You are on page 1of 6

Electrochimica Acta xxx (2018) xxx-xxx

Contents lists available at ScienceDirect

Electrochimica Acta

F
journal homepage: www.elsevier.com

OO
Fundamental parameters governing ion conductivity in polymer electrolytes
A. Kisliuka, V. Bocharovaa, I. Popova, C. Gainarub, A.P. Sokolova, c, ∗
a
Chemical Sciences Division, Oak Ridge National Laboratory, Oak Ridge, TN, 37831, United States
b
Fakultät Physik, Technische Universität Dortmund, D-44221, Dortmund, Germany
c

PR
Department of Chemistry, University of Tennessee, Knoxville, TN, 37996, United States

ARTICLE INFO ABSTRACT

Article history: We analyze conductivity of polymerized ionic liquids with focus on fundamental limitations hindering faster
Received 25 October 2018 charge transport in polymer electrolytes. We emphasize that to achieve the required ionic conductivity
Received in revised form 19 December 2018 ∼10−3 S/cm in dry polymer electrolytes, a decoupling of ion transport from segmental dynamics is required.
Accepted 25 December 2018

ED
We demonstrate that two competing mechanisms control decoupling of ion transport: electrostatic interactions
Available online xxx
that dominates for small ions such Li, and elastic force that dominates for large ions. Our experimental results
indeed confirm significant contribution of the elastic force to the energy barrier controlling transport of large
Keywords: ions. We also emphasize importance of ion-ion correlations that strongly affect charge transport (conductiv-
Polymer electrolytes ity) even at the same ion diffusivity. Our analysis suggests that these correlations suppress ion conductivity
Ion conductivity in polymer electrolytes by about ten times. At the end, we formulate some ideas on design of polymer elec-
Shear modulus trolytes with high ion conductivity.
Haven ratio
CT
© 2018.

1. Introduction 2. What controls ionic conductivity in polymer electrolytes

It is widely recognized that use of polymer electrolytes instead of We start with the analysis of ionic conductivity σ using classical
traditional liquid electrolytes will significantly improve performance Nernst-Einstein equation [7,8]:
RE

and safety of current batteries [1–4]. Among polymer electrolytes,


polymerized ionic liquids (PolyILs), where one ion remains mobile
while the counter ion is attached to the polymer chain, attract signif- (1)
icant attention due to their single-ion conductivity [5,6]. PolyILs do
not require any addition of salts for conductivity because they have
intrinsic ions. Single ion conductance is beneficial for many applica- Here n is the conducting ion concentration, q is their charge, D is
tions, including flow and conventional batteries. It also simplifies data
R

their diffusion coefficient presented as a random-walk terms with the


analysis and theoretical treatment of PolyILs, and make them ideal length of the ion jumps λ divided by the time between ion jumps τi
model systems to study microscopic parameters controlling conduc- (inverse ion jumps rate) and by 6, which accounts for dimensional-
tivity. ity [8]. Mobile ions concentration in PolyILs is usually in the range
CO

In this contribution we discuss the fundamental parameters control- 2–5 nm−3 [9]. Ion jump length in condensed matter should be λ ∼
ling conductivity in polymer electrolytes. We emphasize that decou- 1–2 Å. Using eq. (1) and assuming that all mobile ions contribute to
pling of ion transport from segmental (structural) relaxation is criti-
the conductivity, we can estimate the rate of ion jumps needed to
cal for achieving high ionic conductivity in dry polymer electrolytes.
achieve ion conductivity σ∼10−3 S/cm, usually presented as a mini-
There are two contributions, electrostatic and elastic forces, control-
mum value required for many applications [10]. To achieve this value
ling the decoupled ion transport. We demonstrate that reduction in
corresponding characteristic time between the ion jump should be
elastic force contribution indeed decreases the energy barrier for ion
τi ∼ (0.4–4)*10−10s–40–400 ps. According to the classical approxima-
UN

conductivity in glassy PolyILs. In addition ion-ion correlations signif-


tion for ion transport in liquids, τi is defined by the characteristic struc-
icantly affect charge diffusion and thus the conductivity. This effect is
not actively discussed in literature, but according to our analysis it re- tural relaxation time τα, i.e. τi ∼ τα [11]. The same is assumed to be
duces ionic conductivity in PolyILs by almost 10 times. valid for polymer electrolytes above their glass transition temperature
(Tg) with τα being segmental relaxation time [11,12]. While so fast
structural relaxation time at room temperature is easily achievable for

