Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

E784 Journal of The Electrochemical Society, 165 (14) E784-E792 (2018)

0013-4651/2018/165(14)/E784/9/$37.00 © The Electrochemical Society

Ion-Pair Conductivity Theory V: Critical Ion Size and Range of


Ion-Pair Existence
Andrei Dukhin,1,z Sean Parlia, 2
and Ponisseril Somasundaran2
1 Dispersion Technology Inc., Bedford Hills, New York 10507, USA
2 Columbia University, New York, New York 10027, USA

There are two mechanisms by which ion size affects conductivity of liquids: hydrodynamic resistance, and electrostatic attraction
between cation and anion that leads to the formation of ion-pairs. Increasing ion size reduces conductivity due to growing hydrody-
namic resistance. On the other hand, increasing ion size reduces the probability of ion-pairs formation, which leads to higher ionic
strength and higher conductivity. The opposing nature of these two mechanisms indicates the existence of a “critical ion size” that
corresponds to the maximum conductivity. We apply a theory that was published recently in this Journal with a wide variety of
supporting experimental data for determining value of this “critical ion size”. This theory predicts a very peculiar phenomenon: ions
with sizes that are narrowly distributed around 1 nm (±0.5 nm) must dominate conductivity. This prediction is valid for solutions
with ion-pairs, which according to the theory could occur when relative permittivity is below 10 if ionic strength of the liquids is low
(<10−6 mol/l). This threshold shifts toward higher relative permittivities with increasing ionic strength.
© 2018 The Electrochemical Society. [DOI: 10.1149/2.0821814jes]

Manuscript submitted September 7, 2018; revised manuscript received October 8, 2018. Published November 1, 2018.

When we consider ionization in non-polar media, there are two is not true. The range of liquids that can support ion-pairs is quite
mechanisms by which ion size itself impacts the overall conductivity limited, and will be specified here.
of the system: In order to examine both critical ion size and range of ion-pair exis-
Firstly, ion size determines the hydrodynamic resistance that an ion tence the conductivity theory presented in the paper3 will be used. This
experiences whilst moving relative to the surrounding liquid, when un- theory has been well-tested experimentally, and accordingly verified,
der the influence of an electric field. This hydrodynamic resistance is for a variety of amphiphilic substances that enhance the conductivity
a function of the interfacial area between the ion and the surrounding of non-polar liquids.4–6 These experimental tests confirmed unam-
liquid. Hence, larger ions have more interfacial area and thus larger biguously the validity of this theory across the full composition range
amounts of hydrodynamic resistance, which limits ion mobility and of such liquid mixtures. A short overview of these experiments is
therefore decreases conductivity (as conductivity is a function of the presented below.
ion mobility). Accordingly, the smaller the ion, the higher the con-
ductivity because of the hydrodynamic resistance.
Conductivity Theory
Secondly, ion size controls the distance of closest ion approach in
ion-pairs1 and, consequently, the intensity of cation-anion attraction As we mentioned above, there is a theory3 that describes conduc-
in a potential ion-pair. Smaller ions can approach each other at closer tivity of binary liquid mixtures within the full composition range, from
distances, and therefore the attractive force between them is stronger pure non-polar liquid to pure amphiphile. Here is the final result of
when compared with larger ions, which have distances of closest ion that theory, with equation for conductivity K in units [S/m]:
approach which are farther apart. The stronger the attractive forces
e2
between oppositely charged ions, the more likely they are to combine K = exp
and form an ion-pair. Consequently, the balance between ions and 4π2 η (ϕ)de 3 dh
ion-pairs shifts toward ion-pairs as ion size decreases. As a result, the 
concentration of free ions decreases. An ion-pair is neutrally charged (−λ/de ) ( 1 + 16/3πN A de 3 ϕC A exp (λ/de ) − 1) [1]
entity, and accordingly the ions contained within an ion-pair do not
contribute to the conductivity of a system (i.e. they are not “free ions”). where λ is Bjerrum length which depends on relative permittivity ε,
Therefore, this mechanism leads to a reduction of the conductivity that in turn depends on the volume fraction of amphiphilic liquid ϕ:
with decreasing ion size, which is opposite to the first mechanism e2
pertaining to hydrodynamic resistance. λ= [2]
4πε0 ε (ϕ) kT
As a result of these opposing trends, the conductivity of a given
non-polar system should have a maximum that is a function of ion Other parameters in these equations are:, e is electron charge, ε0 is
size. The value of this “critical ion size” (dcr ) which determines the vacuum permittivity, NA is Avogadro’s number, T is absolute tem-
position (and magnitude) of this conductivity maximum is obviously perature, k is Bolztmann’s constant, η is dynamic viscosity (which
very important. is also dependent on the concentration of amphiphilic liquid ϕ), de
There has been some discussion of the role ion size plays with is distance of closest ion approach in the ion-pair, dh is the effective
regards to non-polar conductivity in the literature, for instance in the hydrodynamic ion size, and CA is the concentration of ions in the pure
extensive review by Prieve et al.2 However, only a limited amount amphiphilic liquid.
of the published data is usually discussed, and the critical ion size There are two parameters that are linked to the geometry of the
that was mentioned above is ignored or omitted entirely. Accordingly, ions: hydrodynamic ion size dh and distance of the closest ion ap-
this paper is dedicated to the theoretical prediction of the value of proach in the ion-pair de , which determines intensity of cation-anion
this critical ion size and the range of concentrations (i.e. amphiphile electrostatic interaction in the pair.
concentrations in nonpolar media) where it is important. Hydrodynamic ion size takes into account potential difference
It should be noted that the notion of critical ion size is relevant only between cation and anion hydrodynamic ion radiuses a+ and a− ,
for those liquids where ion-pairs could potentially exist. This issue respectively, according to the following Equation suggested in the
is also omitted from the relevant literature, creating the impression paper:2
that ion-pairs can exist in any liquid. However, a detailed theoretical 4a + a −
analysis of this subject (presented in this paper) indicates that this dh = [3]
a+ + a−
The relationship between the “distance of the closest approach in
z
E-mail: adukhin@dispersion.com the ion-pair” to ion size is more complicated, because it is known

