16.08.2013 Views

ARTICLE IN PRESS

ARTICLE IN PRESS

ARTICLE IN PRESS

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

CIS-00944; No of Pages 26<br />

Abstract<br />

Characterization of ion-exchange membrane materials:<br />

Properties vs structure<br />

N.P. Berezina ⁎ , N.A. Kononenko, O.A. Dyomina, N.P. Gnusin<br />

Department of Physical Chemistry, Kuban State University, 149, Stavropolskaya St., Krasnodar, 350040, Russia<br />

This review focuses on the preparation, structure and applications of ion-exchange membranes formed from various materials and exhibiting<br />

various functions (electrodialytic, perfluorinated sulphocation-exchange and novel laboratory-tested membranes). A number of experimental<br />

techniques for measuring electrotransport properties as well as the general procedure for membrane testing are also described. The review<br />

emphasizes the relationships between membrane structures, physical and chemical properties and mechanisms of electrochemical processes that<br />

occur in charged membrane materials. The water content in membranes is considered to be a key factor in the ion and water transfer and in<br />

polarization processes in electromembrane systems. We suggest the theoretical approach, which makes it possible to model and characterize the<br />

electrochemical properties of heterogeneous membranes using several transport-structural parameters. These parameters are extracted from the<br />

experimental dependences of specific electroconductivity and diffusion permeability on concentration.<br />

The review covers the most significant experimental and theoretical research on ion-exchange membranes that have been carried out in the<br />

Membrane Materials Laboratory of the Kuban State University. These results have been discussed at the conferences “Membrane<br />

Electrochemistry”, Krasnodar, Russia for many years and were published mainly in Russian scientific sources.<br />

© 2008 Elsevier B.V. All rights reserved.<br />

Keywords: Ion-exchange membrane; Water content; Electroconductivity; Diffusion permeability; Electroosmotic transport<br />

1. Introduction<br />

Charged synthetic membranes with high conductivity and<br />

selectivity are used as separating films in various electromembrane<br />

devices: electrodialyzers, fuel cells and electrolyzers. Two<br />

characteristics of electromembrane processes are particularly<br />

important: (1) they are ecologically safe and (2) the electrical<br />

energy required for these processes is comparatively low. The<br />

efficiency of electromembrane processes depends on physicochemical<br />

characteristics and electrotransport properties of ionexchange<br />

membranes (Fig. 1) placed in the external electric<br />

field, when the electrolyte concentration and temperature are<br />

changed simultaneously [1–5].<br />

Recently, the substantial advances in synthesis and modification<br />

of ion-exchange polymeric membranes have been made.<br />

Nevertheless, the extensive studies are continued to develop<br />

novel materials and methods for membrane modification and to<br />

⁎ Corresponding author. Tel./fax: +7 861 2199573.<br />

E-mail address: berezina@chem.kubsu.ru (N.P. Berezina).<br />

0001-8686/$ - see front matter © 2008 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.cis.2008.01.002<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

www.elsevier.com/locate/cis<br />

produce more complex membrane structures with a wide range<br />

of functionalities and desired operating behaviors. The problem<br />

of membrane characterization has been broadly discussed in<br />

several reviews on the preparation and application of ion-exchange<br />

membranes [6–8]. Physicochemical properties of ionexchange<br />

membranes were extensively investigated by O.<br />

Kedem, P. Meares, A. Narebska, R. Paterson, C. Gardner, R.<br />

Buvet, B. Auclair and others [9–21]. The membrane characterization<br />

problem continues to attract attention in current research<br />

due to the recent significant expansion of the types of polymeric<br />

matrices [22,23] and development of composite membrane<br />

materials [24–28].<br />

Membrane characterization includes the comprehensive studies<br />

on equilibrium and physical–mechanical properties (exchange<br />

capacity, water content, sorption properties, thickness of<br />

films, thermal and chemical stability); those on transport properties<br />

(electroconductivity, diffusion and electroosmotic permeability,<br />

transport numbers of ions) and structural characteristics<br />

measured by various physical methods (X-ray, spectral and<br />

optical methods that allows for investigating the structure of<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


membranes at different length scales). These properties are<br />

commonly mentioned in the catalogs of commercially available<br />

ion-exchange membranes [29–34]. However, there remain several<br />

outstanding questions: is it possible to find a compromise<br />

between exchange capacity, water content, conducting properties<br />

and thermal–mechanical stability of a material? How to<br />

establish the relationships between the transport properties of<br />

membranes (selectivity, conductivity, diffusion and electroosmotic<br />

permeability)? How to measure the changes in electrotransport<br />

properties of membranes in separation process, when the<br />

membrane behavior is influenced by the electric and concentration<br />

fields? To answer these questions, which are very important<br />

for fabricating of membrane materials and controlling their<br />

properties, the approach to the characterization of membranes<br />

remains to be developed using the summarized data on membrane<br />

properties.<br />

The objectives of this review are as follows: (1) to summarize<br />

the results of experimental studies on physicochemical<br />

characteristics and electrotransport properties of ion-exchange<br />

membranes, (2) to sketch some of the important factors<br />

that contribute to the ion and water transfer mechanisms, and<br />

(3) to suggest a model approach that allows the characterization<br />

of electrodiffusion properties of structurally heterogeneous<br />

membranes.<br />

2. Membrane materials and pre-testing procedures<br />

The commercial ion-exchange membranes were obtained from<br />

different companies (Russia, Japan, USA and China). Laboratory<br />

samples were prepared from aromatic polymers (polysulphone<br />

and polyether-ether-ketone) and from perfluorinated materials<br />

(Table 1).<br />

The pre-testing treatment of polymeric membranes is a vitally<br />

important factor for the understanding of several aspects of<br />

these systems including (1) influence of synthesis conditions<br />

on some physicochemical properties of membrane material,<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

2 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 1. Schematic diagram of an electromembrane system.<br />

(2) influence of operating conditions on membrane properties,<br />

(3) reliability of model parameters, and (4) reproducibility of the<br />

experimental results. Membrane materials used with no additional<br />

pre-testing treatment are unstable and often contaminated<br />

with various (organic) substances.<br />

The pre-testing treatment of membranes is a stepwise procedure.<br />

The heterogeneous polymer membranes are first treated<br />

with ethanol for 6 h to extract the monomer residues and<br />

surfactant inclusions that remain in the material after the synthesis.<br />

Then the membranes are treated with the solutions of<br />

salts, acids or bases [46]. Finally, the membranes are rinsed with<br />

distilled water. Single processing step requires about 48 h.<br />

The structure of unlinked polymeric membranes based on<br />

perfluorinated matrices is very sensitive to pre-testing treatment<br />

protocol [47]. We used different methods to process perfluorinated<br />

membranes. For the “salt-based” method, the rinsing of<br />

membranes with 2 М NaCl solutions was followed by equilibrating<br />

with NaCl solutions of lower concentration (down to<br />

0.05 M) under isothermal conditions at 25 °C. In “thermalbased”<br />

procedure, the membranes were subsequently treated<br />

with boiling solutions of 3% Н 2О 2, Н 2О, 0.5 МНNO 3 or<br />

Н 2SO 4, and, finally, Н 2О (each single step required 3 h). The<br />

samples processed in this fashion were thermally equilibrated<br />

with water, and, finally, with NaCl or НCl solutions of various<br />

concentrations at 25 °C.<br />

3. Experimental section<br />

Several standard techniques have been used to test membranes<br />

[19]. The ion-exchange capacity (IEC) was determined<br />

for H + -form samples by titration of the H + -ions produced in a<br />

course of base neutralization [48]. The water was evaporated at<br />

105 °С, and the water content w was determined as the mass<br />

ratio of water to dry membrane by the gravimetric method [49].<br />

The membrane hydration capacity nm (average molar content of<br />

H 2O molecules/1 mol of functional groups) was calculated from<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

Table 1<br />

Membrane materials<br />

Membrane/produced by Membrane type Fixed groups Application Reference<br />

MK-40 “Schekinazot” (Russia) Composites formed from the cation-exchange resins KU-2 (polystyrene<br />

(PS) matrix cross-linked with divinylbenzene (DVB) and fixed groups),<br />

polyethylene and nylon<br />

MK-41 “Schekinazot” (Russia) Composites formed from the cation-exchange resins KF-1 (PS matrix<br />

cross-linked with DVB and fixed groups), polyethylene and nylon<br />

−<br />

–HPO3 MA-41 “Schekinazot” (Russia) Composites fabricated from the anion-exchange resins AV-17 (PS<br />

cross-linked with DVB and fixed groups), polyethylene and nylon<br />

–N + (CH3)3<br />

MA-40 “Schekinazot” (Russia) Composites on the base of resin EDE-10P produced via condensation<br />

` N _ NH–N<br />

of diamine with epichlorhydrine, polyethylene and nylon<br />

+<br />

Neosepta СМ-1 Tokuyama Soda (Japan) Composites prepared by a former method on the base of PS and DVB and<br />

reinforced with polyvinylchloride; these membranes are interpolymer materials<br />

(CH3) 3 20%<br />

–SO3 −<br />

Neosepta АМ-1, АМ-Х, Tokuyama Soda (Japan) –N + (CH3)3<br />

CR 67-HMR-412 “Ionics” (USA) Composite prepared from vinyl monomer and acrylic fiber –SO 3 −<br />

AR 204-SZRA-412 “Ionics” (USA) –N + 3361-BW Shanghai (China) Composite fabricated from powder of cation-exchange resins (PS cross-linked<br />

with DVB) and polyethylene and reinforced with nylon<br />

(CH3)3<br />

–SO3 −<br />

3362-BW Shanghai (China) Composite from powder of anion-exchange resins and polyethylene and<br />

reinforced with nylon, membranes have blue or green color<br />

–N + (CH3) 3<br />

RALEX MEGA (Czechia) Composites formed from ion-exchange resins, polyethylene and<br />

polyamide as the reinforcing material<br />

–SO3 −<br />

RALEX MEGA (Czechia) –N + (CH3) 3<br />

KESD (Poland) Composites formed from the PS cross-linked with DVB and polyethylene –SO 3 −<br />

AESD (Poland) –N + Nafion 115, 117, 120, 425 Dupon de<br />

Films fabricated from polytetrafluorethylene and perfluorvinylether<br />

(CH3)3<br />

–SO3<br />

Nemoure (USA)<br />

with fixed charges<br />

−<br />

MF-4SK various modifications<br />

Films fabricated from polytetrafluorethylene and perfluorvinylether<br />

–SO3<br />

“Plastpolymer” (Russia)<br />

with fixed charges<br />

−<br />

SPS-1, 2 University of Twente (The Netherlands) Films fabricated from polysulphone matrix with varied degree of sulphurization –SO 3 −<br />

SPEEK University of Twente (The Netherlands) Films fabricated from polyether-ether–ketone matrix with varied<br />

degree of sulphurization<br />

Kaspion, Karpov Institute (Russia) Films fabricated from aromatic polyamides with varied phenylon content –SO3 −<br />

–SO 3 −<br />

–SO 3 −<br />

Electrodialysis separation<br />

processes, water desalination<br />

Bipolar electrodialysis<br />

separation processes<br />

Electrodialysis separation<br />

processes, water desalination<br />

Electrodialysis separation<br />

processes, water desalination<br />

Electrodialysis separation<br />

processes, water desalination<br />

Electrodialysis separation<br />

processes, water desalination<br />

Electrodialysis separation<br />

processes, water desalination<br />

[33]<br />

[33,35]<br />

[33,36]<br />

[33,36]<br />

[29]<br />

[31]<br />

[19]<br />

[36]<br />

Electrodialysis separation<br />

process, water desalination<br />

[34]<br />

Diffusion dialysis [37]<br />

Membrane electrolysis, fuel [8,30,38,39]<br />

cells, medicine, sensors and others<br />

Membrane electrolysis, fuel cells [40–42]<br />

and others<br />

Laboratory-made samples [22,43,44]<br />

Laboratory-made samples<br />

Laboratory-made samples [23,45]<br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

