10−16 m2 /s [72], the system relaxes back to equilibrium within
50
51
∼ 100 seconds. This timescale is very short for any application that requires non-volatile
52 operation, such as neuromorphic computing.
Ac

53
54 On the contrary, mixtures that are thermodynamically unstable under no applied electric
55
56 field are non-volatile due to the persistence of the filaments after we stop applying a voltage
57
58 bias. A schematic of the representative current-voltage curve is shown with the orange line
59
60 https://mc04.manuscriptcentral.com/jes-ecs
14
Page 15 of 27 Journal of The Electrochemical Society

1
2
3 in fig. 5. At first, the system is prepared in a state where the interface between the two
4

pt
5 phases is aligned perpendicular to the direction of the electric field. As discussed earlier,
6
7 this configuration corresponds to a circuit with two resistors in series, where the one with
8
the lowest conductivity governs the I − V response of the device. At large enough voltages,
9

cri
10 though, the phase separated mixture becomes thermodynamically unstable and conductive
11
12 filaments are formed that connect the two electrodes. At this point, the macroscopic resis-
13
14 tance of the device drops by orders of magnitude causing an effective dielectric breakdown
15
- from a high resistance state (HRS) to a low resistance one (LRS) [2, 4]. After filament
16

us
17 formation, decreasing the voltage drop does not affect the phase morphology. Hence, the
18
19 persistence of the filamentary state, even after electric fields are not active, demonstrate the
Fo
20
21 potential application of such systems in neuromorphic computing.
22

an
Even though our model is able to predict the correct order of magnitude for the critical
rR
23
24 voltage that causes dielectric breakdown in LTO, it is greatly simplified. First, the functional
25
26 forms for both the electric permittivity and conductivity are purely empirical and do not
ev

27
28 describe an actual material. Therefore, a more complete theory for both ε and σ, which takes
dM
29
iew

30
into account information from first-principles calculations and/or experiments, is needed.
31 A more complete picture would identify the charged species, e.g. bound and/or free
32
33 electrons or ions, that can contribute to the electric conductivity of the medium. For ex-
34
On

35 ample, in the case of quasi-particles such as polaron-ion pairs, if the applied electric field is
36
37
strong enough, e.g. near electrode/bulk interfaces [89] or phase boundaries [27], the pairs
38 can split into its components, i.e. localized electrons and ions. Each of the newly generated
ly
pte

39
40 species can have its own conductivity [90, 91] where its diffusive/conductive motion would
41
42 be described by the corresponding conservation law.
43
44 Another effect we have neglected, which plays an important role when the system is at its
45 LRS, is Joule heating [92–97]. For nanometer scale phase change memristors, Joule heating
46
ce

47 is known to control resistive switching. Hence, the change of the local temperature due to
48
49 dissipation phenomena, such as electric conduction, is expected to affect the thermodynamics
50
51 and, consequently, the phase separation dynamics after the electric field is applied.
52
Most memristive devices are solid state in nature [2, 4, 92]. Thus, it is expected that elas-
Ac

53
54 tic and/or inelastic deformation, as well as the existence of grain boundaries and dislocations
55
56 to influence the dynamics of conductive filament formation. Additionally, phase-separating
57
58 intercalation materials are known to exhibit misfit strains which affect the morphology of
59
60 https://mc04.manuscriptcentral.com/jes-ecs
15
Journal of The Electrochemical Society Page 16 of 27

1
2
3 the formed interfaces [49, 98]. Therefore, there will be a competition between the interface
4

pt
5 orientation defined by the minimum elastic energy state and the one induced by imposed
6
7 electric field.
8
9
Finally, in our model we assumed a closed system where species concentration does not

cri
10
11 change. However, in ion intercalation materials one can change the total number of ions.
12
13 This is known to have a large impact on the phase morphology [18, 49] as well as on the
14
15 electronic conductivity of the material [8], e.g. metal-to-insulator transition. Therefore,
16

us
it would be interesting to explore the effects of species insertion/extraction on the phase
17
18 morphologies at the same time electric fields are applied. All these phenomena have to
19
Fo
20 be examined in greater detail for establishing qualitative design principles for memristive
21
22 devices that are based on the phenomenon of the electric-field induced phase separation.