Corresponding author. Chemical Sciences Division, Oak Ridge National Laboratory, many liquid electrolytes (e.g. τα in water is ∼1 ps [13,14]), it is not
Oak Ridge, TN, 37831, United States.
feasible for the segmental relaxation in dry polymer electrolytes.
Email address: sokolov@utk.edu (A.P. Sokolov)

https://doi.org/10.1016/j.electacta.2018.12.143
0013-4686/ © 2018.
2 Electrochimica Acta xxx (2018) xxx-xxx

The best way to accelerate segmental dynamics is to decrease the


glass transition temperature, and many efforts were focused on devel-
oping polymer electrolytes with reduced Tg. Recent studies [15–17]
suggested that Tg in PolyILs depends on the size of structural units
(monomer plus mobile ion), dielectric constant, and chain rigidity. De-

F
tailed analysis emphasized that it will be difficult, if not impossible
to develop dry PolyIL with Tg below 200 K [17]. In that case, it will
be impossible to have segmental relaxation time faster than ∼10−10 s.

OO
However, it was shown that plasticizing polymer with a solvent can
effectively speed up ion dynamics reaching the desired rate of the ion
transport. As a result, there are many polymer gels that provide suffi-
cient level of ion conductivity, although they lose many advantages of
solid polymer electrolytes [18,19].
There is, however, another mechanism of ion transport that domi-
nates in superionic crystals and glasses [20–22]. In this case ion jumps
in a frozen structure and the rate of their jumps is controlled by a char-

PR
acteristic energy barrier Eσ separating ion sites, τi = τ0*exp(Eσ/kT),
where τ0∼10−12-10−14 s is a characteristic attempt time. Indeed, decou-
pling of ion transport from segmental dynamics has been observed
in crystalline poly(ethylene oxide) (PEO) [23–30], where segmental Fig. 1. Temperature dependence of ion conductivity for PolyILs clearly demonstrates a
mobility is essentially frozen, while ions still exhibit decent mobility. crossover from a Vogel-Fulcher-Tamman (VFT) behavior to an Arrhenius-like behav-
ior at T ∼ Tg. Data are from Refs. [9,37], where detailed description of the PolyILs is
Analysis demonstrates that Li ions move inside the PEO helix [23–25]
presented. The mobile ions are TFSI− and Br−. The structures of polymethylhydrosilox-
with the energy barrier that depends on the counterion [24,26,27]. The ane-graft-5-imidazolium-1-pentene trifluoromethanesulfonimide (polyMHSIm TFSI−),

ED
latter is positioned between the chain helixes [23,24]. Using shorter polyN-vinyl ethyl imidazolium trifluoromethanesulfonimide (PolyEtVIm TFSI−),
PEO chains [24] or orienting PEO chains by stretching or external polyN-vinyl diethylene glycol ethyl methyl ether imidazolium bromide and trifluo-
romethanesulfonimide (PolyEGVIm Br− and TFSI−) are presented in Table 1.
magnetic field [25,28–30] leads to an increase of ion conductivity,
emphasizing again a decoupling of ion motion from segmental dy-
namics. Decoupling of ion conductivity from segmental dynamics has
been also reported for many amorphous polymer electrolytes, includ-
ing PolyILs [12,31–34]. Understanding microscopic parameters that
CT
Table 1
can increase the decoupling and thus enhance ion transport is critical Chemical structure of PolyILs presented in Figs. 1 and 3.
for achieving high ionic conductivity in polymer electrolytes, and we
will discuss possible mechanisms below.

3. Mechanisms controlling decoupling of ion transport from


segmental dynamics
RE

For simplicity we will start with the analysis of ionic conductiv-


ity in PolyILs at temperatures below Tg, i.e. in their glassy state. In
this case segmental relaxation is essentially frozen and ions moves by
over-barrier jumps. Indeed polymer electrolytes show an Arrhenius
temperature dependence of conductivity at T < Tg, σ = σ0*exp(-Eσ/kBT)
R

[16,35,36] (Fig. 1). Thus, studies of ionic conductivity in polymer


electrolytes at T < Tg might reveal microscopic details controlling de-
coupling and the activation energy barrier for ion transport. Advantage
CO

of PolyILs in this study is that only mobile ions provide conductivity


while counter ions are attached to the chain and cannot move at T < Tg.
Detailed analysis of literature data for energy barrier controlling
conductivity below Tg in PolyILs was presented in Ref. [9]. It an-
alyzed polycations with mobile anions and polyanions with mobile
cations. We emphasize that no salt was added to PolyILs in these
studies, and the measured conductivity is intrinsic property of Poly-
UN