3<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong>


Table 2<br />

Schematic drawings of the experimental set-up<br />

Cell schematics Calculation equation Experimental<br />

conditions<br />

Km ¼ l<br />

RS<br />

l, membrane thickness;<br />

S, area;<br />

R, resistance;<br />

Pm ¼ VlDc<br />

ScDt<br />

V, chamber volume;<br />

c, concentration;<br />

t, time;<br />

V ⁎<br />

W ¼<br />

Sit<br />

tw ¼ WF<br />

18<br />

i, current density;<br />

F, Faraday number;<br />

V⁎, volume of water<br />

transfer<br />

Graphic determination<br />

of limiting current and<br />

other parameters of<br />

current–voltage curve<br />

AC frequency<br />

200 kHz<br />

Reference<br />

[64]<br />

i=0 [19,70]<br />

Δc≠0<br />

Δp=0<br />

i≠0 [19,71]<br />

Δc=0<br />

Δp=0<br />

i≠0 [72]<br />

Δc=0<br />

Δp=0<br />

(1) Membrane; (2) measuring electrodes; (3) polarizing electrodes; (4) impedance<br />

meter; (5) potentiostat; (6) measuring capillaries; (7) ammeter; (8) stirrings;<br />

(9) mercury.<br />

the values of the water content and exchange capacity of the<br />

samples; the accuracy of the determination of nm was of 0.1 mol<br />

H2O/mol SO3 − [50–52].<br />

The porous structure of ion-exchange membranes was<br />

studied with various methods: electron microscopy, atomic<br />

force microscopy, small-angle X-ray scattering, mercury<br />

porosimetry etc [2,11,53–59]. The thermodynamic method of<br />

standard contact porosimetry allows the determination of water<br />

distribution in the pores [59–63]. This method has been<br />

developed at the A.N. Frumkin Institute of Electrochemistry by<br />

Yu.M. Volfkovich and E.I. Shkolnikov. This method is useful<br />

for studying the properties of heterogeneous membranes<br />

with S-shaped isotherm of water desorption in a wide range of<br />

effective pore radii as well as that of homogeneous membranes<br />

MF-4SK and Nafion [59,60] with narrow distribution of pore<br />

sizes and low content of water.<br />

In the standard porosimetry method, the water content is<br />

measured for the membranes equilibrated with the standards of<br />

determined porous structure. The curves of integral distribution<br />

of pore volume on the pore radius were obtained for the<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

4 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

standards via independent techniques (mercury porosimetry or<br />

capillary condensation). The size of the molecules (about 1 nm)<br />

of the measuring liquid limits the use of this method.<br />

The specific internal surface area S (m 2 /g) and the distance<br />

between the fixed groups L (nm) were calculated from the<br />

porosimetry curves of the relative water content V (cm 3 Н2О/g)<br />

as a function of the effective radii of pores r (nm) in the<br />

membrane.<br />

Z l<br />

1<br />

S ¼ 2<br />

r2 0<br />

dV<br />

dlnr<br />

Z l<br />

dr ¼ 2<br />

0<br />

dV<br />

r<br />

; L ¼<br />

sffiffiffiffiffiffiffiffiffi<br />

S<br />

;<br />

QNA<br />

ð1Þ<br />

where dV is the differential of water content, Q is the ionexchange<br />

capacity, NA is the Avogadro's number.<br />

Schematic drawings of the electrochemical cells, the<br />

equations used for calculations and some experimental parameters<br />

are summarized in Table 2. The membrane conductivity<br />

κm (S/m) is a common characteristic of electrochemical properties<br />

of these systems. A wide range of experimental techniques<br />

has been developed for the determination of materials<br />

conductivity [17,64–66]. The value of κm is commonly determined<br />

from the value of membrane resistance measured as an<br />

active portion of membrane impedance [64]. We used the mercury-contact<br />

electrode operating at 200 kHz AC. The mercurycontact<br />

method is very attractive for studying polymer films<br />

since this technique provides an ideal contact between the<br />

electrodes and a sample. Moreover, this method is not limited<br />

by the concentration of the equilibrium solution or by the water<br />

content which might vary in different polymer materials.<br />

The polarization effects in membranes translate into the<br />

discrepancy between the conductivity of ion-exchange membrane<br />

measured under alternating current (κ AC) and that measured<br />

under direct current (κ DC) [34,67]. The simple equation<br />

that establishes the relationship between these characteristics<br />

has been derived and confirmed by the experimental data for<br />

heterogeneous electrodialytic membranes [68]:<br />

jDC ¼ jACt f2<br />

i<br />

where ti is transport numbers of counter-ions in solution; f2 is<br />

the volume fraction of equilibrium solution in the membrane.<br />

For homogeneous membranes, when f2→0, it is however<br />

possible to neglect discrepancies between κAC and κDC.<br />

The diffusion of salt through a membrane is described by<br />

several quantitative parameters (the salt diffusion flux ( jm, mol/<br />

m 2 s), the integral (Рm, m 2 /s) and differential (P⁎) coefficients<br />

of diffusion permeability). The relationships between these<br />

parameters are defined by the following equations:<br />

R c<br />

Pm ¼<br />

0 P⁎dc Dc<br />

; jm ¼ Pm ; jm ¼ P<br />

⁎ dc<br />

: ð3Þ<br />

c<br />

l<br />

dl<br />

Several papers regarding the experimental methods for measuring<br />

these diffusion characteristics have been published<br />

[17,69,70]. We studied the diffusion of salt solution through a<br />

membrane to the “pure” water (Table 2). The rate of increase of<br />

the salt concentration in a chamber filled with water (dc/dt) was<br />

controlled via conductometry method.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

ð2Þ


Electroosmotic permeability of the membranes in NaCl<br />

solutions W (m 3 /C) and the number of water transport t w (mol<br />

Н 2О/F) were measured in a two-chamber cell with reversible<br />

silver chloride electrodes by the volumetric method [19,71]<br />

(Table 2).<br />

The voltammetry curves were obtained in a cell equipped<br />

with platinum polarizing and silver chloride measuring electrodes<br />

[72]. The value of the electrodiffusion limiting current<br />

density i lim (A/m 2 ) was found graphically from the voltammetry<br />

curves.<br />

Testing of membranes was carried out in NaCl solutions,<br />

since this salt is one of the most abundant mineral component of<br />

natural waters, industrial and physiological solutions. For the<br />

electrodialytic membranes, the 0.1 М solution of NaCl was<br />

used as a standard. The membranes for other applications must<br />

be studied in more concentrated solutions of salts as well as in<br />

acid and base solutions (NaCl, HCl and H 2SO 4). Herein, the<br />

electrotransport characteristics of membranes were determined<br />

for a wide concentration range.<br />

The isothermal experiments were carried out at 25 °С. The<br />

measurement errors for all determined parameters amounted to<br />

3–5%.<br />

4. Results and discussion<br />

4.1. Structure of ion-exchange membranes investigated by<br />

standard porosimetry method<br />

Ion-exchange membranes are multiphase systems harboring<br />

polar and non-polar components in a polymeric matrix. Dry ionexchangers<br />

are dielectrics with a conductivity of 10 − 5 S/m. The<br />

conducting state of a membrane is induced by water or electrolyte<br />

solution that results in an abrupt increase in the conductivity<br />

of material (by 2–3 orders of magnitude) in a narrow<br />

range of water content values. This effect indicates the formation<br />

of ion and water transport channels in the membrane<br />

structure influenced by the external electric and concentration<br />

fields. The spatial distribution of water molecules and their<br />

specific concentration within the membrane both determine its<br />

structural, mechanical, thermodynamic and electrochemical<br />

properties. More recently, theoretical concepts regarding the<br />

ion aggregation in membranes have been suggested in several<br />

studies on the data obtained by modern physical methods<br />

[8,17,53,55,63,73–76].<br />

The analysis of the standard porosimetry data provides a<br />

rational picture of various interactions between the material and<br />

water. The model represented schematically in Fig. 2 includes<br />

energetic, hydration and geometrical parameters of membrane<br />

microstructure. The concept of “free” and “bonded” water in<br />

ion-exchangers allows simple calculations of the water capacity<br />

of a gel phase and hydration numbers of counter-ions [63].<br />

The “bonded” water confined in hydrate shells of ion–dipole<br />

associates fixed ion–counter-ion is immobilized in channels<br />

with effective radii ranging from 0.4 to 1.5 nm [77]. This water<br />

state is related to the hydrate capacity of a gel phase. The region<br />

of “bonded” water in the curve of water distribution on the pore<br />

radii corresponds to the energy of water bonding A≥1.7 kJ/mol<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

and to the effective value of pore radii r≤1.5 nm (A ¼ 2Vmr<br />

cos H<br />

r , where Vm is the molar volume of water, σ is the<br />

surface tension and Θ is the contact angle). This value of A is<br />

less by approximately two orders of magnitude than the hydration<br />

energy of ions that is comparable with the energy of<br />

hydrogen bonds [78]. The maximum in a region that corresponds<br />

to the macropores indicates that structural defects are<br />

formed at the interface between the resin and polyethylene<br />

particles.<br />

The method of standard porosimetry was successfully used in<br />

several studies of resin-based heterogeneous membranes with<br />

different porosities and degrees of DVB cross-linking. It is also<br />

useful for studying membranes with different chemistry of<br />

matrix, functional groups and counter-ions and for evaluating the<br />

degradation of membrane properties after freezing, sterilization<br />

etc [61,62,79]. The data presented in Fig. 3 confirm the high<br />

sensitivity of standard porosimetry with respect to the synthetic<br />

methods used for membrane fabrication. This sensitivity makes<br />

the standard porosimetry technique very convenient to determine<br />

the structural characteristics of new membrane materials.<br />

Currently, a wide range of Nafion-type perfluorinated membranes<br />

MF-4SK (Russia) is manufactured. The expansion of the<br />

class of novel membrane materials requires careful consideration<br />

of the relationships between cluster morphology and transport<br />

properties of membranes. The content and distribution of<br />

water in these materials are particularly important because they<br />

are commonly used as solid polyelectrolytes in fuel cells. The<br />

morphology of perfluorinated membranes and models of their<br />

structure are the subjects of continuing experiment and theorybased<br />

research. The generally accepted concept of water clusters<br />

(4–5 nm in diameter) connected by narrow channels of 1–<br />

1.5 nm has been extended recently to a new model of Nafion<br />

structure. This model suggests the formation of networked<br />

bundles of elongated polymeric aggregates and cylindrical<br />

channels in the membrane [80]. Fig. 4 shows the porosimetry<br />

Fig. 2. Integral (1) and differential (2) functions of water distribution on the pore<br />

radii and energy of bonded water (A, J/mol) for MK-40 membrane; V is the<br />

relative water content of the swollen membrane (cm 3 /g); r is the effective pore<br />

radius (nm); nm is the water capacity of the swollen membrane (mol H2O/site).<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