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 165 (14) E784-E792 (2018) E785

(via monography by K. Izutsu7 ) that the solvating layers of a cation range from pure non-polar to pure amphiphile. There have been 3 non-
and anion can overlap to certain degree when they build an ion-pair. polar liquids tested: toluene, hexane and heptane. The amphiphilic liq-
We follow here the approach suggested, and successfully employed, uids used in these tests includes 8 different alcohols and 4 non-ionic
by Onsager and Fuoss,8,9 who assumed that hydrodynamic ion size surfactants.
equals to the distance of the closest approach: This selection of amphiphilic liquids allowed for experimental
testing of the roles that permittivity and viscosity, two important pa-
dh = de = d [4] rameters in nonpolar electrochemistry, have in affecting the conduc-
This assumption allows us to introduce an effective ion size d as a tivity of a nonpolar liquid mixture. Alcohols have low viscosity very
single geometric property of the ions. close to the viscosity of the studied non-polar liquids. That is why
A detailed derivation of this theory, and its relationship with the the conductivity of alcohol/non-polar mixtures depends mostly on the
Onsager-Fuoss theory,7,8 can be found in the review.10 Its main dis- composition and permittivity of the mixture. Changes in viscosity as
tinction from the Onsager-Fuoss theory is application of the Aurbach a function of mixture composition are practically unimportant.
equation for ionic strength J11 On the other hand, liquid non-ionic surfactants are very viscous.
√ The conductivity of their mixtures with low viscous non-polar liq-
1 + 4M F Ci − 1 uids depends strongly on the variation of viscosity with the mixture
J= [5] composition.
2M F
Detailed descriptions of these experiments have already been pub-
where MF is the Fuoss constant for equilibrium balance between free lished in the papers.3–6
ions and ion-pairs,9 and Ci is total concentration of ions in particular
liquid mixture. The role of permittivity: mixtures of alcohols with non-polar
The final relationship between nonpolar conductivity and effective liquid.—The widest range of conductivity variation that was measured
ion size d is expressed below in the following equation: in these experiments was achieved using methanol as amphiphilic liq-
 uid. Figure 1 presents the conductivity of methanol-toluene mixtures
e2
K = exp (−λ/d) ( 1+16/3πN A d 3 ϕC A exp (λ/d)−1) within the full composition range. It is seen that conductivity varies 7
4π2 η (ϕ)d 4 orders of magnitude.
[6] The solid line in the Figure 1 represents the best theoretical fit to
the experimental data (dots) using the model described by Equation 6.
There are 2 parameters that are unknown and not easily measur- Two adjustable parameters were used: ion size and ionic strength of
able: ion size d and concentration of ions -or ionic strength- of the methanol. Ion size was found to be equal to 0.63 nm, and the ionic
pure amphiphilic liquid CA . These two parameters serve as adjustable strength of the methanol that was used in that experiment was found
parameters for fitting the experimental data. to be 0.00029 mol/L.
This parameterization differs significantly form the classical elec- Figure 1 clearly demonstrates that suggested the theory fits the
trochemical approach, where an added substance (electrolyte) is char- experimental accurately across the entire concentration range.
acterized via molar concentration, dissociation constant, etc. This Looking more closely at the graph in Fig. 1, there are 2 distinctively
classical parametrization is not applicable to non-aqueous electro- different ranges within this conductivity curve.
chemistry for several reasons: At high concentrations of methanol the conductivity is effectively
First, there is no consensus on the mechanism by which ions form, a linear function of the methanol content. This is true for methanol
and are maintained, in non-polar liquids. The Appendix presents sev- concentrations above approximately 30% (by volume). In this range
eral models that might be applicable, but it is not clear yet which one, the relative permittivity of the mixture is high enough to prevent the
if any, is dominant. formation of ion-pairs. However, conductivity becomes an exponen-
Second, the ionic strength of non-polar liquids is many orders of tial function of methanol content below approximately 30%. This is
magnitude lower than that of aqueous solutions. The contribution of the sign that ion-pairs are appearing due to the stronger electrostatic
impurities to the total ion concentration becomes important and cannot attraction between cations and anions.
be neglected for liquids with an ionic strength below 10−6 mol/l. At 30% methanol in a methanol-toluene mixture the relative per-
Third, it is inadequate to define amphiphilic substances that cause mittivity is roughly equal to 10, as is shown on Figure 2. We can
ionization in non-polar liquids as simply electrolytes. It is possible that assume that a relative permittivity of 10 roughly separates the con-
their molecules breakup gives rise to ions, but impurities mentioned centration range in which ion-pairs exist (<10) from the range that
above complicate the quantitative description of this effect. is free from ion-pairs (>10). If this is true, then higher molecular
However, amphiphiles perform a second important function: they weight alcohols, which typically have lower relative permittivities,
solvate ions. Amphiphiles by definition have a hydrophilic portion must exhibit the transition from linear to exponential dependence of
(typically a polar head group) and a hydrophobic (i.e. nonpolar) por- conductivity at higher alcohol content.
tion. The attraction between their polar heads and the core electric Subsequent measurements confirmed this prediction. Figure 3
charge of the ion creates a steric stabilization layer around ions. In presents conductivity dependence curves for all 8 alcohols. It is seen
other words, the molecules of added amphiphilic substance perform that the linear range shrinks with decreasing alcohol relative permit-
the same role water molecules do in an aqueous solution: They sol- tivity. It disappears completely for hexanol, heptanol, and octanol,
vate free ions and allow them to remain in solution. The addition of which have relative permittivities very close to 10.
amphiphilic substance to non-polar liquid enhances solvation of ions. Quantitative agreement has been also achieved. Figure 1, as previ-
The first theoretical description of such a “self-solvating” effect by an ously stated, presents the theoretical fit based on Eq. 6 for methanol-
amphiphilic substance was given in the review.10 toluene mixtures. Similar theoretical fits were observed for all other
The theoretical model given in Eq. 6 reflects these differences studied alcohols-toluene mixtures, as shown in paper.3
between aqueous electrochemistry and the electrochemistry of non- Later, similar agreement between theory and experiment was
polar liquids. It has already been proven that this theory is valid for a demonstrated for mixtures of alcohol with 2 other non-polar liquids
wide variety of mixtures. The most convincing of these experimental (hexane and heptane).4
verification tests are described in the next section.
The role of viscosity: mixtures of non-ionic surfactant with non-
polar liquid.—In order to test the influence of viscosity on the con-
Experimental Verification of the Conductivity Theory
ductivity of liquid mixtures we needed amphiphilic liquids of high
Experimental verification was conducted for two-component mix- viscosity. Liquid non-ionic surfactants are good candidates for such
tures of non-polar liquid and amphiphile, across the full composition a test. Figure 4 shows the viscosity of toluene mixtures with SPAN