an
rR
23
24
25
26
ev

27
28
dM
29
iew

30
31 VI. SUMMARY
32
33
34
On

35 In summary, we showed that when electric field is applied in a material with concentration-
36
37 dependent permittivity and electric conductivity, phase separation occurs and dielectric
38
ly

breakdown is observed. Through thermodynamics stability analysis we derived phase di-


pte

39
40 agrams in terms of the species concentration and the applied voltage drop between the
41
42 operating electrodes, and we demonstrated that one can de-mix a solid solution mixture.
43
44 Additionally, by performing simulations we predicted that once the system is thermodynam-
45
ically unstable, filament-like structures are formed. These structures percolate across the
46
ce

47 domain and are responsible for the dielectric breakdown by allowing electrons to conduct
48
49 through the metallic phase. Furthermore, we demonstrated the predictions of the theory
50
51 to be in agreement with recent experiments on Li4+3x Ti5 O12 . Finally, we discussed the
52
implications of our results on resistive switching, which can be useful in applications like
Ac

53
54 neuromorphic computing. In particular, we showed that phase separating materials can
55
56 exhibit the desired non-volatile behavior while solid solution materials do not, as they relax
57
58 back to their equilibrium state after the electric field is turned off.
59
60 https://mc04.manuscriptcentral.com/jes-ecs
16
Page 17 of 27 Journal of The Electrochemical Society

1
2
3 ACKNOWLEDGMENT
4

pt
5
6 The authors would like to thank Tao Gao, Neel Nadkarni, Juan Carlos Gonzalez-Rosillo,
7
8 Moran Balaish, and Jennifer L. M. Rupp for insightful discussions.
9

cri
10
11
12 APPENDIX
13
14
15 Jacobian Matrix
16

us
17
18 The components of the Jacobian matrix discussed in Sec. III B are presented here. More
19 specifically,
Fo
20  
21
22
J1,1 0 J1,3 
 

an
J = J2,1 0 J2,3  (16)
rR
23  
24
J3,1 J3,2 J3,3
25
26
ev

where  
27
28 2 2 2 1 2 2
J1,1 = −k Dc ∂c gh + κk − ∂c ε |∇φ| (17a)
dM
29 2
iew

30
31 J1,3 = i∂c εk 2 k · ∇φ (17b)
32
33 J2,1 = i∂c σk · ∇φ (17c)
34
On

35
J2,3 = −σk 2 (17d)
36
37
38 J3,1 = −i∂c εk · ∇φ (17e)
ly
pte

39
40 J3,2 = λ2D (17f)
41
42
43
J3,3 = εk 2 (17g)
44
45
46
ce

47
48
49 [1] J. Lim, Y. Li, D. H. Alsem, H. So, S. C. Lee, P. Bai, D. A. Cogswell, X. Liu, N. Jin, Y.-s. Yu,
50
N. J. Salmon, D. A. Shapiro, M. Z. Bazant, T. Tyliszczak, and W. C. Chueh, Science 353,
51
52 566 (2016).
Ac

53
54 [2] D. Ielmini and H.-S. P. Wong, Nature Electronics 1, 333 (2018).
55
56 [3] E. J. Fuller, F. E. Gabaly, F. Léonard, S. Agarwal, S. J. Plimpton, R. B. Jacobs-Gedrim,
57
58 C. D. James, M. J. Marinella, and A. A. Talin, Advanced Materials 29, 1604310 (2017).
59
60 https://mc04.manuscriptcentral.com/jes-ecs
17
Journal of The Electrochemical Society Page 18 of 27

1
2
3 [4] R. Waser, R. Dittmann, S. Menzel, and T. Noll, Faraday discussions 213, 11 (2019).
4

pt
5 [5] S. Zafeiratos, S. Piccinin, and D. Teschner, Catalysis Science & Technology 2, 1787 (2012).
6
7 [6] M. Hildebrand, M. Kuperman, H. Wio, A. Mikhailov, and G. Ertl, Physical review letters
8
9 83, 1475 (1999).