5


<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

6 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 3. Integral (a, b, c) and differential (a, d, e) functions of water distribution on the effective pore radii (r, nm) and energy of bonded water (A, J/mol)<br />

for heterogeneous membranes with various counter-ions (a, b, c) and altered DVB content (d and e). The hydrate microstructures are drawn schematically (a, b, c).<br />

a) MK-40 in Li + (1), Na + (2) and K + (3) ionic form; b) MK-40 in Na + (1), Ni 2+ (2), Cu 2+ (3) and Zn 2+ (4) ionic form; c) MK-41 in Na + (1), Ni 2+ (2), Cu 2+ (3) and Zn 2+<br />

(4) ionic form; d) MK-40 (2. 4 and 8% DVB); e) MK-41 (4 and 8% DVB).<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


Fig. 4. Distributions of integral pore-volume (V, cm 3 H 2O/cm 3 membrane) on pore radii (r, nm) for (1) Nafion-115, (2) Nafion-112, (3) Nafion-117 and (4) MF-4SK<br />

membranes (a) [60], and the scheme of Nafion membrane structure (b) [80].<br />

curves for membranes Nafion and MF-4SK described in [60],<br />

and a model of Nafion structure which accounts for membrane<br />

heterogeneity scaling from nanometers to microns.<br />

The effects of water redistribution between structural elements<br />

are even more pronounced when the organic ions are<br />

introduced into the membrane material. Fig. 5 represents the<br />

integral functions of water distribution on the effective pore radii<br />

for membranes MK-40 and MF-4SK with varied degree of<br />

Fig. 5. Integral functions of water distribution on effective pore radii for the<br />

membranes MF-4SK (a) and MK-40 (b). The saturation (θ) of membranes with<br />

ТBА + ions varies as follows: (a) (1) 0, (2) 0.17, (3) 0.25, (4) 0.49, (5) 0.70,<br />

(6) 1.00 and (b) (1) 0, (2) 0.24, (3) 0.72, (4) 0.88.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

membrane saturation with tetrabutylammonium ions (TBA + ).<br />

The degree of saturation (θ) is the ratio of the number of absorbed<br />

organic ions q (mol/g) to the maximum IEC of a membrane<br />

with respect to organic counter-ions (Qorg): θ=q/Qorg<br />

[81]. The values of internal specific surface S, distance between<br />

the fixed groups L and the amount of TBA + ions on the internal<br />

surface in membrane cs (cs=q/S) calculated from porosimetric<br />

curves for various θ values are summarized in Table 3.<br />

Several suggestions regarding the localization of organic<br />

ions at the internal interfaces in charged membranes have been<br />

made on the basis of the experimental data. Different scales of<br />

heterogeneity in polystyrene and perfluorinated membranes<br />

results in the accumulation of organic ions at the interface<br />

between gel phase and internal solution in the macro- and<br />

mesopores (Fig. 6a). This hypothetical interface is somewhat<br />

discrete because 40% of the membrane volume is occupied with<br />

polyethylene [33], which does not interact with organic counterions;<br />

instead, it is involved in the formation of transport<br />

channels.<br />

The saturation of MF-4SK membrane with TBA + ions is<br />

accompanied by the polymer dehydration and structural reorganization<br />

that decreases the diameter of hydrophilic channels<br />

approximately down to that of organic ions (1 nm) (Fig. 6b). This<br />

Table 3<br />

Structural characteristics of sulphocationic membranes with a varied degree of<br />

saturation with TBA + ions<br />

θ V0,cm 3 /g S, m 2 /g L, nm cs10 6 , mol/m 2<br />

MF-4SK<br />

0 0.27 146 0.50 0<br />

0.17 0.24 129 0.47 0.9<br />

0.25 0.20 72 0.34 2.4<br />

0.49 0.16 55 0.31 6.0<br />

0.70 0.13 33 0.24 14.5<br />

1.00 0.05 12 0.14 55.3<br />

MK-40<br />

0 0.46 121 0.27 0<br />

0.24 0.36 113 0.27 3.6<br />

0.72 0.33 99 0.25 13.8<br />

0.88 0.30 96 0.24 17.3<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

7


decrease in the distance between functional groups indicates the<br />

shrinking of material that inhibits the hydrophilic transport and<br />

results in the complete loss of membrane conductivity.<br />

Overall, the consideration of water distribution within the<br />

equilibrated membrane provides a comprehensive understanding<br />

of changes in electrotransport properties of membranes<br />

influenced by various internal and external factors.<br />

4.2. Ion transport in ion-exchange membranes<br />

4.2.1. Membrane conductometry<br />

The electrochemical behavior of ion-exchange membranes is<br />

generally described by a number of macroscopic properties:<br />

conductivity, transport numbers of ions and water, diffusion and<br />

electroosmotic permeability, value of limiting current in the<br />

current–voltage curve. These properties depend on the concentration<br />

of equilibrium electrolyte solution, which concentrations<br />

are altered in a course of separation process (Fig. 1).<br />

The membrane conductivity thus depends on the solution<br />

concentration, exchange capacity, type of counter-ions, fixed ions<br />

and polymeric matrix [17,19,43,45,82–84]. The most important<br />

factor is the dependence of membrane conductivity on concentration.<br />

The measurements of membrane conductivity in a wide<br />

range of NaCl concentration revealed three different concentration<br />

regions for the Κ M−С dependence (Fig. 7). For diluted<br />

solutions with the concentration ≤0.5M,anincreaseoftheΚ M<br />

value has been observed for all measured samples. When the<br />

membrane conductivity (κ M) is equal to the equilibrium solution<br />

conductivity (κ), the “isoconductivity” point (κ iso) is a characteristics<br />

of conducting properties of membranes in diluted solutions,<br />

jisoujm ¼ j: ð4Þ<br />

In most cases, the conductivity increased at the medium<br />

concentrations ranging from 0.5 M to 1.5 M. For the con-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

8 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 6. Schematic representation of the ТBА + ions immobilized in a heterogeneous cation-exchange membrane (а) and the ТBА + ions moving in cluster-channelcluster<br />

structure of a perfluorinated membrane (b).<br />

centrations exceeding 1.5 M, the extreme character of the Κ M−С<br />

function has been observed for some membranes [84]. The<br />

conductivity decrease in the concentrated solutions is related to<br />

the increase in the Donnan's sorption of electrolyte, membrane<br />

dehydration and ionic association. This study of conductivity VS<br />

concentration dependences (Figs. 7, 8) confirmed the key role of<br />

water molecules and their content in the membrane material on<br />

the conductivity of ion-exchange membranes.<br />

There are two different ways to vary water content in the<br />

membrane: (1) synthetically (by varying the degree of polymeric<br />

matrix cross-linking and the amount of inert component<br />

in heterogeneous membranes; by changing the conditions<br />

of alkaline saponification of sulphonyl fluoride groups or<br />

by varying the time of exposure in glycol at 110 °C for<br />

Fig. 7. Membrane conductivity vs concentration dependences obtained in<br />

equilibrium NaCl solutions: (1) MK-40 with 4% DVB, (2) MK-40 with 8%<br />

DVB, (3) MA-41, (4) MK-41 with 8% DVB, (5) MK-41 with 4% DVB, (6) BW-<br />

3361, (7) MA-40, (8) MF-4SK with nm=11.9, (9) MF-4SK with nm=36.5,<br />

(10) MF-4SK with nm=20.2, (11) MF-4SK with nm=7.1, (12) Kaspion.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


perfluorinated membranes), and (2) by post-synthetic processing<br />

of the already synthesized membranes (temperature treatment,<br />

pre-testing processing of perfluorinated membranes,<br />

immobilization of hydrophobic organic ions etc.) [19,41,43,<br />

47,63,71,85].<br />

The conductivity as function of water capacity for perfluorinated<br />

membranes preliminary rinsed with water is shown<br />

in Fig. 9. The nm value was varied both by changing the<br />

counter-ion type and by using different pre-testing procedures.<br />

The n m increased in the H + -form compared to that of K + -form,<br />

especially after successive boiling of membranes in acidic<br />

solutions and water [47]. The κm increases with nm (Fig. 9), that<br />

corroborates the general relationship between the hydrate characteristics<br />

and electrotransport properties of the ion-exchange<br />

membranes.<br />

The analysis of anion-exchange membranes conductivity has<br />

been described in details [36] and the type of functional groups<br />

and polymeric matrix have been suggested to be the key factors<br />

contributed to the ionic transfer.<br />

4.3. Percolation effects in ion-exchange membranes<br />

Studying the percolation transitions during the preparation of<br />

a polymeric composite is very important for the optimization of<br />

a composition of raw materials. The changes in conductivity are<br />

commonly observed when the ratio between the conducting and<br />

non-conducting components is varied for ion-exchange sticks<br />

(granulates), membranes from a resin and polyethylene, or<br />

membranes formed from sulpho-containing polyarylenamides<br />

(conducting component) dispersed in phenylone (inert component)<br />

[86].<br />

Percolation transition in the mixed system composed of<br />

swollen gel, solution and inert component particles is sche-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 8. Concentration dependences of the conductivity of electrodialytic membranes (a) and perfluorinated membranes reinforced with tetrafluoroethylene fibers (b): a)<br />

(1) Neosepta AM-1, (2) MK-40, (3) Neosepta CM-1, (4) CR 67-HMR-412, (5) MA-41, (6) AR 204-SZRA-412, (7) NaCl solutions; b) (1) Nafion-117, (2) Nafion-425,<br />

(3) MF-4SK, (4, 5, and 6) MF-4SK with fibers, (7) NaCl solution.<br />

matically represented in Fig. 10a. Fig. 10b shows a clear set of<br />

dependences of the relative conductivity value (κ m/κ) on the<br />

volume fraction of the conducting phase in the composite<br />

(factor fc).<br />

The equation proposed by S. Kirkpatrick on the basis of<br />

the percolation phenomena theory developed by S. Broadbent<br />

and J. Hammersly [87,88] has been used for the calculations of<br />

membrane parameters:<br />

jm ¼ j0 ð/ / jÞ<br />

t ; ð5Þ<br />

where κ0 is the parameter dependent on the mechanism of<br />

conductivity, which has been determined experimentally (the κ 0<br />

Fig. 9. Conductivity vs hydration capacity curve for perfluorinated membranes<br />

in various ionic forms.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

9


Fig. 10. 2D schematic representation of the swollen membranes as percolating<br />

systems (a). Percolation curves (b) for various membrane materials:<br />

1 — heterogeneous ion-exchange sticks (granulates), fabricated from extruded<br />

mixture of sulphocationic resin KU-2 and polyethylene; 2 — heterogeneous ionexchange<br />

membranes with the varied ratios of the resin KU-2 and polyethylene;<br />

3 — homogeneous cation-exchange films formed from sulphonate-containing<br />

polyaryleneamide.<br />

value is the same as electroconductivity of the conducting<br />

phase); ϕ and ϕ k are the volume fractions of conducting phase<br />

in the polymer and the critical value, at which the percolation<br />

transition isolator–conductor is observed; t is critical index of<br />

electroconductivity or universal constant.<br />

It has been found, that percolation transitions occur upon<br />

changing both water contents and the energetic state of water<br />

molecules in ion-exchange membranes. Fig. 11 (curve 3) illustrates<br />

percolation transition for the MK-40 membrane during<br />

freezing and thawing processes. We did not observe percolation<br />

effects for Neosepta СМ-1 and MF-4SK membranes under the<br />

similar experimental conditions [89].<br />

These observations suggest that the “free” water is responsible<br />

for the percolation nature of ion transport in hetero-<br />

Fig. 11. Time-dependent curve of the logarithm of specific membrane conductivity<br />