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E786 Journal of The Electrochemical Society, 165 (14) E784-E792 (2018)

3
10

2
10

1
10
Conductivity (10E-06 S/m)

0
10

-1 Figure 1. Measured conductivity of binary methanol-


10 Experimental toluene mixture with different alcohol content and theo-
Theoretical retical fit achieved with Eq. 6. Symbols are experimental
-2 data, lines are theoretical fit. Published in the paper.2
10

-3
10

-4
10

-5
10
0 10 20 30 40 50 60 70 80 90 100
Pure Toluene Pure Methanol
Weight Fraction (%)

20 and SPAN 80, which are typical non-ionic surfactants. Since the tial function of the relative permittivity and, consequently, surfactant
viscosity varies 3 orders of magnitude, the impact is has on the con- volume concentration.
ductivity of those systems is quite substantial. Figure 5 combines data for all 4 surfactants that we studied. It
However, the relative permittivity of these mixtures is below 10 in is seen that Walden product is indeed an exponential function of the
both cases because the liquid non-ionic surfactants have low relative surfactant content for all four mixtures on a linear scale (i.e. it is
permittivities themselves. Therefore, it is expected that the existence a linear function when plotted on logarithmic scale). This can be
of ion-pairs in these mixtures, and their dependence on the volume considered as qualitative confirmation of the theory.
fraction of surfactant, must exhibit the same exponential relationship Quantitative confirmation was observed for the solutions of all 4
that was seen in the alcohols describe previously. surfactants as well. Figure 6 demonstrates the best theoretical fit for
It appears that the viscosity variation masks this exponential depen- the conductivity of the toluene mixtures with SPAN 80. A similar fit
dence on permittivity. This was confirmed by multiplying conductivity was achieved for all 4 surfactants.
by viscosity, creating the so called “Walden product12 ”. According to There is one peculiar feature of viscosity for these liquid mix-
the Eq. 6, such product (conductivity x viscosity) must be an exponen- tures that has an interesting effect on conductivity dependence. As