cri
10
[7] A. S. Mikhailov and G. Ertl, ChemPhysChem 10, 86 (2009).
11
12 [8] N. Nadkarni, T. Zhou, D. Fraggedakis, T. Gao, and M. Z. Bazant, Advanced Functional
13
14 Materials 29, 1902821 (2019).
15
16 [9] J. C. Gonzalez-Rosillo, M. Balaish, Z. D. Hood, N. Nadkarni, D. Fraggedakis, K. J. Kim,

us
17
18 K. M. Mullin, R. Pfenninger, M. Z. Bazant, and J. L. Rupp, Advanced Materials 32, 1907465
19
(2020).
Fo
20
21 [10] Q. Lu, S. Huberman, H. Zhang, Q. Song, J. Wang, G. Vardar, A. Hunt, I. Waluyo, G. Chen,
22

an
rR
23 and B. Yildiz, Nature Materials , 1 (2020).
24
25 [11] D. T O’Connor, M. J. Welland, W. K. Liu, and P. W. Voorhees, Modelling and Simulation
26
ev

27 in Materials Science and Engineering 24, 035020 (2016).


28
dM
29
[12] C. V. Di Leo, E. Rejovitzky, and L. Anand, International Journal of Solids and Structures
iew

30 67, 283 (2015).


31
32 [13] R. Koerver, W. Zhang, L. De Biasi, S. Schweidler, A. O. Kondrakov, S. Kolling, T. Brezesinski,
33
34 P. Hartmann, W. G. Zeier, and J. Janek, Energy and Environmental Science 11, 2142 (2018).
On

35
36 [14] D. Kondepudi and I. Prigogine, Modern thermodynamics: from heat engines to dissipative
37
38
structures (John Wiley & Sons, 2014).
ly
pte

39 [15] R. W. Balluffi, S. M. Allen, and W. C. Carter, Kinetics of materials (John Wiley & Sons,
40
41 2005).
42
43 [16] M. Mirzadeh and M. Z. Bazant, Physical review letters 119, 174501 (2017).
44
45 [17] T. Gao, M. Mirzadeh, P. Bai, K. M. Conforti, and M. Z. Bazant, Nature communications 10,
46
1 (2019).
ce

47
48 [18] P. Bai, D. A. Cogswell, and M. Z. Bazant, Nano letters 11, 4890 (2011).
49
50 [19] M. Z. Bazant, Faraday discussions 199, 423 (2017).
51
52 [20] D. Fraggedakis and M. Z. Bazant, The Journal of Chemical Physics 152, 184703 (2020).
Ac

53
54 [21] J. M. Carmack and P. C. Millett, Soft matter 14, 4344 (2018).
55
56 [22] X. Gu, W. Liu, and K. Liang, Materials Science and Engineering: A 278, 22 (2000).
57
[23] D. D. Thornburg and R. M. White, Journal of Applied Physics 43, 4609 (1972).
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
18
Page 19 of 27 Journal of The Electrochemical Society

1
2
3 [24] H. Hori, O. Urakawa, O. Yano, and Q. Tran-Cong-Miyata, Macromolecules 40, 389 (2007).
4

pt
5 [25] I. Aranson, B. Meerson, P. Sasorov, and V. Vinokur, Physical review letters 88, 204301
6
7 (2002).
8
9 [26] Y. Tsori and L. Leibler, Proceedings of the National Academy of Sciences of the United States

cri
10
of America 104, 7348 (2007).
11
12 [27] Y. Tsori, F. Tournilhac, and L. Leibler, Nature 430, 544 (2004).
13
14 [28] Z.-G. Wang, The Journal of Physical Chemistry B 112, 16205 (2008).
15
16 [29] B. Khusid and A. Acrivos, Physical Review E 54, 5428 (1996).

us
17
18 [30] B. Khusid and A. Acrivos, Physical Review E 60, 3015 (1999).
19
[31] A. Kumar, Z. Qiu, A. Acrivos, B. Khusid, and D. Jacqmin, Physical Review E 69, 021402
Fo
20
21 (2004).
22

an
rR
23 [32] M. D. Johnson, X. Duan, B. Riley, A. Bhattacharya, and W. Luo, Physical Review E -
24
25 Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics 69, 7 (2004).
26
ev

27 [33] C. Liedel, C. W. Pester, M. Ruppel, V. S. Urban, and A. Böker, Macromolecular Chemistry


28
dM
29
and Physics 213, 259 (2012).
iew

30 [34] J. Tomlinson, Journal of Electronic Materials 1, 357 (1972).