(κm, S/m) in a course of freezing–thawing processes: (1) MF-4SK,<br />

(2) Neosepta СМ-1, (3) МK-40.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

10 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

geneous membrane (MK-40). Dehydrated heterogeneous<br />

membranes exhibit an abrupt dependence of conductivity on<br />

water content. We observed this so-called “antipercolation transitions”<br />

in successively dried Nafion-117 and heterogeneous<br />

membranes MK-40, MA-41 and MA-41⁎ (isoporous structure<br />

of resin) equilibrated with water (Fig. 12), and membranes MA-<br />

40 equilibrated with sodium, nickel (II), copper (II) and zinc (II)<br />

chloride solutions [90–92]. As indicated in Fig. 12, the conductivity<br />

changed by more than four orders of magnitude with<br />

increasing water content. For highly hydrated membranes, the<br />

conductivity values depend on a chemistry of co-ions because<br />

of their complexation with the membrane material.<br />

The accumulation of organic ions in membranes causes the<br />

formation of structural elements with lower conductivity. This,<br />

in turn, induces the changes both in the membrane hydrophilicity<br />

and electrotransport properties. The membrane conductivity<br />

decreases in a course of saturating of the material with<br />

low-mobility organic counter-ions. In fact, the membranes with<br />

θ≤0.4 generally exhibit similar behavior (Fig. 13a). However,<br />

the ТBА + ions somewhat decrease the conductivity of sulphocationic<br />

membranes MK-40 and MF-4SK, when the saturation<br />

exceeds θ=0.4. We hypothesized, that this decrease was caused<br />

by an influence of membrane structure.<br />

For the system TBA + /MF-4SK, the change of membrane<br />

hydration state induced by the corresponding change in saturation<br />

with organic counter-ions results in an abrupt decrease in<br />

conductivity. This effect might be attributed to the percolation<br />

conductor–dielectric transition. The calculations on the percolation<br />

theory equation confirmed that the experimental value of<br />

parameter ϕk for all membranes correlated well with the<br />

theoretical ones [86]. The calculated value of the conductivity<br />

critical index (parameter t) however exceeded those mentioned<br />

in the literature for typical percolation systems. This discrepancy<br />

follows, most likely, from an approximate character of<br />

calculating the volume fraction (ϕ) of conducting component in<br />

perfluorinated membranes. The parameters in percolation theory<br />

equation do not correspond well to an actual geometry and<br />

packing of conducting regions in the membrane [93,94]. We<br />

assumed that the volume fraction of water, which formed<br />

Fig. 12. Dependence of the membrane conductivity (κ m, S/cm) on the water<br />

content: (1) MA-41, (2) MK-40 and (3) MA-41⁎ membranes impregnated with<br />

water [90].<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


hydrophilic channels for ion transfer, corresponded to the volume<br />

fraction of the conducting phase. The ϕ value was determined<br />

using the experimental dependences of the water content<br />

on the degree of saturation with organic ions and on membrane<br />

density [95,96].<br />

In summary, the introduction of TBA + ions into membrane<br />

material provided an opportunity for direct observation of new<br />

percolation conductor–dielectric transitions related to the<br />

interplay between the membrane dehydration and formation<br />

of the phase with low conductivity in homogeneous membranes.<br />

The small value of the t parameter indicates the strong<br />

influence of membrane dehydration on the type of a conductivity<br />

decrease. We note, that the similar effects were not<br />

observed for heterogeneous МK-40 membranes under the same<br />

experimental conditions.<br />

The experimental data on the conductivity of ion-exchange<br />

membranes in dry and swollen states in a wide range of electrolyte<br />

concentrations provide a basis for a clear set of most<br />

important factors for understanding polymer behavior: (1) critical<br />

value of water content corresponding to the percolation<br />

conductor–dielectric transition, (2) “isoelectroconductivity”<br />

point corresponding to the equilibration of the conductivity of<br />

membrane and that of solution, and (3) membrane conductivity<br />

maximum.<br />

4.4. Water electrotransport and selectivity of membranes<br />

The transport of ions through membranes under external<br />

electric field is accompanied by the transport of solvent molecules.<br />

The mechanism of water electrotransport depends on the<br />

membrane hydrophilic properties as well as on the degree of ion<br />

hydration [71]. The electroosmotic permeability of ion-exchange<br />

membranes has been studied systematically in several<br />

works [19,41,71,97,98].<br />

The concentration dependences of water transport numbers<br />

in a wide range of NaCl concentration for electrodialytic and<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

perfluorinated membranes are presented in Fig. 14. For diluted<br />

solutions, water transport numbers (t w), which depend strongly<br />

on the water content, remain constant for the majority of the<br />

investigated membranes. When the concentration of solution<br />

decreases, the tw increases substantially (up to 10 mol H2O/F) in<br />

strongly swelling membranes (MF-4SK, N M=36.5 mol H2O/<br />

mol SO3 − , KESD N M=19 mol H2O/mol SO3 − ). For the higher<br />

values of concentrations (from 0.5 M to 1.5 M of NaCl), this<br />

value remains stable in electrodialytic membranes. However,<br />

the decrease of t w was observed for some perfluorinated membranes.<br />

For the concentrated solutions, the t w value approaches<br />

the number of primary hydration of ions (4 for Na + and 3 for<br />

Cl − ) for all investigated membranes.<br />

The type of counter-ions may have an essential influence on<br />

the electroosmotic permeability of membranes. The high numbers<br />

of water transport by tetrametylammonium and trimethylbenzylammonium<br />

cations have been found in [98]. A similar<br />

effect has been described in [99] for Nafion-117 membrane<br />

saturated with alcylammonium cations (the number of hydrocarbon<br />

radicals varied from 1 up to 4). The authors hypothesized<br />

that electroosmotic flow increased because of hydrodynamic<br />

conditions that induced “the pumping effect”, i.e., the transport<br />

of the excessive amount of solvent. However, the detailed<br />

mechanism for this type of water transport has not been proposed.<br />

High electroosmotic flow has been also observed for<br />

dicarboxylic acids [100] and amino acids [101,102] transferred<br />

through membranes.<br />

Herein, we studied the transport of water by TEA + ions<br />

through perfluorinated MF-4SK membrane. The anomalous<br />

electroosmotic transport as compared to NaCl solution was<br />

observed. For tetraethylammonium chloride (TEACl), the tw<br />

value reached 22 mol of H2O/F that correlated well with the data<br />

obtained in [98,99]. To determine the contribution of “the<br />

pumping effect” to the overall water transport, we studied the<br />

electroosmotic transport in mixed NaCl and TEACl solutions by<br />

the method of isonormal series. A total electrolyte concentration<br />

Fig. 13. Conductivity vs saturation degree curves for membranes with immobilized organic counter-ions TBA + and DDS − : (1) MK-40, (2) MA-41, (3) MF-4SK<br />

(a). Graphical determination of the parameters of percolation theory (b). (The parameter κm refers to the membrane conductivity after saturation with the TBA + ions<br />

and parameter κm 0 refers to the same characteristic of initial membrane).<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

11


Fig. 14. Concentration dependence of the water transport numbers in NaCl solutions for (a) electrodialytic and (b) perfluorinated membranes: a) (1) MK-40 with<br />

8% DVB, (2) MK-40 with 4% DVB, (3) MK-41 with 8% DVB, (4) MA-41, (5) MA-40, (6) BW-3361; b) (1) MF-4SK with nm=11.9, (2) MF-4SK with nm=20.2,<br />

(3) MF-4SK with n m=36,5, (4) Nafion 425, (5) Nafion 117, (6) MF-4SK with n m=9.8, (7) MF-4SK with n m=10.1.<br />

was kept constant (0.1 M). The experimental results are shown<br />

in Fig. 15. The convex dependence of tw on the equivalent<br />

fraction of organic ions in the mixture indicates that, unlike<br />

individual NaCl and TEACl solutions, the t w values obtained<br />

for the mixture of Na + and TEA + ions are not additive ones The<br />

conservative value of tw can be related to the membrane selectivity<br />

toward organic counter-ions [103,104]. For the mixture of<br />

the TEA + ions with hydrophilic Na + , there are two parallel<br />

flows transferring ions through the membrane.<br />

Table 4 summarizes the equilibrium and dynamic hydration<br />

characteristics for the MF-4SK membrane (Na + -form) saturated<br />

with TEA + ions. The t w/n m coefficient [105] facilitates the<br />

comparison between dynamic and equilibrium hydration characteristics.<br />

The transport number of water, tw, characterizes the<br />

flow of solvent through the labyrinth of transport channels in a<br />

Fig. 15. Dependences of (1) the water transport numbers and (2) conductivity of<br />

the MF-4SK membrane on the equivalent fraction of TEACl in 0.1 M NaCl<br />

solution.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

12 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

membrane. The water capacity nm is determined by the amount<br />

of water absorbed by the membrane upon swelling under equilibrium<br />

conditions.<br />

As evidenced by the data given in Table 4, the t w/n m coefficient<br />

for Na + is smaller than 1. This value indicates that the<br />

water is only partially involved into electroosmotic transport<br />

with Na + ions. Here, the transport of water immobilized in the<br />

hydration shells of ions is the primary mechanism of electroosmotic<br />

process. For the TEA + , the tw/nm coefficient increases<br />

to 1.93 that suggests the additional mechanism of water trans-<br />

port. The electroosmotic transport in perfluorinated membranes<br />

is, most likely, induced by the pressure gradient ( dP)<br />

promoting<br />

dx<br />

the opening of additional channels that are highly selective to<br />

the transport of Na + ions [78].<br />

The formation of aqueous clusters around tetraalkylammonium<br />

ions in a free solution thus manifests itself in the electromigration<br />

of TEA + ions through the membrane. The proposed<br />

mechanism of water transport with Na + and TEA + ions through<br />

ion-exchange membrane is drawn schematically in Fig. 16. We<br />

suggest, that the ‘pumping effect” ( dP ¼ 0) does not occur during<br />

dx<br />

the translational motion of Na + ions when the water molecules<br />

are transported only in immobilized form, that is, in ion hydration<br />

shells.<br />

The function of the equivalent fraction of TEA + ions obtained<br />

by conductometry measurements for the membranes in<br />

a mixed Na + -TEA + form is shown in Fig. 15. Even small<br />

amounts of TEA + ions cause a considerable decrease in the<br />

Table 4<br />

Hydration characteristics of the MF-4SK membrane<br />

Ions Ion radius, nm n m, mol Н 2О/mol SO 3 −<br />

Na +<br />

TEA +<br />

t w, mol Н 2О/F t w/n w<br />

0.095 16.6 10.60 0.64<br />

0.41 11.4 22.92 1.93<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


membrane conductivity. This might be attributed to the membrane<br />

selectivity toward organic counter-ions. The degree of<br />

membrane saturation by TEA + ions tends to unity (θ→1) after<br />

the contact of the membrane with the TEACl solution of<br />

minimal concentration. A similar loss of conductivity was<br />

observed also for Nafion-117 membrane in NaBr and TMABr<br />

solutions [42]. The anomalous water transport through MF-<br />

4SK membrane and the simultaneous decrease in membrane<br />

conductivity can be therefore explained by specific interaction<br />

between tetraalkylammonium ions and water and by the pressure<br />

gradient acting as an additional force in the electromembrane<br />

system.<br />

The characteristics of water electrotransport are directly<br />

related to the selectivity of ion-exchange membranes. This<br />

parameter is especially important for the membranes used as<br />

ion-selective electrodes in electromembrane processes (e.g.<br />

water desalination, concentrating of electrolyte solutions etc.).<br />

The transport numbers of ions (ti ⁎) are quantitative and complex<br />

characteristics of the selectivity of membranes. This parameter<br />

represents the fraction of current transferred by certain ions and<br />

depends strongly on the experimental conditions. In fact, the<br />

effective transport numbers of ions [17] are “apparent” (t+app)<br />

because ion transport is accompanied by water transport. For<br />

this reason, some authors argue that these characteristics should<br />

not be used in technical data sheets. Nevertheless, the development<br />

of new methods to measure actual transport numbers still<br />

continues since these values are important for calculating the<br />

electric energy required for the electromembrane processes. A.<br />

Narebska et al. [4,106] analyzed the dependences of transport<br />

numbers of counter-ions on concentration and electric current<br />

density and suggested that this parameter as well as electroosmotic<br />

permeability might be included into the set of six<br />

“compulsory” properties for membrane characterization. The<br />

following equation suggested in [107] provides yet another<br />

approach for calculating transport numbers using the values of<br />

conductivity, diffusion and electroosmotic permeability coefficients:<br />

t ⁎ i ¼<br />

z2 þL⁎þ c ðÞ<br />

z2 þL⁎ þðÞþz c 2 L⁎ ; ð6Þ<br />

ðÞ c<br />

Here, the electrodiffusion coefficients Li ⁎ (с) are defined in the<br />

following equations using the values of the specific conductiv-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 16. Schematics illustrating the mechanism of water transport with Na + and TEA + ions in ion-exchange membrane.<br />