40
Methanol
Ethanol
Propanol
Butanol
30 Pentanol
Hexanol
Heptanol
Dielectric Constant

Octanol

Figure 2. Measured relative permittivity of binary


20
alcohol-toluene mixtures with different alcohol con-
tent. Lines illustrate trends, they are not theoretical
fits, paper.2

10

0
0 10 20 30 40 50 60 70 80 90 100
Pure Toluene Pure Alcohol
Weight Fraction of Alcohol (%)

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 165 (14) E784-E792 (2018) E787

-3
10

-4
10

-5
10

-6
Conductivity (S/m)

10

10
-7
Figure 3. Measured conductivity of binary alcohol-
toluene mixtures with different alcohol content.
Methanol Lines illustrate trends, they are not theoretical fit.
10
-8
Ethanol Published in the paper.2
Propanol
-9 Butanol
10
Pentanol
Hexanol
-10
10 Heptanol
Octanol
-11
10
0 10 20 30 40 50 60 70 80 90 100
Pure Toluene Pure Alcohol
Weight Fraction of Alcohol (%)

demonstrated in Fig. 4, viscosity is an exponential function of sur- Discussion and Theoretical Predictions
factant volume concentration. Somehow this exponential function
The extensive experimental verification of this theory described
compensates the exponential dependence of conductivity on relative
above justifies using it for deriving theoretical predictions regarding
permittivity, specifically in the range of low volume fraction of the
ion size and the range in which ion-pairs exist in liquids. According
surfactant (i.e., the range in which ion-pairs are present). The result-
to this theory, there are several parameters that affect the conductivity
ing effect is quite peculiar: these phenomena appear to “cancel each
of non-polar liquids mixed with amphiphilic substances:
other out,” and conductivity becomes a linear function of the surfac-
tant content in this concentration regime. This unexplained linearity
r Viscosity of the mixture
is documented in prior research, and several electrochemical models
have been suggested for explaining it.10 Our tests indicate that this
r Permittivity of the amphiphilic material
r Volume fraction of the amphiphilic material
linearity is simply the result of the added exponential dependence of
r Concentration of ions in the pure amphiphilic liquid
viscosity on the surfactant volume fraction. To date there appears to
be no theory that can explain such a phenomenon.
r Ion size

SPAN 20 & SPAN 80: Viscosity vs Surfactant


Concentraon - Log Scale
10000

1000

Figure 4. Measured viscosity for mixtures of SPAN


100 non-ionic surfactants with toluene, published in the
Viscosity (cP)

paper.5
SPAN 20
SPAN 80
10

1
0 20 40 60 80 100

0.1
Wt Fracon of SPAN (%)

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E788 Journal of The Electrochemical Society, 165 (14) E784-E792 (2018)

K*η vs Surfactant Concentraon


1.00E+00

1.00E-01

1.00E-02

1.00E-03

1.00E-04
K*η

1.00E-05 SPAN 20

1.00E-06 SPAN 80
OFX-5098
1.00E-07
OFX-0400
1.00E-08

1.00E-09

1.00E-10

1.00E-11
0 20 40 60 80 100
Wt. % of Surfactant

Figure 5. Measured conductivity multiplied with measured viscosity for mixtures of SPAN non-ionic surfactants with toluene and for Xiameter surfactants with
toluene, from the papers.4,5 Logarithmic scale.

In order to graphically demonstrate conductivity trends, it is nec- Next, conductivity was calculated (using Eq. 6) versus ion size
essary to specify all of these parameters. at two ionic strengths (of the liquid mixture) in order to determine a
As a first step, theoretical “Walden products” (conductivity mul- value for the “critical ion size”. In these calculations ionic strength of
tiplied by viscosity) were presented graphically. These plots follow the liquid mixture is the product φ∗ CA in Eq. 6.
theoretical predictions, which was alluded to earlier and states that Finally, similar Walden products are graphed versus relative
conductivity multiplied by viscosity has an exponential dependence permittivity, in order to examine the range in which ion-pairs
on volume fraction when ion-pairs are present, and a linear depen- exist.
dence when they are not. This is discussed in more detail in a previ-
ously published paper on the role of viscosity on the conductivity of Critical ion size value.—Figures 7 and 8 demonstrate the depen-
non-polar liquids.12 dences of conductivity multiplied by viscosity (Walden product) on

SPAN 80 - Theory vs Experiment


1

0.1
Conducvity (10-6 S/m)

0.01 Figure 6. Measured data and theoretical models for


conductivity of mixtures of SPAN 80 non-ionic sur-
factant with toluene. Symbols are as follows: ( ) –
Measured Data Measured conductivity; ( ) – Theoretical con-
Theorecal Data
ductivity. Published in the paper.5
0.001