31
32 [35] M. Z. Bazant, Accounts of chemical research 46, 1144 (2013).
33
34 [36] A. L. Kupershtokh and D. A. Medvedev, Physical Review E 74, 021505 (2006).
On

35
36 [37] C. Liedel, C. W. Pester, M. Ruppel, V. S. Urban, and A. Böker, Macromolecular Chemistry
37
38
and Physics 213, 259 (2012).
ly
pte

39 [38] S. M. Sze and K. K. Ng, Physics of semiconductor devices (John wiley & sons, 2006).
40
41 [39] J. J. O’Dwyer, The theory of electrical conduction and breakdown in solid dielectrics (Claren-
42
43 don Press, 1973).
44
45 [40] A. Milewska, K. Świerczek, J. Tobola, F. Boudoire, Y. Hu, D. Bora, B. Mun, A. Braun, and
46
J. Molenda, Solid State Ionics 263, 110 (2014).
ce

47
48 [41] R. Amin and Y.-M. Chiang, Journal of The Electrochemical Society 163, A1512 (2016).
49
50 [42] M. G. Verde, L. Baggetto, N. Balke, G. M. Veith, J. K. Seo, Z. Wang, and Y. S. Meng, ACS
51
52 nano 10, 4312 (2016).
Ac

53
54 [43] D. Young, A. Ransil, R. Amin, Z. Li, and Y.-M. Chiang, Advanced energy materials 3, 1125
55
56 (2013).
57
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
19
Journal of The Electrochemical Society Page 20 of 27

1
2
3 [44] D. A. Goldhammer, Dispersion und Absorption des Lichtes in ruhenden isotropen Körpern:
4

pt
5 Theorie und ihre Folgerungen, Vol. 16 (Teubner, 1913).
6
7 [45] K. Herzfeld, Physical Review 29, 701 (1927).
8
9 [46] A. Manthiram, Nature Communications 11, 1 (2020).

cri
10
[47] N. Nitta, F. Wu, J. T. Lee, and G. Yushin, Materials Today 18, 252 (2015),
11
12 arXiv:arXiv:1011.1669v3.
13
14 [48] Y. Li, H. Chen, K. Lim, H. D. Deng, J. Lim, D. Fraggedakis, P. M. Attia, S. C. Lee, N. Jin,
15
16 J. Moškon, Z. Guan, W. E. Gent, J. Hong, Y. S. Yu, M. Gaberšček, M. S. Islam, M. Z. Bazant,

us
17
18 and W. C. Chueh, Nature materials 17, 915 (2018).
19
[49] N. Nadkarni, E. Rejovitsky, D. Fraggedakis, C. V. Di Leo, R. B. Smith, P. Bai, and M. Z.
Fo
20
21 Bazant, Physical Review Materials 2, 085406 (2018).
22

an
rR
23 [50] N. J. de Klerk, A. Vasileiadis, R. B. Smith, M. Z. Bazant, and M. Wagemaker, Physical
24
25 Review Materials 1, 025404 (2017).
26
ev

27 [51] Y. Li, E. J. Fuller, S. Asapu, S. Agarwal, T. Kurita, J. J. Yang, and A. A. Talin, ACS Applied
28
dM
29
Materials & Interfaces 11, 38982 (2019).
iew

30 [52] P. M. Chaikin, T. C. Lubensky, and T. A. Witten, Principles of condensed matter physics,


31
32 Vol. 10 (Cambridge university press Cambridge, 1995).
33
34 [53] J. D. van der Waals, Journal of Statistical Physics 20, 200 (1979).
On

35
36 [54] J. W. Cahn and J. E. Hilliard, The Journal of chemical physics 31, 688 (1959).
37
38
[55] J. W. Cahn and J. E. Hilliard, The Journal of chemical physics 28, 258 (1958).
ly
pte

39 [56] J. H. Hildebrand, , and R. Scott, Solubility of Non-electrolytes (Rheinhold Publishing Co.,


40
41 New York, 1936).
42
43 [57] T. Zhou, M. Mirzadeh, D. Fraggedakis, R. J.-M. Pellenq, and M. Z. Bazant, arXiv preprint
44
45 arXiv:1909.09332 (2019).
46
[58] I. M. Gelfand, R. A. Silverman, and S. Fomin, Calculus of variations (Courier Corporation,
ce

47
48 2000).
49
50 [59] C. Brosseau, Journal of applied physics 75, 672 (1994).
51
52 [60] W. M. Deen, Analysis of transport phenomena (Oxford University Press New York, 1998).
Ac