ILs provided exclusively by the mobile ions. This analysis reveals that
Eσ has non-monotonous dependence on the size of the mobile ions
Rion: It goes through a minimum, and strongly increases at smaller
and larger Rion (Fig. 2). This behavior can be explained by two com- Fig. 2. Dependence of the activation energy for conductivity at T < Tg on the mobile ion
peting mechanisms controlling ion transport: electrostatic ion-ion in- size in polymerized ionic liquids, mobile ions are named in the Figures. The data are
from Ref. [9], and references therein. The dashed line shows contribution of electrosta-
teractions and elastic force. The electrostatic contribution decreases tic interactions with ε = 6 and R1+R2 = 2*Rion. Figure (A) presents elastic force follow-
with increase in Rion while the elastic part increases. This idea has ing shoving model with αG∞ = 1 GPa and 1.5 GPa (eq. (4)), while figure (B) presents
been proposed by Anderson and Stuart to describe ion con Anderson-Stuart Model with G = 1 GPa and 1.5 GPa, λ∼1 Å and RD = 0 (eq. (3)). The
solid lines show the sum of electrostatic and elastic force contributions for the corre-
sponding models.
Electrochimica Acta xxx (2018) xxx-xxx 3

These simplistic models indeed provide a good description of the


dependence of Eσ on the mobile ion radius Rion [9,38,41], with rea-
sonable for polymer electrolytes parameters ε = 6, and αG∞ or G
∼1–1.5 GPa (Fig. 2). These two model approaches provide different
dependence of the elastic energy barrier on the mobile ion size Rion:
It should increase as ∼ Rion2 in Anderson-Stuart model (eq. (3)), while

F
it should increase as ∼ Rion3 in shoving model (eq. (4)). However, both
models predict the same dependence on the shear modulus.

OO
To verify the latter prediction we measured Brillouin Light Scat-
tering (BLS) spectra of several PolyILs. The synthesis and charac-
terizations of these PolyILs have been described in Ref. [17]. BLS
spectra were measured in symmetric geometry at 90° angle using
a tandem Fabry-Pérot interferometer and a solid state laser (Verdi,
λlaser = 532 nm). The samples were measured in cryostat Janis ST-100
with Lakeshore controller. The power of the incident laser beam was
∼5 mW. The HV polarization was used to measure the transverse

PR
modes (TA). The advantage of the symmetric scattering geometry is in
the compensation of the refractive index that provides an estimate of
the sound velocity VTA from the Brillouin peak frequency νTA without
knowledge of the materials refractive index [42], VTA = νTA*λlaser/√2
Fig. 3. BLS spectra of studied PolyILs measured at T ∼ Tg. Spectrum for PolyEtVIm Br (this equation is valid for symmetric scattering at 90°). The spec-
is presented at 297 K due to the very poor signal at Tg. In this case, several spectra
tra (Fig. 3) show clear TA mode, that allows estimate G∞ = VTA2*ρ,
below Tg were measured and peak position was extrapolated to Tg. The solid lines rep-
resent fits to the damped harmonic oscillator model. where ρ is the density of the material reported in Ref. [17].

ED
Analysis of the Eσ vs. shear modulus (Fig. 4A) shows no corre-
lation: PolyIL with Br mobile ion has higher shear modulus than the
ductivity in superionic glasses [38]. It predicts that the energy barrier same PolyIL with TFSI mobile ion, but has lower Eσ. However, it is
for conductivity can be presented as: important to realize that elastic contribution to Eσ depends also on the
size of the mobile ion (Eqs. (3) and (4)). We estimated expected elas-
tic contribution to the activation energy, assuming α = 1 for the shov-
(2)
CT
ing model (eq. (4)), and assuming jump length λ∼1 Å and RD = 0 in the
Anderson-Stuart model (eq. (3)). Comparison of the experimental Eσ
with estimated elastic contribution (Fig. 4B) shows good agreement
Here the first term presents the electrostatic interaction of mobile of both models for the case of TFSI ion, while they both significantly
ion with charge q1 and radius R1 with the matrix ion with charge q2 underestimate energy barrier for Br ion. This is consistent with sig-
and radius R2, ε is the dielectric constant, and β is the ‘Madelung’ nificantly larger electrostatic contribution to Eσ in the case of smaller
constant, which indicates charge neutralization between the ion and its Br anion than in the case of larger TFSI- anion (Fig. 2). Unfortu-
RE