ity and diffusion permeability coefficient at certain concentration<br />

of the solution:<br />

L ⁎ þ c ðÞ¼j⁎ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ðÞ c<br />

2P<br />

⁎ 1 þ 1<br />

zþF 2 ⁎ðÞcF c 2<br />

RTj⁎ " s<br />

#<br />

; ð7Þ<br />

ðÞpF c ðÞ c<br />

L ⁎ ðÞ¼ c<br />

j⁎ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ðÞ c<br />

2P<br />

⁎ 1 1<br />

z F2 ⁎ðÞcF c 2<br />

RTj⁎ " s<br />

#<br />

: ð8Þ<br />

ðÞpF c ðÞ c<br />

where π ±(c) is the function accounted for the non-ideality of<br />

equilibrium solution:<br />

pFðÞ¼1 c þ dlngF=dlnc: ð9Þ<br />

The data on the ti ⁎ value calculations for membranes with<br />

various structures are summarized in Table 4. Herein, the t+ ⁎<br />

values were obtained using the well-known Scatchard equation<br />

[108]:<br />

t ⁎ þ ¼ tþapp þ 0:018mFtw; ð10Þ<br />

This equation establishes the relationship between “true” and<br />

“apparent” transport numbers of counter-ions and it accounts<br />

for the additional contribution of water transport. We used the<br />

experimental data obtained in our laboratory and those published<br />

in [43] for our calculations. The transport numbers calculated<br />

from the concentration dependences of specific conductivity and<br />

diffusion permeability of counter-ions are in very close agreement<br />

with those estimated from the experimental data with Eq.<br />

(10); the discrepancy amounted to 3–5% (Table 5). The concentration<br />

dependences of conductivity, diffusion and electroosmotic<br />

permeability provides enough data for calculations of<br />

ion transport numbers, and the experimental determination of<br />

these latter parameters is therefore not required.<br />

We successfully used this approach in [107] for perfluorinated<br />

membranes (MF-4SK) with various contents of water and<br />

for heterogeneous membrane МK-40 on polystyrene matrix.<br />

Our results as well as those of A. Narebska et al. for Nafion-120<br />

membrane [109,110] were used in a recent work [13] for the<br />

calculation of ion transport numbers. The authors suggested an<br />

equation, which was proposed earlier by R. Paterson and O.<br />

Kedem [9,10], though in different mathematical form.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

13


Table 5<br />

Transport numbers of counter-ions and water transport numbers for sulphocation<br />

membranes with varied type of polymer matrix in 0.5 M NaCl solution<br />

Membrane t + ⁎ calculation<br />

by Eq. (6)<br />

t+app<br />

tw, mol<br />

Н 2О/F<br />

SPEEK-1 0.99 0.97 4.8 0.99<br />

SPEEK-2 0.93 0.91 9.0 0.95<br />

SPS-2 0.99 0.96 4.0 0.98<br />

Nafion-117 0.97 – 4.0 –<br />

MF-4SK 0.98 0.94 10.3 0.99<br />

t+ ⁎ calculation<br />

by Eq. (10)<br />

Fig. 17 demonstrates the concentration dependence of the<br />

“true” transport numbers, which was calculated with Eq. (6) for<br />

sulphocationic membranes formed from polysulphone and polyether-ether-ketone<br />

matrix. The t+ ⁎ values indicated by dots were<br />

calculated from the experimental data on the water transport<br />

numbers and “apparent” transport numbers of ions for the same<br />

membranes [43]. Overall, this method enables, in principle, the<br />

determination of the “true” transport numbers of counter-ions for<br />

ion-exchange membranes with various structures.<br />

The concept of simultaneous water and ion transfer allows<br />

calculations of the hydration characteristics of ions from the<br />

concentration dependences of water transport numbers. It has<br />

been found that the dependence of transport number of water<br />

(transporting through a gel zone) on “true” transport number of<br />

ions is linear for many membranes when the transport number<br />

of counter-ions is below 0.9 [97]. The experimental results can<br />

be plotted in tw−t+ ⁎ coordinates according to the equation<br />

jw ¼ jþw þ j w ¼ t ⁎ þ hþ t ⁎ h i=F: ð11Þ<br />

The dynamic hydration numbers can be therefore found<br />

graphically: h − as an intersection point of the linear function<br />

with the ordinate axis (tw), and h+ as a tangent of the linear<br />

function slope to the x axis (Fig. 18). In Eq. (10), jw is a density<br />

of the water flow across the membrane, j+w and j−w are the<br />

densities of the water flow transferred by cations and anions, i is<br />

the current density, and F is the Faraday constant. The dynamic<br />

Fig. 17. Concentration dependence of the ion transport numbers in ion-exchange<br />

membranes with a varied degree of sulphurization (1) SPS-2 with nm=7.0, (2)<br />

SPEEK-1 with nm=13.7 and (3) SPEEK-2 with nm=20.0 mol H2O/mol.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

14 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 18. Water transport number vs transport number of counter-ions (Na + )<br />

correlation for various membranes: (1, 3) MF-4SK with nm=(1) 36.5 and<br />

(3) 20.2 mol H2O/mol, (2) SPEEK-2 and (4) MK-40.<br />

characteristics h+ and h− consist of two parts: (1) volume of<br />

water involved in hydration, and (2) volume of water, which<br />

participates in the electrotransport process.<br />

The calculated values of the dynamic hydration numbers of<br />

counter-ions and co-ions for sulphocationic membranes with<br />

different polymeric matrices and water capacity nm are presented<br />

in Table 6. The dynamic hydration numbers of counterions<br />

somewhat depend on the nm factor. The h+ values range<br />

from 7 to 11 mol Н2О/mol Na + , while the differences between<br />

the dynamic hydration numbers of co-ions Cl − are more pronounced.<br />

We note that the dynamic hydration numbers of<br />

counter-ions Na + listed in Table 5 are in agreement with those<br />

calculated with the model equations for the electroosmotic<br />

permeability [40,66]. In summary, the values of dynamic hydration<br />

numbers of ions depend on the water content in a<br />

membrane and therefore they must be mentioned among other<br />

technical characteristics.<br />

4.5. Diffusion permeability of membrane materials<br />

The concentration field induces the diffusion transfer of the<br />

electrolyte solution in a course of ion separation, which, in turn,<br />

influences the membrane behavior. The data on the salt diffusion<br />

into the chamber filled with distillate water are shown in<br />

Fig. 19. The integral coefficients of diffusion permeability as<br />

Table 6<br />

Dynamical hydration numbers of ions in membranes<br />

Membrane nm, mol<br />

Н2О/mol SO3 −<br />

hNa+, mol<br />

Н2О/mol Na +<br />

МK-40 13.0 6.9 7.0<br />

МF-4SK 36.5 11.4 29.2<br />

МF-4SK 20.2 7.0 11.2<br />

SPEEK-2 20.0 8.5 6.0<br />

hCl<br />

− ,mol<br />

Н2О/mol Cl −<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


Fig. 19. Concentration dependences of the integral diffusion permeability<br />

coefficient for various membranes in NaCl solution: (1) MK-40 with 8% DVB,<br />

(2) MA-40, (3) BW-3361, (4) Neosepta AM-1, (5) Neosepta AM-X, (6) MF-<br />

4SK with nm=11.9, (7) MF-4SK with nm=20.2, and (8) MF-4SK with<br />

nm=36.5.<br />

well as other transport characteristics of different membrane<br />

depend strongly on the concentration. The interpolymeric<br />

KESD membrane generally used for the diffusion dialysis exhibited<br />

the highest Р m value in all measured solutions (Р m of<br />

KESD membrane is 10 times higher than those of electrodialytic<br />

membranes). In the case of membranes with high water capacity<br />

(e.g., MC-3361 type), a pronounced maximum of Рm was<br />

observed in the concentration range from 0.5 to 1.5 M. These<br />

data suggest that the average water content and, importantly, the<br />

state and location of water molecules are the factors that<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

dominate the dependences of conductivity and diffusion permeability<br />

on concentration (Fig. 7). Fig. 20a shows the dependences<br />

of the diffusion flow j on concentration of NaCl<br />

solution for electrodialytic membranes. The similar curves of<br />

the integral coefficient of diffusion permeability for perfluorinated<br />

membranes filled with reinforcing fluorethylene fibers are<br />

shown in Fig. 20b.<br />

The concentration dependences of the diffusion permeability<br />

of about 40 membrane types have been studied systematically in<br />

[70]. The logarithm of j m value depends linearly on the logarithm<br />

of electrolyte concentrations. The tangent of these linear<br />

functions (β=dlnjm/dlnc) varies from 0.5 up to 1.5 depending<br />

on a type of membrane and chemistry of co-ions, which thereby<br />

influences the β value. For βN1, the concentration profile in a<br />

membrane is a convex function, and for βb1, it is of concave<br />

character (Fig. 21) — the inference, that has been supported by<br />

the results of theoretical calculations [65]. In the case of poorly<br />

hydrated membranes (SPS, SPEEK or Kaspion), the β value lies<br />

near 1. The similar β values were measured for MF-4SK membrane<br />

with ultra-low content of water (nm≈5 mol H2O/mol) and<br />

reinforced with thick fluorethylene network (Fig. 20b). The<br />

concentration profiles and the β parameter can be therefore used<br />

for the membrane characterization.<br />

For all membranes, the specific conductivity and diffusion<br />

permeability increase up to maximal critical value in concentrated<br />

solutions, while the activity of water and mobilities of<br />

counter-ions and co-ions decrease. Similarly, the water transport<br />

numbers and ion transport numbers decrease simultaneously<br />

with increasing concentration. For cb1 M, there is a defined<br />

region of stable tw values in tw−c curve. The threshold character<br />

of these curves is due to the difference in the physical<br />

properties of conducting phases in structurally heterogeneous<br />

membranes, e.g. viscosity.<br />

Fig. 20. Concentration dependences of diffusion flow for electrodialytic membranes (a) and integral diffusion permeability coefficient (b) for perfluorinated<br />

membranes in NaCl solution: a) (1) AM-1, (2) Neosepta СM-1, (3) МА-41, (4) MK-40, (5) CR 67-HMR-412, and (6) AR 204-SZRA-412 membranes; b) (1) Nafion-<br />