53
54 [61] J. D. Jackson, Classical electrodynamics (John Wiley & Sons, 2007).
55
56 [62] S. R. De Groot and P. Mazur, Non-equilibrium thermodynamics (Courier Corporation, 2013).
57
[63] L. Onsager, Physical review 37, 405 (1931).
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
20
Page 21 of 27 Journal of The Electrochemical Society

1
2
3 [64] L. Onsager, Physical review 38, 2265 (1931).
4

pt
5 [65] S. Adams, N. Dirr, M. Peletier, and J. Zimmer, Philosophical Transactions of the Royal
6
7 Society A: Mathematical, Physical and Engineering Sciences 371, 20120341 (2013).
8
9 [66] J. Melcher and G. Taylor, Annual review of fluid mechanics 1, 111 (1969).

cri
10
[67] D. Saville, Annual review of fluid mechanics 29, 27 (1997).
11
12 [68] J. R. Melcher, Continuum electromechanics, Vol. 2 (MIT press Cambridge, MA, 1981).
13
14 [69] O. Schnitzer and E. Yariv, Journal of Fluid Mechanics 773, 1 (2015).
15
16 [70] M. Z. Bazant, Journal of Fluid Mechanics 782, 1 (2015).

us
17
18 [71] C. Kittel, Introduction to solid state physics, Vol. 8 (Wiley New York, 1976).
19
[72] A. Vasileiadis, N. J. de Klerk, R. B. Smith, S. Ganapathy, P. P. R. Harks, M. Z. Bazant, and
Fo
20
21 M. Wagemaker, Advanced Functional Materials 28, 1705992 (2018).
22

an
rR
23 [73] Y. Liu, J. Lian, Z. Sun, M. Zhao, Y. Shi, and H. Song, Chemical Physics Letters 677, 114
24
25 (2017).
26
ev

27 [74] M. Wagemaker, E. R. van Eck, A. P. Kentgens, and F. M. Mulder, The Journal of Physical
28
dM
29
Chemistry B 113, 224 (2009).
iew

30 [75] W. Schmidt, P. Bottke, M. Sternad, P. Gollob, V. Hennige, and M. Wilkening, Chemistry of


31
32 Materials 27, 1740 (2015).
33
34 [76] M. Wilkening, W. Iwaniak, J. Heine, V. Epp, A. Kleinert, M. Behrens, G. Nuspl, W. Bensch,
On

35
36 and P. Heitjans, Physical Chemistry Chemical Physics 9, 6199 (2007).
37
38
[77] M. Z. Bazant, K. Thornton, and A. Ajdari, Physical review E 70, 021506 (2004).
ly
pte

39 [78] D. Fraggedakis, J. Papaioannou, Y. Dimakopoulos, and J. Tsamopoulos, Journal of Compu-


40
41 tational Physics 344, 127 (2017).
42
43 [79] D. Fraggedakis, C. Kouris, Y. Dimakopoulos, and J. Tsamopoulos, Physics of Fluids 27,
44
45 082102 (2015).
46
[80] L. G. Leal, Advanced transport phenomena: fluid mechanics and convective transport processes,
ce

47
48 Vol. 7 (Cambridge University Press, 2007).
49
50 [81] P. W. Voorhees, Journal of Statistical Physics 38, 231 (1985).
51
52 [82] G. I. Taylor, Proceedings of the Royal Society of London. Series A. Mathematical and Physical
Ac

53
54 Sciences 280, 383 (1964).
55
56 [83] Y. Zhou and S. Ramanathan, Proceedings of the IEEE 103, 1289 (2015).
57
[84] S. Z. Bisri, S. Shimizu, M. Nakano, and Y. Iwasa, Advanced Materials 29, 1 (2017).
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
21
Journal of The Electrochemical Society Page 22 of 27

1
2
3 [85] J. S. Lee, S. Lee, and T. W. Noh, Applied Physics Reviews 2 (2015), 10.1063/1.4929512.
4

pt
5 [86] Z. Wang, H. Wu, G. W. Burr, C. S. Hwang, K. L. Wang, Q. Xia, and J. J. Yang, Nature
6
7 Reviews Materials (2020), 10.1038/s41578-019-0159-3.
8
9 [87] W. Sun, B. Gao, M. Chi, Q. Xia, J. J. Yang, H. Qian, and H. Wu, Nature Communications

cri
10
10, 1 (2019).
11
12 [88] S. Green and J. B. Aimone, Nature Electronics 2, 96 (2019).
13
14 [89] A. C. Luntz, J. Voss, and K. Reuter, “Interfacial challenges in solid-state li ion batteries,”
15
16 (2015).