0.0001

0.00001
0 20 40 60 80 100
Wt. % of SPAN 80

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 165 (14) E784-E792 (2018) E789

2E-11

1.5E-11

Es = 5
Es = 10
Figure 7. Conductivity multiplied by viscosity ver-
1E-11 sus ion diameter. Ionic strength of the liquid mixture
is 10−8 mol/l. Non-polar liquid relative permittivity
K*η (C2/m4)

Es = 15
is 2.36. Amphiphilic liquid relative permittivities are
Es = 20
shown in the legend.
Es = 25
Es = 30
5E-12

0
0 1 2 3 4 5
Ion Size d (nm)

the ion size. The ionic strength of the liquid mixture is 10−8 mol/L in However, the range still remains very narrow, from roughly 0.5 nm
Figure 7, and 10−7 mol/L in Figure 8. up to 1.5 nm.
Relative permittivity of the amphiphile is not important at lower This simple analysis predicts that ions with diameters that are
ionic strengths, as all curves overlap. The value of the critical ion di- closely clustered around 1 nm (±0.5 nm) contribute most to the con-
ameter that ensures the highest conductivity is approximately 1.5 nm. ductivity of non-polar liquids. These theoretical predictions can be
This is true for all curves in Figure 7, with only slight variation due to compared with values of ion size published in literature.
relative permittivity. We found 17 papers3–6,15,21–32 that published data on ion size in
Increasing the ionic strength leads to a spreading of the critical ion non-polar liquids with the addition of various amphiphilic substances,
size, making it dependent on the relative permittivity of the additive. both solid and liquid. While no claims are made that these papers

4E-10

3.5E-10

3E-10

2.5E-10

Es = 5
Figure 8. Conductivity multiplied by viscosity
Es = 10
2E-10 versus ion diameter. Ionic strength of the liquid
K*η (C2/m4)

Es = 15 mixture is 10−7 mol/l. Non-polar liquid relative


Es = 20 permittivity is 2.36. Amphiphilic liquid relative
permittivities are shown in the legend.
1.5E-10 Es = 25
Es = 30

1E-10

5E-11

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Ion Size d (nm)

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E790 Journal of The Electrochemical Society, 165 (14) E784-E792 (2018)

14

12

10
Frequency

Figure 9. Frequency of appearance of certain ion


6
diameters in the published paper.

More
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
3.2
3.4
3.6
3.8
4
4.2
4.4
4.6
4.8
5
ion diameter [nm]

are all encompassing of such data, it seems reasonable to consider below approximately 10 in order to support ion-pairs. This empirical
them representative of the findings on this subject. Some papers prediction is verified in the next section using the abovementioned
present numbers for ion radii, which was converted into ion diam- theory.
eters. It is worth noting that the methods applied for collecting this
data have not been considered, and that differences between anion Range of ion-pairs existence.—Figures 10 and 11 depict theoret-
and cation size were ignored. The purpose here is simply to see if ical Walden products plotted versus the relative permittivity of the
agreement exists between theory and published data on an elementary liquid mixture. The liquid mixture ionic strength is 10−6 mol/L in Fig.
level. 10, and 10−2 mol/L in Fig. 11.
Figure 9 presents these values for ion diameters (in nanometers) The graphs show that for low ionic strength liquids (Fig. 10),
organized as a histogram. It is seen that the published data for ion the relative permittivity value of 10 is indeed the threshold that
sizes is also, for the most part, clustered around 1 nm, which agrees determines the range in which ion-pairs exist, i.e. they could exist
with theoretical prediction formulated above. only in liquids with relative permittivity below 10 in low conducting
As stated earlier, the existence of a critical ion size is due to liquids.
opposing mechanisms, which both affect conductivity and are ion Increasing ionic strength shifts this ion-pair threshold to the higher
size dependent: hydrodynamic resistance, and electrostatic cation- relative permittivity values. Fig. 11 indicates that it becomes close to
anion attraction that leads to ion-pairs formation. 30 in the liquids with a high ionic strength of 10−2 mol/L, which
Critical ion size phenomenon occurs, consequently, only in liquids could correspond to even aqueous solutions. However, it is still well
with ion-pairs. The experiment presented above in the The role of vis- below relative permittivity of 80 which would typically characterize
cosity: mixtures of non-ionic surfactant with non-polar liquid section the dielectric range of aqueous solutions. This implies that ion-pair
indicates that the relative permittivity of the liquid mixture must be existence in aqueous solutions is unlikely.

Figure 10. Conductivity multiplied by viscosity versus relative permittivity of the liquid. Ionic strength of the liquid mixture is 10−6 mol/l. Non-polar liquid
relative permittivity is 2.36. Ion size values are shown in the legend.