117, (2) Nafion-425, (3) MF-4SK, (4, 5, and 6) MF-4SK with tetrafluoroethylene fibers as the reinforcing material.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

15


Fig. 21. Concentration profile in membranes with varied β parameter: c is the<br />

solution concentration, c 1≈0, c 2 is the electrolyte solution concentration in the<br />

chamber of a cell, x is the transport axis and l is the thickness of membrane.<br />

The data on concentration dependences of electrotransport<br />

properties and those on critical values of physicochemical properties<br />

provide, in principle, a set of guidelines for the optimization<br />

of membranes testing protocols, membrane composition<br />

and concentration ranges corresponding to more efficient electrochemical<br />

separation.<br />

4.6. Polarization phenomena in electromembrane systems<br />

4.6.1. Membrane voltammetry<br />

The membrane voltammetry method is capable of measuring<br />

the properties of polarized membranes under the conditions<br />

approximating those of actual membrane operating. The<br />

literature on testing of membranes via current–voltage measurements<br />

(CVC) is however limited by only a few works [111–<br />

117]. This is due to the difficulties related to the correct interpretation<br />

of measured CVC signals, which are influenced by the<br />

variety of factors.<br />

The current, which reaches certain limiting value and then<br />

increases sharply, and the following system transition to the socalled<br />

“overlimiting” state are the distinguishing characteristics<br />

of CVC curves of ion-exchange membranes. Each portion of the<br />

voltammetric curve provides information about the electrotransport<br />

properties of the membrane. The ohmic portion helps to<br />

determine the membrane resistance. The value of the limiting<br />

electrodiffusion current (ilim) depends on the membrane selectivity,<br />

concentration and chemistry of the electrolyte solutions<br />

as well as on hydrodynamic conditions, which influence the<br />

thickness of the diffusion layer on the membrane surface. The<br />

increase of a current above i lim is caused by the coupled effects<br />

of the concentration polarization inducing the electro-, termoand<br />

gravitational convective flows. These effects are the generation<br />

of H + ,OH − ions and formation of the macroscopic spatial<br />

charge in solution at the membrane/solution interface, and the<br />

limiting current exaltation [118–128]. Hydrodynamic and structural<br />

features of the membrane surface dominate the behavior of<br />

the membrane at the overlimiting currents.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

16 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

In the case of cation-exchange membrane, the characteristics<br />

of hydrodynamic instability, which is proportional to the length<br />

of a CVC plateau and to the ratio of “overlimiting” resistance to<br />

“underlimiting” one, depend on the cation radii [129]. The<br />

intensity of hydrodynamic instability decreases when the<br />

Stokes' ion radius increases. The coupled convection, which<br />

manifests itself in the electrochemical noise on CVC curves,<br />

depends on the ratio of counter-ion/co-ion diffusion coefficients<br />

[130]. The parameters of CVC curves to be used for testing of<br />

membrane materials are as follows: the slope of the ohmic<br />

portion, the limiting current value, the length of limiting current<br />

plateau, and the potentials, at which the transitions into limiting<br />

and overlimiting state occur.<br />

It is well-known fact that the shape of current–voltage curves<br />

and the value of limiting current density (ilim) substantially<br />

depend on the amount and chemistry of surfactants [95,131–<br />

133]. The addition of amphiphiles alters significantly the<br />

properties of the external membrane/solution interface<br />

[134,135]. The effect of water redistribution within thin surface<br />

layers of membranes modified with surfactants has been found<br />

in [136] using the contact angle measurement and corresponding<br />

calculations of wetting energy. The adsorption of such<br />

compounds on a membrane surface can, in principle, favor the<br />

decrease as well as the increase of the limiting current compared<br />

to that for non-modified membrane in the NaCl standard<br />

(Fig. 22). We now have an original “scaling bar” based on<br />

Fig. 22. Current–voltage characteristics of (1, 3, 5) MK-40 and (2, 4, 6) MA-41<br />

ion-exchange membranes in (3, 4) 0.05 M NaCl solution and in the same<br />

solution containing various organic surfactants: (1, 2) camphor, (5) tetrabutylammonium<br />

iodide, and (6) sodium dodecyl sulphate.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


Fig. 23. Schematically drawn structure of a membrane/solution interface with<br />

adsorbed camphor. The drawing illustrates an effect of adsorbed substance on<br />

the diffusion layer thickness (from δ′ to δ″).<br />

limiting current values in the current–voltage curves that helps<br />

to evaluate the activity of organic surfactants in membranes<br />

[95,131].<br />

The presence of organic counter-ions decreases the value of<br />

ilim and diminishes the slope of ohmic portion of the CVC curve<br />

with respect to x axis. This is so, because the interactions<br />

between surfactants and ion-exchange membranes are accompanied<br />

by the Coulomb's forces thereby increasing the membrane<br />

resistance. The continuous loss of i lim observed in a<br />

course of membrane saturation with organic ions is related to<br />

electrostatic shielding effect. The membrane selectivity however<br />

decreases at large θ values and in surfactant solutions<br />

concentrated over CCM threshold. This selectivity loss translates<br />

into degeneration of the CVC dependencies, which become<br />

linear [137].<br />

The organic counter-ion can be immobilized in a membrane<br />

either at the equilibrium [72,95,135] or under applied electrical<br />

field [135,138]. The degree of membrane saturation might be<br />

controlled through changing the time, for which the membrane<br />

is exposed to surfactant solution, or by altering the amount of<br />

electrical charge transferred through the membrane. The chemistry,<br />

geometry and activity of organic ions contribute significantly<br />

to the electrochemical response of modified membranes;<br />

for instance, the measured values of the limiting current are<br />

different in a presence of TBA + and DDS − ions, respectively.<br />

Some substances cause, however, a significant increase in the<br />

ilim (Fig. 22). These substances belong to the class of camphorlike<br />

compounds (non-electrolyte) with high attraction constant<br />

[131,133,139]. Camphor itself is capable of assembling into 2D<br />

“islands” (or layers) on a membrane surface (Fig. 23).<br />

The role of the membrane surface on the coupled effects of<br />

concentration polarization remains an open question [134,140–<br />

145]. Nevertheless, modern physical methods confirmed that<br />

chemical and mechanical modification of the membrane surface<br />

changed its properties and behavior [124,143]. The cluster-like<br />

adsorption of camphor molecules on membrane surfaces results<br />

in strong electrical heterogeneity and, consequently, in hydrodynamic<br />

instability at the interface (Fig. 23). These effects are<br />

responsible for corresponding decrease in thickness of diffusion<br />

layer [146] that favors the enhancement of mass transport.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

These effects (so-called the Marangoni–Gibbs effects) are a<br />

well-known phenomena for the mercury electrode system<br />

[147,148]. We note, however, that the adsorption of camphor<br />

molecules on a polymer surface leads to the formation of specific<br />

discrete layer, which dimensions are comparable to those<br />

of electrical double layer.<br />

To establish the relationship between CVC parameters and<br />

physicochemical properties of membranes, we studied the<br />

properties of different MF-4SK membranes. Several methods<br />

were used to modify membranes:<br />

(a) various chemical pre-testing protocols,<br />

(b) immobilization of tetrabutylammonium cations,<br />

(c) polymerization of aniline in perfluorinated matrix.<br />

The shape of CVC curves and the value of ilim of MF-4SK<br />

membranes are both sensitive to pre-testing treatment protocol<br />

(Fig. 24). Thermal pretreatment causes an increase in i lim while<br />

the potential corresponding to the transition into the overlimiting<br />

state decreases. Several factors can affect the limiting<br />

current. Specifically, the water content increases in membranes<br />

after thermal processing that, in turn, induces an increase in<br />

conductivity and diffusion permeability [47]. Backward diffusion<br />

of the electrolyte solution can increase the ilim value in<br />

accordance with the following equation [149]:<br />

ilim ¼ DcF<br />

t⁎ ð i tiÞd<br />

þ P⁎cF t⁎ : ð12Þ<br />

ð i tiÞl<br />

where l is the membrane thickness, δ is the thickness of<br />

diffusion layer.<br />

The decrease in overlimiting potential is directly related to<br />

increased amount of water in the membrane. The data obtained<br />

with porosimetry method indicate that the number of pores with<br />

radii more than 100 nm is about 38% greater in thermally-<br />

Fig. 24. Current–voltage curves of MF-4SK membrane in 0.05 М HCl solutions<br />

after different pre-testing treatments: (1) thermal pretreatment and (2) salt<br />

pretreatment.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

17


treated membrane compared to that in membranes after the<br />

treatment with salt. These cavities in membrane material<br />

provide a free volume for unbound water which readily dissociates<br />

under applied electrical field.<br />

The current–voltage characteristics of the MF-4SK membranes<br />

variously saturated with the TBA + ions are represented<br />

in Fig. 25a. An increase in saturation degree θ results in the<br />

decrease in limiting current by two orders of magnitude. The<br />

slope of ohmic portion with respect to the x axis in CVC curve<br />

is also decreased thus pointing to the loss of the conductivity.<br />

Fig. 25b demonstrates the dependences of the limiting current,<br />

water transport number and conductivity on the saturation<br />

degree. The specific conductivity values obtained graphically<br />

from the ohmic portion of the curves correlate well with those<br />

measured by the contact technique using the alternating current.<br />

The conductivity decreases with the increasing θ (other memrane<br />

properties changes similarly) and the transition to the nonconducting<br />

state occurs at the θ=0.7. The i lim value decreases<br />

abruptly and the membrane becomes electrochemically inactive.<br />

Therefore, there is a clear percolation-overlimiting transition<br />

correlation in the perfluorinated MF-4SK membranes<br />

containing TBA + ions. Both effects are firmly related with the<br />

polymer dehydration induced by the immobilization of TBA +<br />

ion in the membrane cavities. The standard porosimetry results<br />

(Fig. 5) confirmed this suggestion. The change in saturation<br />

degree also causes the decrease in potential of overlimiting<br />

conductivity. This is so, because the TBA + ions act as catalytically<br />

active nitrogen-containing centers that form bipolar<br />

contacts [123]. The more pronounced transition of the system<br />

into the overlimiting state might arise because of the electroconvection<br />

resulted from electrical heterogeneity of the membrane<br />

surface [118,119].<br />

The development of the polymers with mixed electron-ion<br />

conductivity is important for fabricating technologically rele-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