us
17
18 [90] T. Maxisch, F. Zhou, and G. Ceder, Physical review B 73, 104301 (2006).
19
[91] R. Malik, A. Abdellahi, and G. Ceder, Journal of the electrochemical society 160, A3179
Fo
20
21 (2013).
22

an
rR
23 [92] K. M. Kim, D. S. Jeong, and C. S. Hwang, Nanotechnology 22, 254002 (2011).
24
25 [93] D. Ielmini, IEEE Transactions on Electron Devices 58, 4309 (2011).
26
ev

27 [94] J. P. Strachan, D. B. Strukov, J. Borghetti, J. J. Yang, G. Medeiros-Ribeiro, and R. S.


28
dM
29
Williams, Nanotechnology 22, 254015 (2011).
iew

30 [95] M. D. Pickett, G. Medeiros-Ribeiro, and R. S. Williams, Nature materials 12, 114 (2013).
31
32 [96] W. Wang, M. Laudato, E. Ambrosi, A. Bricalli, E. Covi, Y. H. Lin, and D. Ielmini, IEEE
33
34 Transactions on Electron Devices 66, 3795 (2019).
On

35
36 [97] W. Wang, M. Laudato, E. Ambrosi, A. Bricalli, E. Covi, Y. H. Lin, and D. Ielmini, IEEE
37
38
Transactions on Electron Devices 66, 3802 (2019).
ly
pte

39 [98] D. A. Cogswell and M. Z. Bazant, ACS nano 6, 2215 (2012).


40
41
42
43
44
45
46
ce

47
48
49
50
51
52
Ac

53
54
55
56
57
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
22
pt
electrode 1 electrode 2

e- ion
p1
<latexit sha1_base64="(null)">(null)</latexit>

"1 <latexit sha1_base64="(null)">(null)</latexit>

cri
e- ion
p2 <latexit sha1_base64="(null)">(null)</latexit>

"2
<latexit sha1_base64="(null)">(null)</latexit>

L
<latexit sha1_base64="(null)">(null)</latexit>
<latexit

us
<latexit sha1_base64="(null)">(null)</latexit>
<latexit

FIG. 1. Schematic representation of a phase separating dielectric medium of length L placed


between two ion blocking electrodes. Each of the formed phases has different dielectric ε and

an
conductive σ properties. The electric permittivity for each of the phases is described by a simple
model based on two overlapping spheres of negative (green bound e− ) and positive (red ion) charges.
The application of electric field E induce polarization pi in each of the materials.
dM
pte
ce
Ac

23
Journal of The Electrochemical Society Page 24 of 27

1
2
3
4 (a) (b) (c)
= 300

pt
5
| V | = 0.0
6

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
7
300
8
gv/kB T

9 Binodal

| V|
400 Spinodal

cri
Point 400
10

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
11 500
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
12 Common 500 Binodal

13 Tangent

14
15
16 c c c

us
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
latexit
<
17
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
latexit
<
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
latexit
<
18
19 FIG. 2. (a) Homogeneous free energy diagram as a function of the species concentration. Different
Fo
20
21 curves correspond to different applied dimensionless voltage drop ∆V (scaled to kB T /e) between
22
the two electrodes. With increasing electric field, the free energy loses its convexity, indicating

an
rR
23
24
the formation of two phases with concentrations determined by the common tangent construction.
25
26
ev

(b) The electrostatic contribution to the free energy for three different values of dimensionless
27
28 voltage drop ∆V . (c) Thermodynamics phase diagram in terms of the species concentration and
dM
29
iew

30 the applied voltage drop ∆V . The light brown region corresponds to the miscibility gap, which
31
32 changes with increasing voltage drop, and the blue region is the binodal region, where the solution
33
is metastable. The present phase diagram is generated using the model of eq. 9 with Ω = 0.
34
On

35
36
37
38
ly
pte

39
40
41
42
43
44
45
46
ce

47
48
49
50
51
52
Ac

53
54
55
56
57
58
59
60 https://mc04.manuscriptcentral.com/jes-ecs
24
9
8
7
6
5
4
3
2
1