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 165 (14) E784-E792 (2018) E791

Figure 11. Conductivity multiplied by viscosity versus relative permittivity of the liquid. Ionic strength of the liquid mixture is 10−2 mol/l. Non-polar liquid
relative permittivity is 2.36. Ion size values are shown in the legend.

An additional interesting conclusion can be drawn from these re- are in consensus with previous studies of non-aqueous electrochemistry conducted in the
sults. There are two known mechanisms for describing ion interaction: early 20th century by Bjerrum, Onsager, and Fuoss.1,8,9 They conclude that the balance
Bjerrums’ ion-pairs and Debye-Huckels’ ion-clouds.32 Our analysis between free ions and ion-pairs is a controlling factor of ionic strength in non-polar
here confirms that these two mechanisms do not overlap, which justi- liquids.
Unfortunately, there is no such concurrence on the mechanism that is responsible for
fies their being studied independently. Bjerrums’ ion-pairs dominate
ion formation in non-polar liquids in the first place. Below we present some comments
ion interaction in liquids with low ionic strength and low relative regarding the variety of models on this subject.
permittivity, whereas the Debye-Huckel effect is important for high Early works on non-aqueous electrochemistry stressed the importance of electrostatic
conducting aqueous solutions. interactions between cations and anions due to the low relative permittivity of the non-
polar liquid. This leads to the understanding that ions should have a sufficiently large size
in order to exist as free entities in such liquids. That is why early pioneering works by
Conclusions Fuoss, Krauss, and Onsager were conducted with large picrate molecules.8,9,13,14 They
assumed that the dissociation of such molecules would yield large ions that could withstand
Ion-pairs conductivity theory created in the paper3 and experimen- electrostatic attraction. Effectively, they applied the classical “dissociation model” for
tally verified with a variety of liquid mixtures predicts the existence of explaining ion formation in non-polar liquids with one additional requirement: that ions
a critical ion diameter at which the conductivity of a non-polar liquid must be sufficiently large.
with amphiphilic additive, such as surfactant, alcohol, etc., reaches a However, they neglected to discuss the solvation aspect of ionization. There is practi-
maximum. The value of this critical ion size is approximately 1 nm cally no discussion about the state of the small counter-ions that split from the large picrate
± 0.5 nm. This phenomenon of the “critical ion size” reflects the molecules. The fact that these ions cannot be solvated by the non-polar liquid molecules
balance between hydrodynamic resistance and electrostatic attraction due to the lack of molecular dipole moments (in the non-polar liquid molecules) was
simply ignored.
between cations and anions leading to ion-pair formation. These two
On the other hand, Fuoss and Krauss realized that ion-pairs carry dipole moments
effects contribute to the ionic conductivity of the system with the that could lead to the formation of more complicated structures due to ion-dipole and
opposing trends. dipole-dipole interactions. This approach to non-aqueous electrochemistry, based on the
Ions with sizes in said range dominate the conductivity of liquid dissociation model, is still being applied in relevant peer-reviewed literature, as can be
mixtures that contain ion-pairs, and notion of “critical ion size” is seen in recent (2009) publications.15
valid only for such liquids. There is a threshold value for relative However, this model ignores the existence of the other objects with dipole moments
permittivity of the liquid mixture that marks the range of such liquids in such non-polar solutions: neutral molecules of the added substances. It turns out that
that contain ion-pairs. This threshold value is 10 for the liquids with practically all substances that causes ionization of non-polar liquids are amphiphilic. They
low ionic strength below 10−6 mol/L, as only liquids with a relative must have a sufficiently long carbon chain (hydrophobic) on one side for solubility in
non-polar liquid, and a polar group that is capable of dissociation (hydrophilic) on the
permittivity below 10 can support ion-pairs at low ionic strength. This
other side. This polar group has a dipole moment, which may play an important role in
conclusion is supported by both theory and experiment. ion formation of various structures.
This threshold value increases up to approximately 30 for liquids The earliest paper known to us on the subject of the interactions of dipole moments of
with high ionic strength (e.g. 10−2 mol/L), which corresponds to aque- added molecules in non-polar liquids is by Nelson and Pink in 1951.16 In that paper they
ous solutions. This means that the existence of ion-pairs in aqueous wrote that “. . . in a non-polar solvent, aggregation is clearly the result of a balance between
solutions is unlikely. On the other hand, it also implies that the two the solubilizing power of the hydrocarbon chains and the attractive forces between polar
effects of ion interactions, Bjerrum’s ion-pairs and Debye-Huckels’ parts of the soap [surfactant] molecules. . . ” They also wrote that “. . . Like the coulombic
ion-clouds, do not overlap. force between ions, the attractive force due to the dipoles is exerted indiscriminately on
all other polar molecules depending only on the distance of separation. . . ”
This dipole-dipole interaction leads to the formation of “inverse micelles” in non-
polar liquids, which is supported by many detailed experimental studies. These “inverse
Appendix: Mechanisms of Ion Formation in Non-Polar Liquids
micelles” are neutral entities. However, it is possible that could become charged some-
There are two recently published reviews2,10 on the electrochemistry of non-polar times.
liquids. Both of them present an analysis of the current understanding of the mechanisms According to Morrison,17 a paper by Mattoon and Mathews from 194918 was the
by which ions form and are maintained in non-polar liquids. Both reviews are in agreement first to mention “charged inverse micelles”. They discovered that AOT micelles behave as
regarding the importance of ion-pair formation in non-polar liquids. On this subject they positively charged particles. The mechanism describing this charging of inverse micelles