18 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

vant materials (e.g. those for electrocatalysis, fuel cells, sensors<br />

etc.) [150–152]. Electrochemical behavior of composite membranes<br />

(MF-4SK/polyaniline) has been investigated by the<br />

membrane voltammetry [153]. The shift of overlimiting transition<br />

potential was clearly observed Fig. 26). In this particular<br />

case, the slopes of the ohmic portion of CVC curve and the ilim<br />

values are similar for the common membranes and composite<br />

ones.<br />

The observed effect might be attributed to the specific energetic<br />

state of water in the membrane structure. The experimental<br />

data obtained with standard porosimetry and those found in<br />

literature suggest that the water is immobilized in the structural<br />

cavities of a composite. The aromatic nature of polyaniline<br />

chains promotes the assembling of water molecules into clusters<br />

that inhibits the proton transfer. The polyaniline forms nanosize<br />

interpolymer complexes with ion–dipole associates in these<br />

cluster zones. This change in energetic state of water (water<br />

dissociation), most likely, hinders the transition of the system<br />

into the overlimiting state. The CVC parameters as well as all<br />

electrotransport properties therefore depend on the state of<br />

water.<br />

The analysis of electrochemical behavior of MF-4SK/<br />

polyaniline composites evidences the enhanced properties of<br />

this material when compared with common membrane matrices<br />

[154–156]. The composite membranes have high specific conductivity<br />

and selectivity and low diffusion permeability; the<br />

water immobilized in the interpolymer network, which is<br />

formed in such materials, is stable with respect to dissociation<br />

under applied electrical field. These properties make composites<br />

operationally more efficient for fuel cell systems and separation<br />

processes than the comparable non-composite systems.<br />

In summary, we believe, that the membrane voltammetry<br />

method has conclusively proven very useful for membrane<br />

testing and characterization.<br />

Fig. 25. Current–voltage curves for MF-4SK membrane with varied saturation with TBA + ions measured in 0.05 M HCl solution (a). Saturation dependences of<br />

dimensionless parameters: (1) water transport number, (2) density of limiting current, (3) conductivity from CVC (DC), and (4) conductivity under AC (b). (The<br />

parameter f refers to the membrane property after saturation with the TBA + ions and parameter f0 refers to the same characteristics of initial membrane): a) (1) θ=0,<br />

(2) θ=0.2, (3) θ=0.3, (4) θ=0.5.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


4.7. Theoretical modeling of electrotransport in heterogeneous<br />

membranes<br />

The behavior of membranes operating in real separation<br />

processes is influenced by the electric, concentration, and temperature<br />

gradients when the circulating flows of electrolyte<br />

solution pass through the inter-membrane space. Theoretical<br />

models of the electrotransport processes in membranes should<br />

therefore rely on an approach provided by the nonequilibrium<br />

thermodynamics which define the fluid flows as linear functions<br />

of generalized forces. Nonequilibrium membrane processes are<br />

accompanied by various cross-effects and associated phenomena<br />

(electro-, thermo-, barodiffusion, electro- and thermoosmosis<br />

etc.) [4,17,106,109,110,157–160]. The well-known Nernst–<br />

Planck equation<br />

ji ¼ L ⁎ i dA i=dx ð13Þ<br />

can be used as the differential transport equation for i-type ions<br />

that move under the electrochemical potential gradient (dµi/dx).<br />

The ji characterizes the ionic flow and the Li ⁎ is the coefficient<br />

of proportionality between the ionic flow and the force acting on<br />

the ion. Phenomenological coefficients Li ⁎ are the electrodiffusion<br />

characteristics of the ion mobility in heterogeneous<br />

medium. It has been suggested [161] that these factors are the<br />

functions of conducting properties of membrane “pseudophases”<br />

and of membrane geometry.<br />

The theoretical analysis of some known equations (Kedem–<br />

Katchalsky, Onsager and Nernst–Planck) in [13,17] evidences<br />

that the consideration of the structural heterogeneity of membranes<br />

is more important with respect to estimating transport<br />

values than the calculations using cross-coefficients — a traditional<br />

nonequilibrium thermodynamic approach to the description<br />

of trans-membrane transport [71,107].<br />

The studies on the transport properties vs solutions concentration<br />

that have been carried out for various membranes<br />

Fig. 26. Current–voltage curves of (1) MF-4SK membrane and composites<br />

formed from MF-4SK and polyaniline in the emeraldine form (2) and overoxide<br />

form (3) measured in 0.025 М H2SO4 solution.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

supported the assumption that the heterogeneity of membrane<br />

structure translated into the specific character of these dependences<br />

[17,19,65,71,84,162,163]. The Nernst–Planck equation<br />

was therefore extended to account for Li ⁎ coefficients — the<br />

functions dependent on the solution concentration. The mathematical<br />

form of the equation remains unaltered, while the Li ⁎<br />

coefficients are now considered as the kinetic electrodiffusion<br />

characteristics in the structurally heterogeneous medium.<br />

The closed system of the Nernst–Planck equations defines<br />

each ion with a charge z i participating in the ion transport and<br />

accounts for electroneutrality of membrane system. The<br />

theoretical model therefore relies on the determination of the<br />

flow values at certain parameters and boundary conditions in<br />

the system [149,164–167]. When Li ⁎ coefficients are defined as<br />

functions of concentration, they account for the structural properties<br />

of material. The values of Li ⁎ coefficients are determined<br />

from the experimental data on the concentration dependences<br />

of membrane conductivity (κ m) and salt diffusion flow ( j). The<br />

κ m−c and j−c dependences represent the power functions:<br />

jm ¼ Ac a<br />

ð14Þ<br />

jm ¼ Bc b : ð15Þ<br />

The parameters A, a, B and b are constants that depend on a<br />

membrane morphology and solution type. The electrodiffusion<br />

coefficients are expressed as follows [149,166]:<br />

L ⁎ ðÞ¼ c<br />

Bblzþ<br />

RTðzzþÞz cb<br />

ð16Þ<br />

L ⁎ þ c<br />

A<br />

ðÞ¼<br />

z2 ca<br />

ð17Þ<br />

2<br />

þF<br />

where l is the membrane thickness.<br />

This approach, which is based on concentration dependences<br />

of electrodiffusion coefficients of counter-ions and co-ions,<br />

provides an adequate description of electromasstransfer in the<br />

membrane system. The conductivity values, differential coefficients<br />

of diffusion permeability (P⁎) and the transport numbers<br />

of ions at certain concentration (с) are calculated with the<br />

following equations [149,166]:<br />

jm ¼ F 2 z 2 þL⁎þ c ðÞþz2L ⁎ ðÞ c<br />

ð18Þ<br />

P ⁎ ¼<br />

RTðz zþz c<br />

zþÞ<br />

t ⁎ i ¼<br />

1<br />

z 2 þ L⁎ þ c<br />

ðÞþ<br />

1<br />

z2 L⁎ ðÞ c<br />

ð19Þ<br />

z2 þL⁎þ c ðÞ<br />

z2 þL⁎ þðÞþz c 2 L⁎ : ð19:aÞ<br />

ðÞ c<br />

This approach has been recently corrected for a non-ideality<br />

of solution [168]. The authors suggested a set of equations to<br />

calculate the electrodiffusion coefficients L i ⁎ (с) from the data on<br />

κ m and P⁎ at randomly varied concentrations.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

19


However, the model based on the set of physically relevant<br />

parameters instead of concentration dependences of electrodiffusion<br />

coefficients L i ⁎ of the counter-ions and co-ions remains<br />

to be developed. In this respect, the charged membranes<br />

are now generally considered as the phase-separated systems in<br />

which transfer processes occur. The membrane harbors several<br />

structural fragments which number depends on the preparation<br />

protocol.<br />

For instance, the heterogeneous polymeric compositions for<br />

electrodialysis consist of ion-exchange resin, reinforcing polyethylene<br />

or polyvinylchloride and equilibrated solution distributed<br />

in membrane cavities. Perfluorinated ion-exchange<br />

membranes contain the hydrophobic matrix, charged groups<br />

and chains of side segments. The hydration of charged entities<br />

promotes the formation of ion-water clusters connected by<br />

narrow channels [8,53,78,169]. To model transfer phenomena<br />

in such heterogeneous structures, the number of structural<br />

elements is commonly considered as a single conducting phase<br />

[73,170,171]. The combination procedure varies in different<br />

studies. In some cases, the certain volume fraction of polymer is<br />

defined as a portion impermeable for the diffusion of particles<br />

[17,172]. The so-called winding factor is used to estimate the<br />

decrease in diffusion coefficient in a membrane when compared<br />

to that in a free solution. However, this model considering<br />

separate conducting and non-conducting phases helps to understand<br />

the decreasing conductivity alone, while the dependences<br />

of these properties on concentration of equilibrated solutions<br />

remain largely uncharacterized.<br />

The theory of generalized conductivity for heterogeneous<br />

system [161,173] provides a basis for quantitative characterization<br />

of the electrochemical properties of membranes. The effective<br />

properties are defined as the characteristics of separated<br />

phases by the interpolating function (Eq. (21)):<br />

L ⁎ m ¼ f1L a 1 þ f2L a 2<br />

1=a<br />

ð20Þ<br />

where Lm is the coefficient of generalized conductivity (electroconductivity<br />

or diffusion permeability coefficient); L 1, L 2 are<br />

the properties of individual phases; f 1 and f 2 are the volume<br />

fractions of phases; α is the parameter characterizing the arrangement<br />

of phases in the material (α=+1 for parallel orientation<br />

of conducting phases, α=−1 for serially orientated<br />

phases). The parameters f and α are related with the geometry of<br />

the dispersion.<br />

The structure of two conducting “pseudophases” is dictated by<br />

the mechanism of conductivity. The hypothetical phase I<br />

comprises: (1) gel phase with high conductivity, that is, gel/<br />

cluster structural fragments representing the ion–dipole associates<br />

fixed ion–counter-ion, and (2) inert, uncharged fragments<br />

(polyethylene and reinforcing material in heterogeneous membranes<br />

or microcrystallites of fluorocarbon chains and hydrophobic<br />

segments in the perfluorinated membranes). The ion<br />

transport in phase I is a flow of counter-ions. The volume fraction<br />

of this phase in the swollen membrane is equal to f 1=V 1/V sw.m.<br />

Phase II represents an internal “free” solution distributed in<br />

structural cavities and pores of the swollen membrane. This<br />

solution is not influenced by the local electrical fields of charged<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

20 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

polymer matrix. Both cations and anions are responsible for the<br />

charge transfer in phase II. The volume fraction of this phase is<br />

defined as f 2=V II/V sw.m. Consequently, both phases constitute a<br />

single integrated conducting structure ( f1+f2=1).<br />

When influenced by electric potential and concentration<br />

gradients, the properties of such system depend on the interplay<br />

between the spatial arrangement of both phases and the directions<br />

of electrical current or diffusion flow. The conducting<br />

phases can be assembled in parallel, serially or chaotically; the<br />

model accounts for the phase arrangement by using a specific<br />

parameter α. Eq. (20) represents the interpolation between the<br />

conductivity of the selected pseudophases.<br />

The first model representations of ion-exchange membrane as a<br />

biphase system have been published in the monographs [174,175]<br />

aimed to interpret correctly the conductivity of heterogeneous<br />

membranes. Similar approach has been used to describe conducting<br />

properties of the perfluorinated membranes with cluster<br />

morphology and those of membranes from aromatic polymers<br />

[41,43,71]. Fig. 27 shows the two-dimensional structural image of<br />

the membranes applied in electrodialysis. The volume fraction of<br />

phase I and an equilibrium solution (phase II) are also indicated.<br />

In the case of 1:1 electrolyte, the resulting membrane conductivity<br />

and diffusion permeability coefficients can be expressed<br />

as follows:<br />

jm ¼ f1j a iso þ f2j a 1=a ; ð21Þ<br />

P ⁎ ¼ f1ðGCÞ a þf2D a<br />

½ Š 1=a : ð22Þ<br />

Here κ iso and κ are specific conductivities of the solution<br />

and of phase I, respectively; c¯ −, c are co-ions concentrations in<br />

Fig. 27. Schematically drawn structure of a heterogeneous electrodialytic cationexchange<br />

membrane.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


Fig. 28. The data from the computer processing of the dependences of<br />

dimensionless characteristics Y and X at various solution concentration for<br />