60
59
58
57
56
55
54
53
52
51
50
49
48
47
46
45
44
43
42
41
40
39
38
37
36
35
34
33
32
31
30
29
28
27
26
25
24
23
22
21
20
19
18
17
16
15
14
13
12
11
10
Page 25 of 27

(a)
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
kr

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
latexit
<
Ac !>0
c = 0.8

(Unstable)

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
ce
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
kx
400
sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
500
800

600
| V | = 1000
!<0

align with the applied electric field.


pte Fo
=0
(Stable)

rR
(b)

25
dM iew
ev
t = 1.0
t = 0.0

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
<latexit
Journal of The Electrochemical Society

an

https://mc04.manuscriptcentral.com/jes-ecs
On q
| V |= 600

ly us
t = 0.1

t = 20.0

sha1_base64="(null)">(null)</latexit>
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
sha1_base64="(null)">(null)</latexit><latexit
latexit
<
cri
c

formed filaments accumulate into larger islands which at the end merge to form filaments that
into multiple filaments with with concentrations 0.12 and 0.998. As time increases, the initially
state at c = 0.8. A voltage drop of |∆V | = 600 is applied. At first, the homogeneous state splits
unstable for a larger set of (kx , kr ). (b) Temporal evolution of an initially unstable homogeneous
described in eq. 9 with Ω = 0. With increasing electric field, the homogeneous state becomes
curves correspond to different values of the applied electric field. The thermodynamics model is
rate ω = 0 for different values of the wave numbers kx and kr = ky2 + kz2 . Different colored
an electric field E. The contour lines correspond to the locus where the non-dimensional growth
FIG. 3. (a) Linear stability analysis of a homogeneous mixture with c = 0.8 under the influence of

pt
(a) c = 0.5 300
(b)
Dielectric

pt
<latexit sha1_base64="(null)">(null)</latexit>
<latexit

Breakdown
J · ndA 200

cri
Z

| V|
I=
<latexit sha1_base64="(null)">(null)</latexit>

<latexit sha1_base64="(null)">(null)</latexit>
Current

| Vc |
<latexit sha1_base64="(null)">(null)</latexit>

us
| V |= 100
<latexit sha1_base64="(null)">(null)</latexit>

Binodal

Time c
<latexit sha1_base64="(null)">(null)</latexit>
<latexit

R
FIG. 4. (a) Temporal evolution of macroscopic current defined as I = − σE · ndA, in response to
an applied voltage |∆V | (scaled to kB T /e), indicated by different colors. The system is initially at a
phase separated state as shown in the inset for the earliest time instance. The average concentration
an
dM
is c = 0.5. At low voltage drop, e.g. |∆V | = 100, the initial morphology remains intact. For |∆V |
above the critical value, any interface perpendicular to the applied E-field is unstable. This leads to
the formation of filaments that connect the two electrodes and increase the electrical conductivity of
the device by orders of magnitude, causing effectively dielectric breakdown. (b) Thermodynamics
phase diagram in terms of the species concentration and the applied voltage drop ∆V . The light
brown region corresponds to the miscibility gap, which changes with increasing voltage drop, and
pte

the blue region is the binodal region, where the solution is metastable. The critical voltage drop
∆Vc is defined as the value of |∆V | that shifts one of the binodal points at zero bias into the
spinodal region. For both the simulations and the phase diagram, NA Ωv = 5.601 kJ/mol, while
the electric conductivity is described by σ = σ0 eσ1 c , where σ1 = 10−7 and σ1 = 16.11.
ce
Ac

26
pt
Phase
Separated

cri
Current
Solid
Solution

us
| Vc |
<latexit sha1_base64="(null)">(null)</latexit>

Voltage

an
FIG. 5. Schematic of the device-level current-voltage response for a mixture with electric permit-
tivity that depends on species concentration ε (c). Both curves correspond to the parameters used
in Secs. III-IV. The blue curve corresponds to a system initially prepared as a solid solution mix-
ture, while the orange line is that for a phase separated system. The arrows indicate the direction
dM
of the voltage sweep. In both cases, there is a critical applied voltage above which filaments are
formed between the two electrodes and the resistance of the device drops by orders of magnitude.
This phenomenon corresponds to resistive switching, which is the key operation for the operation
of memristors.
pte
ce
Ac

27

You might also like