Downloaded on 2018-12-25 to IP 130.63.180.147 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
E792 Journal of The Electrochemical Society, 165 (14) E784-E792 (2018)

is called the “disproportionation model”. It assumes that two micelles exchange electric 6. S. Parlia, A. S. Dukhin, and P. Somasundaran, “Ion-Pair Conductivity Theory IV:
charge during collision. SPAN Surfactants in Toluene and the Role of Viscosity.”, J. of the Electrochemical
The disproportionation model is most widely used in Colloid Science. This field Society, 165(5), H1 (2018).
became involved into non-polar liquid electrochemistry because amphiphilic substances, 7. K. Izutsu, “Electrochemistry in Nonaqueous Solutions”, Willey-VCH, (2007).
8. L. Onsager, “Report on revision of the conductivity theory”, Trans. Faraday Soc, 23,
such as surfactants, are common.
341 (1927).
However, the disproportionation model is not the only mechanism used to describe 9. R. M. Fuoss, “Ionic association. I. Derivation of constants from conductance data”,
these phenomena. The “Fluctuation model” reflects the potential role of water in the J. Am. Chem. Soc., 79(13), 3301 (1957).
ionization process. The addition of water to a non-polar liquid that contains surfactant (with 10. A. S. Dukhin and S. Parlia, “Ions, ion-pairs and inverse micelles in non-polar media”,
a certain HLB number) leads to the formation of microemulsion droplets. Such droplets Current Opinion in Colloid and Interface Science, 18, 93 (2013).
are similar to inverse micelles to a degree, except with a water core. And similarly to 11. D. Aurbach, “Nonaqueous Electrochemistry”, Marcel Dekker, (1999).
inverse micelles these droplets can exchange electric charge during collisions and become 12. P. Walden, Bull. Acad. Imp. Sci., St. Petersburg, 7, 934 (1913), referenced according
ions. The fluctuation model theory had been derived and discussed more thoroughly in to P. Walden, “Salts, Acids, and Basis; Electrolytes; Stereochemistry”, McGraw-Hill,
New York, (1929).
the papers.19,20 This model can be important even when water is present in very small
13. R. M Fuoss and C. A. Kraus, “Properties of Electrolytic Solutions. III. The Dissoci-
amounts, as was shown in the paper.21 ation constant”, J. Am. Chem. Soc., 55, 1019 (1933).
Lastly, the classical “dissociation model” described earlier can be extended to so- 14. C. A. Kraus and R. M. Fuoss, “Properties of electrolytic solutions. 1. Conductance
lutions that include amphiphilic substances. The polar heads of these molecules can as influenced by the dielectric constant of the solvent medium”, J. Am. Chem. Soc.,
occasionally dissociate like, for instance, what happens with picrate, alcohols, and ionic 55(1), 21 (1933).
surfactants. The nearby amphiphilic molecules, while still neutrally charged because they 15. N. O. Mchedlov-Petrossyan, I. N. Palval, A. V. Lebed, and E. M. Nikiforova, “Asso-
have not dissociated, are attracted to the newly formed ions due to the dipole moment in ciation of the picrate ion with cation of various nature in solvents of medium and low
their polar heads. relative permittivity. An UV/Vis spectroscopic and conductometric study”, Journal
of Molecular Liquids, 145, 158 (2009).
This attraction leads to a buildup of a neutral molecule shell around each of the
16. S. M. Nelson and R. C. Pink, “Solutions of metal soaps in organic solvents. Part III.
charged (dissociated) ions. This, in turn, would enlarge such ions, sterically protecting The aggregation of metal soaps in toluene, isobutyl alcohol, and pyridine”, J. of the
them from re-aggregation. Chem. Soc, 1744 (1952).
In electrochemical terms, the added amphiphilic substance can “solvate” ions in non- 17. M. B. Mathews and R. W. Mattoon, J. Chem. Phys., 17, 496 (1949).
polar liquids, which molecules have no such capacity due to their lack of dipole moments. 18. H. F. Eicke, M. Borkovec, and B. Das-Gupta, “Conductivity of water-in-oil mi-
In contrast to aqueous electrochemistry, the added substances serve as “solvating agents” croemulsions: A quantitative charge fluctuation model”, J. Phys. Chem., 93, 314
and not necessarily as “electrolytes”. This “dissociation model with self-solvation” was (1989).