surroundings. The second term presents the elastic force contribution, nately experimental data cannot differentiate whether Anderson-Stu-
where G is the shear modulus, and RD is an effective radius of unopen art or shoving model describes better the elastic force contribution.
‘doorway’ for mobile ion to move from one site to another. This model Both models have enough adjustable parameters, and the difference in
assumes that elastic contribution to the energy barrier can be approx- ion size is not large enough to differentiate between ∼ Rion2 or ∼ Rion3
imated by an opening of a ‘doorway’ with the size of the mobile ions {{}}dependence.
[38]. Later the model was modified by McElfresh & How it by includ- To describe the ion transport above Tg we need to include two
R

ing ion jump length (λ) as a more appropriate parameter [39]: possible mechanisms of conductivity: (i) the liquid-like mechanism
with ion diffusion being controlled by polymer segmental mobility,
i.e. τi ∼ τα; and (ii) the quasi-solid-like mechanism with ions jumping
CO

(3)

A similar model was recently proposed in Ref. [9], where elas-


tic contribution was described following the shoving model [40]. This
model assumes that the energy barrier is controlled by fluctuations
needed to create a volume increase, Eel ≈ G∞V, where G∞ is the
UN

high-frequency shear modulus and V is a volume comparable to the


volume of the moving molecule. As a result, the activation energy
controlling ionic conductivity can be presented as [9]:

(4)
Fig. 4. The measured energy barrier for conductivity below Tg vs shear modulus at Tg
(A) and vs elastic contribution estimated using eq. (3) (open squares) with λ∼1 Å and
Here α is a constant, α < 1, that accounts for a possible decrease in the RD = 0, and eq. (4) with α = 1 (open triangles) (B). The line in (B) presents Eσ = Elastic
elastic energy barrier due to frustration in polymer chain packing. Contribution. Strong difference for the Br anion is related to a significant electrostatic
contribution to Eσ for this ion.
4 Electrochimica Acta xxx (2018) xxx-xxx

over energy barriers in a slowly rearranging environment. In that case Here V is the volume of the system, sgn(qi) and vi(t) are the sign and
conductivity can be expressed as: velocity of an ion i. Thus conductivity is defined by velocity-veloc-
ity correlations of all ions, while ion diffusion is defined by self-cor-
relation function. In that respect Nernst-Einstein equation (eq. (1)) is
valid only for dilute solutions where correlations between different
ions is negligible and only self-correlation (i = j) contributes to con-

F
ductivity. In concentrated ionic systems ion-ion correlations are not
negligible and might strongly affect ion conductivity [16,47,48]. It is

OO
(5) usually characterized by the so-called inverse Haven ratio H−1 = Dσ/D
[44,49], where Dσ = σkT/(n0q2) is the averaged charge diffusion esti-
mated from DC conductivity assuming that all mobile ions (n0) con-
tribute. The inverse Haven ratio was analyzed for many ionic liquids
and is often called “ionicity”. It is always below 1 suggesting that not
all ions contribute to the conductivity [16,47,48]. In most cases this ef-
fect is ascribed to diffusion of ion pairs that does not contribute to con-
Here the first term in square brackets presents Vogel-Fulcher-Tamman ductivity. However, recent simulations [45] suggested that anti-corre-

PR
(VFT) temperature dependence of the segmental dynamics, while the lated motion of ions with the same charge also play an important role
second term presents over-barrier relaxation with constants C1(Rion) in reducing conductivity in ionic liquids.
and C2(Rion) that depend on the mobile ion size (eqs. (3) and (4)); n(T) In that respect analysis of the inverse Haven ratio in PolyILs [9,50]
is an ‘effective’ concentration of ions contributing to conductivity and reveals an interesting puzzle (Fig. 5): Although motion of ion pairs
reflects temperature dependent ion-ion correlations (will be discussed can be completely excluded in this case, the inverse Haven ratio is sig-
in the next section). We want to emphasize that shear modulus G∞(T) nificantly below 1, Dσ/D ∼0.1–0.2. This result clearly indicates sig-
has significant temperature dependence above Tg [43], leading to a nificant anti-correlated motion of mobile ions in PolyILs, and authors

ED
VFT-like behavior of conductivity even in the case of strongly decou- of [9] suggested a kind of backflow motions of mobile ions that does
pled ion transport. Thus VFT-like behavior of σ(T) itself is not a sign not contribute to the charge diffusion despite ion diffusion. Similar
of ion transport being coupled to segmental dynamics, as often mis- idea was proposed to explain the low values of the inverse Haven ra-
interpreted in literature. However, at T < Tg the structure is frozen and tio for proton transport in oxides [51]. It is interesting that many su-
the shear modulus has weak temperature variations, leading to an Ar- perionic crystals and glasses exhibit inverse Haven ratio larger than
rhenius-like temperature dependence of σ(T) (Fig. 1). eq. (5) also ex- 1 [44,46,52]. In that case mobile ions correlation is positive lead-
plains the observation that decoupling is usually stronger in high-Tg ing to a chain-like motion that strongly enhances ionic conductivity
CT
polymers than in low-Tg polymers: while segmental dynamics (first [44,46,52].
term) slows down drastically with increase in Tg, the probability of the The presented analysis (Fig. 5) reveals that ionic conductivity in
over-barrier jump (second term) increases significantly with increase polymer electrolytes suppressed by ∼10 times relative to that in su-
in temperature. perionic ceramics due to negative mobile ions correlations. Over-
As was discussed in the previous section, it is not feasible for dry coming this problem is critical to achieve high ionic conductivity in
polymer electrolytes to achieve required ion jump rate τi∼10−10 s at polymer electrolytes. However, it is not obvious how can one design
RE

ambient conditions based on the liquid-like mechanism of conduc- PolyILs with positive mobile ion correlations. It is possible that cre-
tivity alone. To achieve this rate with the over-barrier jumps ating local channels with high mobile ion concentration will bring
(quasi-solid-like mechanism) at room temperature requires the energy these systems closer to superionic ceramics. Enhancement of ionic
barrier to be less than ∼25 kJ/mol (assuming attempt time τ0∼10−14 s). conductivity in highly ordered crystalline PEO might be partially
For PolyILs with large mobile ions, such as TFSI, where energy bar-
rier is dominated by elastic force it would require αG∞(T) ∼ 0.3 GPa
R

(eq. (4)). This might be possible at T > Tg where G∞(T) softens signif-
icantly. At the same time, electrostatic interactions dominate the en-
ergy barrier for small ions such as Li+ and Na+, and can be reduced
CO

by an increase of the dielectric constant. According to the present es-


timates (eq. (4) and Fig. 2), an increase of the dielectric constant to
ε > 40 might reduce the activation energy barrier for Li ion to the re-
quired value. Even smaller ε might work for Na ion.

4. Role of ion-ion correlations


UN

An important point often overlooked in studies of ionic conductiv-


ity in polymers is that conductivity is defined by a charge diffusion
and not directly by the ion diffusion [44–46]. The ion conductivity is
defined by the velocity-velocity correlation of ions [44–46]:

(6)
Fig. 5. The inverse Haven ratio for several PolyILs calculated using diffusion measured
by NMR (open symbols) and estimated from the conductivity relaxation time [50]. Data
from Refs. [9,50].
Electrochimica Acta xxx (2018) xxx-xxx 5

caused by positive ion-ion correlations, where Li ions moves prefer- Acknowledgments


ably inside the PEO chain helix [23–25] and ‘backflow’ effects might
be excluded. Another important point is the extremely high concentra- Authors thank K. D. Kreuer and C. A. Angell for helpful discus-
tion of mobile ions in superionic ceramics, much higher than what can sions and suggestions. This work was supported by the U.S. Depart-
be achieved in polymers. However, one might be able to create local ment of Energy, Office of Science, Basic Energy Sciences, Materials
Science and Engineering Division.

F
concentration of mobile ions in ionic channels that will be comparable
to that in superionic ceramics. Another option might be an increase in
the dielectric constant of the polymer matrix. This should decrease Tg References

OO
[17], significantly reduce the energy barrier for decoupled ion motions
(eqs. (3) and (4)), especially for small ions such as Li+ and Na+, and [1] I. Osada, H. de Vries, B. Scrosati, S. Passerini, Ionic-liquid-based polymer elec-
trolytes for battery applications, Angew. Chem. Int. Ed. 55 (2016) 500.
might additionally decrease the degree of mobile ions correlations.
[2] A. Manthiram, X. Yu, S. Wang, Lithium battery chemistries enabled by
solid-state electrolytes, Nat. Rev. Mater. 2 (2017) 16103.
[3] J. Yi, S. Guo, P. He, H. Zhou, Status and prospects of polymer electrolytes for
5. Conclusions solid-state Li–O2 (air) batteries, Energy Environ. Sci. 10 (2017) 860.
[4] J.C. Bachman, S. Muy, A. Grimaud, H.H. Chang, N. Pour, S.F. Lux, O. Paschos,
F. Maglia, S. Lupart, P. Lamp, L. Giordano, Y. Shao-Horn, Inorganic solid-state
Presented analysis demonstrates importance of two mechanisms of