suggested in the paper.10 The term “self-solvation” is derived from the amphiphilic ion 19. D. G. Hall, “Conductivity of microemulsions: An improved fluctuation model.”
J. Phys. Chem., 94, 429 (1990).
being solvated by similar amphiphilic molecules, e.g. a dissociated ethanol ion would be
20. M. Gacek, D. Bergsman, E. Michor, and J. C. Berg, “Effects of trace water on
solvated by other ethanol molecules. charging of silica particles dispersed in nonpolar medium”, Langmuir, 28, 11633
The structure of an ion that is “solvated with a shell of amphiphilic molecules” is (2012).
identical to the “charged inverse micelle”. In this sense “disproportionation model” and 21. S. Kaufman and C. R. Singleterry, “Micelle formation by sulfonates in a nonpolar
“dissociation model with self-solvation” lead to the same result, due to the same sterically solvent”, Journal of Colloid Science, 10(2), 139 (1955).
stabilized ions. 22. K. Miyoshi, “Comparison of the conductance equations of Fuoss-Onsager, Fuoss-
However, there is one important difference. “Dissociation model with self-solvation” Hsia and Pitts with the data of bis(2,9-dimethyl-1,10-phenanthroline)Cu(1) Perchlo-
does not require the preliminary step of neutral micelle formation. That is why it is rate”, Bulletin of the Chem. Soc. Japan, 46, 426 (1973).
23. N. O. Mchedlov-Petrossyan, I. N. Palval, A. Lebed, and E. M. Nikiforova, “Asso-
applicable in cases when micelles do not form such as, for instance, solutions of alcohols in
ciation of the picrate ion with cations of various nature in solvents of medium and
non-polar liquids.3 It can also explain ionization in a solution with surfactant concentration low relative permittivity. An UV/Vis spectroscopic and conductometric study”, J. of
below the Critical Micelle Concentration (CMC). Molecular Liquids, 145, 158 (2009).
24. H. Sadek and R. Fuoss, “Electrolyte solvent interaction. VI. Tetrabutylammonium
bromide in nitrobenzene-carbon tetrachloride mixtures”, pp. 5005.
ORCID 25. M. F. Hsu, E. R. Dufresne, and D. A. Weitz, “Charge stabilization in nonpolar sol-
vents”, Langmuir, 21(11), 4881 (2005).
Sean Parlia https://orcid.org/0000-0001-5763-0167 26. A. Denat, B. Gosse, and J. P. Gosse, “Electrical conduction of solutions of an ionic
surfactant in hydrocarbons”, J. of Electrostatics, 12, 197 (1982).
27. S. K. Sainis, J. W. Merrill, and E. R. Dufresne, “Electrostatic interactions of colloidal
References particles at vanishing ionic strength”, Langmuir, 24, 13334 (2008).
28. P. M. Bowyer, A. Ledwith, and D. C. Sherrington, “Ion-pair dissociation equilibria for
1. N. Bjerrum, “A new form for the electrolyte dissociation theory”, Proceedings of the hexachloroantimonate salts of stable organic cations in dichloromethane”, J. Chem.
7th International Congress of Applied Chemistry, Section X, p.55-60, London, 1929. Soc., B, 1511 (1971).
2. D. C. Prieve, B. A. Yezer, A. S. Khair, P. J. Sides, and J. W. Schneider, “Formation 29. C. A. Kraus and G. S. Hooper, “The dielectric properties of solutions of elec-
of Charge Carriers in Liquids”, Advances in Colloid and Interface Science, (2016). trolytes in non-polar solvent”, Proc. of the National Academy of Sci., 19(#11), 939
3. A. Dukhin and S. Parlia, “Ion-Pair Conductivity Theory Fitting Measured Data for (1933).
Various Alcohol-Toluene Mixtures across Entire Concentration Range.” Journal of 30. A. P. Abbot and D. J. Schiffrin, “Conductivity of tetra-alkyammonium salts in pol-
The Electrochemical Society, 162(4), H1 (2015). yaromatic solvents”, J. Chem. Soc. Faraday Trans., 86(9), 1453 (1990).
4. S. Parlia, A. Dukhin, and P. Somasundaran, “Ion-Pair Conductivity Theory: Mixtures 31. K. Mpoukouvalas, D. Turp, M. Wagner, K. Mullen, H-J. Butt, and G. Floudas,
of Butanol with Various Non-Polar Liquids and Water.” Journal of The Electrochem- “Dissociation and charge transport in salts of dendronized ions in solvents of low
ical Society, 163(7) H1 (2016). polarity”, J. of Physical Chemistry, 115, 5801 (2011).
5. S. Parlia, A. S. Dukhin, and P. Somasundaran, “Ion-pair conductivity theory: mixtures 32. P. Debye and E. . H¨
of non-polar media with non-ionic surfactant”, J. of the Electrochemical Society,
164(12), E1 (2017).

You might also like