PR
electrolytes for lithium batteries: mechanisms and properties governing ion con-
ionic conductivity in polymer electrolytes, a liquid-like when ion mo- duction, Chem. Rev. 116 (2016) 140.
tion is coupled to segmental dynamics, and a quasi-solid-like when [5] A. Eftekhari, Polymerized Ionic Liquids, RSC Publishing, Cambridge, 2018.
ion can jump over energy barriers in a frozen or slow moving en- [6] J.Y. Yuan, D. Mecerreyes, M. Antonietti, Prog. Polym. Sci. 38 (2013) 1009.
vironment. However, to achieve the required by many applications [7] J. Daintith, A Dictionary of Chemistry, 6 ed., Oxford University Press, Oxford,
2008.
conductivity σ∼10−3S/cm, the first mechanism requires extremely fast [8] F. Kremer, A. Schönhals, Broadband Dielectric Spectroscopy, Springer-Verlag
segmental dynamics that is not feasible in dry polymer electrolytes. Berlin Heidelberg, Berlin, 2003.
Thus, employment of the second mechanism is critical. Energy barrier [9] E.W. Stacy, C.P. Gainaru, M. Gobet, Z. Wojnarowska, V. Bocharova, S.G.

ED
for ion transport in this case depends on electrostatic interactions that Greenbaum, A.P. Sokolov, Fundamental limitations of ionic conductivity in
polymerized ionic liquids, Macromolecules (2018).
dominate for small ions and on elastic force contribution that domi- [10] P.V. Wright, MRS Bull. 27 (2002) 597.
nates for larger ions (eqs. (3) and (4)). Increase in the dielectric con- [11] M.A. Ratner, D.F. Shriver, Ion transport in solvent-free polymers, Chem. Rev.
stant of the polymer matrix should strongly reduce electrostatic con- 88 (1) (1988) 109–124.
tribution to the energy barrier and might significantly enhance trans- [12] Y. Wang, F. Fan, A.L. Agapov, T. Saito, J. Yang, X. Yu, K. Hong, J. Mays, A.P.
port of small ions, such as Li and Na. Presented here studies revealed Sokolov, Examination of the fundamental relation between ionic transport and
segmental relaxation in polymer electrolytes, Polymer 55 (16) (2014)
a good agreement between the predictions for elastic force contribu-
CT
4067–4076.
tion and the activation energy controlling ion conductivity at T < Tg. [13] J.S. Hansen, A. Kisliuk, A.P. Sokolov, C. Gainaru, Identification of structural re-
Thus reducing shear modulus of glassy polymer matrix will enhance laxation in the dielectric response of water, Phys. Rev. Lett. 116 (23) (2016),
ion transport, especially in the case of large ions, such as TFSI. An- 237601.
[14] T. Iwashita, B. Wu, W.-R. Chen, S. Tsutsui, A.Q.R. Baron, T. Egami, Seeing
other important point we want to emphasize is that the ion conduc- real-space dynamics of liquid water through inelastic x-ray scattering, Sci. Adv.
tivity is defined by a cumulative charge diffusion, not by individual 3 (12) (2017).
ion drift. Ion-ion correlations are not negligible in highly concentrated [15] U.H. Choi, A. Mittal, T.L. Price, M. Lee, H.W. Gibson, J. Runt, R.H. Colby,
RE

ionic systems, including polymer electrolytes, and unfortunately re- Molecular volume effects on the dynamics of polymerized ionic liquids and their
monomers, Electrochim. Acta 175 (2015) 55–61.
duce by almost 10 times ionic conductivity in PolyILs. The micro-
[16] U.H. Choi, Y. Ye, D. Salas de la Cruz, W. Liu, K.I. Winey, Y.A. Elabd, J. Runt,
scopic mechanism of mobile ion correlations in PolyILs is not obvi- R.H. Colby, Dielectric and viscoelastic responses of imidazolium-based
ous, and simulations might be critical in unraveling their details. Un- ionomers with different counterions and side chain lengths, Macromolecules 47
derstanding this mechanism might be critical for design of polymer (2) (2014) 777–790.
electrolytes with positive mobile ion correlations that should enhance [17] V. Bocharova, Z. Wojnarowska, P.-F. Cao, Y. Fu, R. Kumar, B. Li, V.N.
Novikov, S. Zhao, A. Kisliuk, T. Saito, J.W. Mays, B.G. Sumpter, A.P. Sokolov,
R

ion conductivity at the same ion diffusivity. Influence of chain rigidity and dielectric constant on the glass transition tempera-
ture in polymerized ionic liquids, J. Phys. Chem. B 121 (51) (2017)
11511–11519.
Notice [18] A. Manuel Stephan, Review on gel polymer electrolytes for lithium batteries,
CO

Eur. Polym. J. 42 (1) (2006) 21–42.


[19] L. Long, S. Wang, M. Xiao, Y. Meng, Polymer electrolytes for lithium polymer
This manuscript has been authored by UT-Battelle, LLC under batteries, J. Mater. Chem. 4 (26) (2016) 10038–10069.
Contract No. DE-AC05-00OR22725 with the U.S. Department of En- [20] J.B. Boyce, B.A. Huberman, Superionic conductors: transitions, structures, dy-
ergy. The United States Government retains and the publisher, by ac- namics, Phys. Rep. 51 (4) (1979) 189–265.
cepting the article for publication, acknowledges that the United States [21] H. Stephen, Superionics: crystal structures and conduction processes, Rep. Prog.
Phys. 67 (7) (2004) 1233.
Government retains a non-exclusive, paid-up, irrevocable, world-wide [22] C.A. Angell, Mobile ions in amorphous solids, Annu. Rev. Phys. Chem. 43 (1)
license to publish or reproduce the published form of this manuscript,
UN

(1992) 693–717.
or allow others to do so, for United States Government purposes. The [23] Z. Gadjourova, Y.G. Andreev, D.P. Tunstall, P.G. Bruce, Ionic conductivity in
Department of Energy will provide public access to these results of crystalline polymer electrolytes, Nature 412 (2001) 520.
[24] Z. Stoeva, I. Martin-Litas, E. Staunton, Y.G. Andreev, P.G. Bruce, Ionic conduc-
federally sponsored research in accordance with the DOE Public Ac-
tivity in the crystalline polymer electrolytes PEO6:LiXF6, X = P, as, Sb, J. Am.
cess Plan (http://energy.gov/downloads/doe-public-access-plan). Chem. Soc. 125 (15) (2003) 4619–4626.
[25] D. Golodnitsky, E. Livshits, Y. Rosenberg, E. Peled, S.H. Chung, Y. Wang, S.
Bajue, S.G. Greenbaum, A new approach to the understanding of ion transport in
semicrystalline polymer electrolytes, J. Electroanal. Chem. 491 (1) (2000)
203–210.
6 Electrochimica Acta xxx (2018) xxx-xxx

[26] A.M. Christie, S.J. Lilley, E. Staunton, Y.G. Andreev, P.G. Bruce, Increasing the [39] D.K. McElfresh, D.G. Howitt, A structure based model for diffusion in glass and
conductivity of crystalline polymer electrolytes, Nature 433 (2005) 50. the determination of diffusion constants in silica, J. Non-Cryst. Solids 124 (2)
[27] C. Zhang, E. Staunton, Y.G. Andreev, P.G. Bruce, Raising the conductivity of (1990) 174–180.
crystalline polymer electrolytes by aliovalent doping, J. Am. Chem. Soc. 127 [40] J.C. Dyre, Colloquium: the glass transition and elastic models of glass-forming
(51) (2005) 18305–18308. liquids, Rev. Mod. Phys. 78 (3) (2006) 953–972.
[28] D. Golodnitsky, E. Peled, Stretching-induced conductivity enhancement of LiI [41] M.L.F. Nascimento, S. Watanabe, Test of the Anderson–Stuart model and corre-
(PEO)-polymer electrolyte, Electrochim. Acta 45 (8) (2000) 1431–1436. lation between free volume and the ‘universal’ conductivity in potassium silicate

F
[29] D. Golodnitsky, E. Livshits, Y. Rosenberg, I. Lapides, E. Peled, Stretching-in- glasses, Mater. Chem. Phys. 105 (2) (2007) 308–314.
duced changes in ion–polymer interactions in semicrystalline LiI–P(EO)n poly- [42] K.K. Jan, E. Jan, B. Justin, J. Rafael, A new Brillouin scattering technique for
mer electrolytes, Solid State Ionics 147 (3) (2002) 265–273. the investigation of acoustic and opto-acoustic properties: application to poly-

OO
[30] E. Livshits, R. Kovarsky, N. Lavie, Y. Hayashi, D. Golodnitsky, E. Peled, New mers, J. Phys. Appl. Phys. 31 (15) (1998) 1913.
insights into structural and electrochemical properties of anisotropic polymer [43] B. Xu, G.B. McKenna, Evaluation of the Dyre shoving model using dynamic
electrolytes, Electrochim. Acta 50 (19) (2005) 3805–3814. data near the glass temperature, J. Chem. Phys. 134 (12) (2011), 124902.
[31] J.R. Sangoro, C. Iacob, A.L. Agapov, Y. Wang, S. Berdzinski, H. Rexhausen, V. [44] G.E. Murch, The haven ratio in fast ionic conductors, Solid State Ionics 7 (3)
Strehmel, C. Friedrich, A.P. Sokolov, F. Kremer, Decoupling of ionic conductiv- (1982) 177–198.
ity from structural dynamics in polymerized ionic liquids, Soft Matter 10 (20) [45] H.K. Kashyap, H.V.R. Annapureddy, F.O. Raineri, C.J. Margulis, How is charge
(2014) 3536–3540. transport different in ionic liquids and electrolyte solutions?, J. Phys. Chem. B
[32] C.T. Imrie, M.D. Ingram, G.S. McHattie, Ion transport in glassy polymer elec- 115 (45) (2011) 13212–13221.
trolytes, J. Phys. Chem. B 103 (20) (1999) 4132–4138. [46] A. Marcolongo, N. Marzari, Ionic correlations and failure of Nernst-Einstein re-

PR
[33] H. Sasabe, S. Saito, Relationship between ionic mobility and segmental mobility lation in solid-state electrolytes, Phys. Rev. Mater. 1 (2) (2017), 025402.
in polymers in the liquid state, Polym. J. 3 (1972) 624. [47] J.R. Sangoro, F. Kremer, Charge transport and glassy dynamics in ionic liquids,
[34] Y. Wang, A.L. Agapov, F. Fan, K. Hong, X. Yu, J. Mays, A.P. Sokolov, Decou- Acc. Chem. Res. 45 (4) (2012) 525–532.
pling of ionic transport from segmental relaxation in polymer electrolytes, Phys. [48] Y. Wang, C.-N. Sun, F. Fan, J.R. Sangoro, M.B. Berman, S.G. Greenbaum, T.A.
Rev. Lett. 108 (8) (2012), 088303. Zawodzinski, A.P. Sokolov, Examination of methods to determine free-ion dif-
[35] Z. Wojnarowska, K.J. Paluch, E. Shoifet, C. Schick, L. Tajber, J. Knapik, P. fusivity and number density from analysis of electrode polarization, Phys. Rev.
Wlodarczyk, K. Grzybowska, S. Hensel-Bielowka, S.P. Verevkin, M. Paluch, 87 (4) (2013), 042308.
Molecular origin of enhanced proton conductivity in anhydrous ionic systems, J. [49] C.D. Jeppe, M. Philipp, R. Bernhard, L.S. David, Fundamental questions relating
Am. Chem. Soc. 137 (3) (2015) 1157–1164. to ion conduction in disordered solids, Rep. Prog. Phys. 72 (4) (2009), 046501.

ED
[36] F. Fan, W. Wang, A.P. Holt, H. Feng, D. Uhrig, X. Lu, T. Hong, Y. Wang, N.-G. [50] C. Gainaru, E.W. Stacy, V. Bocharova, M. Gobet, A.P. Holt, T. Saito, S. Green-
Kang, J. Mays, A.P. Sokolov, Effect of molecular weight on the ion transport baum, A.P. Sokolov, Mechanism of conductivity relaxation in liquid and poly-
mechanism in polymerized ionic liquids, Macromolecules 49 (12) (2016) meric electrolytes: direct link between conductivity and diffusivity, J. Phys.
4557–4570. Chem. B 120 (42) (2016) 11074–11083.
[37] Z. Wojnarowska, H. Feng, Y. Fu, S. Cheng, B. Carroll, R. Kumar, V.N. [51] M. Spaeth, K.D. Kreuer, T. Dippel, J. Maier, Proton transport phenomena in
Novikov, A.M. Kisliuk, T. Saito, N.-G. Kang, J.W. Mays, A.P. Sokolov, V. pure alkaline metal hydroxides, Solid State Ionics 97 (1) (1997) 291–297.
Bocharova, Effect of chain rigidity on the decoupling of ion motion from seg- [52] N. Kuwata, X. Lu, T. Miyazaki, Y. Iwai, T. Tanabe, J. Kawamura, Lithium dif-
mental relaxation in polymerized ionic liquids: ambient and elevated pressure fusion coefficient in amorphous lithium phosphate thin films measured by sec-
studies, Macromolecules 50 (17) (2017) 6710–6721.
CT
ondary ion mass spectroscopy with isotope exchange methods, Solid State Ionics
[38] O.L. Anderson, D.A. Stuart, Calculation of activation energy of ionic conductiv- 294 (2016) 59–66.
ity in silica glasses by classical methods, J. Am. Ceram. Soc. 37 (12) (1954)
573–580.
R RE
CO
UN

You might also like