membranes (1) Neosepta CL-25T, (2) MK-40, (3) SPEEK, (4) SPS, (5) CR 67-<br />

HMR-412, (6) MF-4SK (7) Nafion-117: ● — electroconductivity data<br />

(Yj ¼ jm<br />

jiso ; Xj ¼ j<br />

jiso ), ○ — diffusion permeability data (YP ¼ P⁎<br />

Gc ; XP ¼ D<br />

Gc ).<br />

phase I and in solution, respectively; D is the electrolyte diffusion<br />

coefficient. G is the complex parameter [176] which<br />

contains the Donnan constant (kD), the co-ion diffusion coefficient<br />

in the phase I (D¯ −) and the membrane exchange capacity<br />

P<br />

G ¼ kDD<br />

=Q: ð23Þ<br />

The G parameter plays the same role in the diffusion as does<br />

the κ iso parameter the conductivity. The κ iso is a characteristic<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

of the counter-ions transfer and the G is that of the co-ions<br />

transport in phase I.<br />

Using the conductivity and diffusion permeability functions<br />

normalized on the corresponding characteristics of phase I, the<br />

(21) and (22) equations can be generalized and represented as<br />

uniform function [177],<br />

Y ¼ f1 þ f2X a<br />

½ Š 1=a : ð24Þ<br />

For conductivity, we have the following equations:<br />

Yj ¼ j⁎<br />

juMj<br />

; Xj ¼ j<br />

juMj<br />

; ð25Þ<br />

and for diffusion permeability, the relationships are expressed as<br />

follows:<br />

Yp ¼ P⁎<br />

Gc ; Xp ¼ D<br />

: ð26Þ<br />

Gc<br />

These Eqs. (24)–(26) confirm the close similarity between<br />

the conductivity and diffusion transport.<br />

Fig. 28 shows the dependencies of the Y-function on the X<br />

for the number of ion-exchange membranes in logarithmic<br />

coordinates. The shape of the curves depends on some<br />

transport-structural parameters (TSP): f, κiso, α and G.<br />

To describe electromasstransfer in membrane systems correctly,<br />

the model must account for the electroosmotic phenomena.<br />

The transport of water is expressed by the dynamic<br />

transport numbers of water and ionic flows. Resulting set of<br />

TSP contains also the dynamical hydration characteristics of<br />

ions h+ and h− [107,178]. Membrane characterization based on<br />

Fig. 29. Scheme of calculation procedure for transport-structural parameters of ion-exchange membrane.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

21


Fig. 30. Dependencies of the transport-structural parameters on the water capacity for the set of membranes MF-4SK: κiso (a), G (b), f2 (c), α (d).<br />

concentration dependencies of specific conductivity, salt diffusion<br />

flux and electroosmotic permeability has been described in<br />

details in [179,180]. Fig. 29 represents a scheme for the determination<br />

of TSP-set of ion-exchange membrane. We note that<br />

the approach discussed here is based on the large set of our<br />

experimental data obtained for various membranes.<br />

The aforementioned method has been recently used to test<br />

new membrane materials. In [82], the polyamidoacid membranes<br />

have been characterized using f2 and κiso, parameters;<br />

similarly, the membranes fabricated from fibrous materials<br />

[181] and the heterogeneous membranes pre-treated with polycharged<br />

ions have been studied [182]. The behavior of MF-<br />

4SK/polyaniline composites, which comprises the properties of<br />

ionic and electronic conductors, has been broadly investigated<br />

with the TSP-set [155]. Yet another approach has been suggested<br />

in [183,184] to simplify calculations.<br />

The TSP-set is sensitively influenced by an average content<br />

of water and its distribution within the membrane. The dependences<br />

of the TSP of perfluorinated sulphocationic MF-4SK<br />

membranes on the water capacity are shown in Fig. 30. The<br />

water content was altered by changing the type and concentra-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

22 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

Fig. 31. Dependence of transport-structural parameters ( f2 and α) on a<br />

membrane structure.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


tion of base in a course of alkaline saponification of sulphonyl<br />

fluoride groups [85]. An increase in the water content enhances<br />

the transport of counter-ions (Na + ) and that of co-ions (Cl − )<br />

through phase I; these types of transport are related with the κiso<br />

and the G parameters, respectively (Fig. 30). Importantly, the<br />

enhanced transfer of co-ions can be accompanied by the loss of<br />

selectivity (Fig. 30b).<br />

In our model, the f2 parameter changes in the same manner<br />

as the κ iso and G parameters (Fig. 30c), while the dependence of<br />

the α parameter on the n m (Fig. 30d) exhibits an opposite trend.<br />

This behavior suggests an evolution of structural morphology<br />

from more or less regular arrangement to the chaotic one<br />

(α→0). The relationship between f and α parameters illustrated<br />

by Fig. 31 has been established using the database of the<br />

Membrane Materials Laboratory of Kuban State University.<br />

5. Conclusions<br />

This review summarized the state of field and general<br />

understanding of the physicochemical behavior of charged ionexchange<br />

membranes for a wide range of applications. In particular,<br />

we emphasized some of the important questions regarding<br />

structural and dynamic properties of ion-exchange<br />

membranes and discussed the theoretical and experimental elements<br />

that expanded the commonplace representation of these<br />

systems.<br />

Charged polymers and composite materials are the essential<br />

components of many developments related to rapidly advancing<br />

and demanding technological fields (fabrication of fuel cells,<br />

water purification and separation, analytical and medical<br />

sensing etc.). The outstanding progress in polymer synthesis<br />

during past decades considerably expands the types of materials<br />

used in electrochemical devices and the variety of functions<br />

exhibited by membrane structures. The development of common<br />

approach to optimize membrane structure and composition<br />

and to test its operating properties is therefore a significant<br />

challenge for modern analytical chemistry and material science.<br />

One strategy that provides a convenient experimentally-based<br />

data-set for comprehensive and reliable analysis of membrane<br />

properties has been described here in details. Further work will<br />

be, however, needed to generalize membrane handling protocols<br />

and to find a common guideline for rationalizing the diversity<br />

of structures used in electrochemical processes.<br />

Future progress of the technologies utilizing ion-exchange<br />

membranes will also depend on theoretical advancing the<br />

understanding of structural and dynamic aspects of existing<br />

systems and new types of membrane materials, especially those<br />

representing nanocomposites. The theoretical consideration of<br />

nanoscale structures must account for the interactions and<br />

processes at the multiple inner and external interfaces formed in<br />

hybrid (organic/inorganic) composite membranes.<br />

We believe that our approach for theoretical modeling of<br />

electrotransport in membranes makes it possible to characterize<br />

adequately the relationship between the structure and electrochemical<br />

properties of membrane materials. In this method, which<br />

is based on nonequilibrium thermodynamics consideration, the<br />

membrane structure is modeled with two conducting pseudo-<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

phases thus accounting for different mechanisms of ion transport.<br />

The structure–properties correlations and the proposed mechanisms<br />

of conductivity of heterogeneous membranes rely on the<br />

hypothesis that the phenomenological coefficients are the functions<br />

of polymeric structure. The model enables the characterization<br />

of electrochemical processes with a set of structurally-related<br />

parameters and provides a reasonably clear understanding of<br />

subtle interplay between the colloidal-scale structure and macroscopic<br />

properties of various membrane systems.<br />

Finally, we believe, that the development of new nanostructured<br />

materials will significantly encourage the theoretical<br />

and experimental progress in this field in coming years and that<br />

it will have a large impact on the way the research on ionexchange<br />

materials approach technologies.<br />

List of symbols<br />

A Energy of membrane-bonded water<br />

c Concentration of electrolyte solution<br />

c¯ − Co-ion concentration in phase I<br />

cs Amount of TBA + ions on the internal membrane surface<br />

Di Ion diffusion coefficient in solution<br />

D¯ − Co-ion diffusion coefficient in phase I<br />

F Faraday's constant<br />

f1 Volume fraction of phase I<br />

f2 Volume fraction of the inner solution phase<br />

G Gnusin's parameter, characterizing the diffusion properties<br />

of phase I<br />

h+ Dynamical hydration numbers of counter-ions<br />

h− Dynamical hydration numbers of co-ions<br />

i Current density<br />

ilim Electrodiffusion limiting current density<br />

ji Density of ionic flow<br />

jm Density of salt diffusion flow<br />

jw Density of water flow across membrane<br />

j +w Density of water flow transferred by cations<br />

Density of water flow transferred by anions<br />

j−w<br />

kD<br />

Donnan constant<br />

L Distance between the fixed groups<br />

Li ⁎ Phenomenological coefficients (electrodiffusion characteristics<br />

of the ion mobility in heterogeneous<br />

L m<br />

membrane)<br />

Membrane generalized conductivity coefficient (electroconductivity<br />

or diffusion permeability coefficient)<br />

L1, L2 Conducting properties of individual phases<br />

l Membrane thickness<br />

NA Avogadro's constant<br />

nm Membrane water capacity<br />

P⁎ Differential coefficient of diffusion permeability<br />

Integral coefficient of diffusion permeability<br />

Р m<br />

Q Ion-exchange capacity (IEC)<br />

Q org<br />

Ion-exchange capacity of membrane with respect to<br />

organic counter-ions<br />

q Amount of absorbed organic ions<br />

r Effective radius of pores<br />

S Area of the internal specific surface<br />

t Critical index of conductivity, or universal constant in<br />

the percolation theory<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

23


ti Transport number of counter-ions in solution<br />

ti⁎ Transport number of counter-ions in membrane<br />

t +app “Apparent” transport number of counter-ions in<br />

membrane<br />

tw Electroosmotic transport number of water<br />

V Relative water content<br />

Vm Molar volume of water<br />

W Electroosmotic permeability<br />

w Water content<br />

zi Ion charge<br />

α Parameter characterizing the spatial arrangement of<br />

membrane phases<br />

β Slope of the concentration dependences of the salt<br />

diffusion flow in logarithmic coordinates<br />

δ Diffusion layer thickness<br />

θ Membrane saturation with organic ions<br />

Θ Contact angle<br />

κ Conductivity of electrolyte solution<br />

κ0 Parameter of the percolation theory (conductivity)<br />

κm Conductivity of membrane<br />

κiso Conductivity of phase I and at isoelectroconductivity<br />

point<br />

κAC Conductivity measured under alternating current<br />

κDC Conductivity measured under direct current<br />

µ i Electrochemical potential<br />

π ±(c) Function accounted for non-ideality of equilibrium<br />

solution<br />

σ Surface tension<br />

ϕ Volume fractions of conducting phase in the polymer<br />

Critical value of volume fraction of conducting phase<br />

ϕk<br />

Acknowledgments<br />

We gratefully thank the Russian Foundation for Basic<br />

Research for the financial support (projects N 06-08-01424 and<br />

N 06-03-96675) and to Professor Victor Starov for the kind<br />

invitation in contributing to this special issue of the Journal.<br />

References<br />

[1] Strathmann H. Ion-exchange membrane separation processes. Membrane<br />

Science and Technology Series, 9, Elsevier; 2004.<br />

[2] Mulder MHV. Introduction in the membrane technology. Mir; 1999<br />

[in Rus.].<br />

[3] Hwang ST, Kammermeier K. Membrane separation processes. Mir; 1981<br />

[in Rus.].<br />

[4] Membrany i membranowe techniki rozdzialu. Praca zbiorowa pod<br />

redakcja A.Narebskiej. Torun; 1997.<br />

[5] Grebenyuk VD. Electrodialysis. Technika; 1976 [in Rus.].<br />

[6] Nagarale RK, Gohil GS, Shahi VK. Adv Colloid Interface Sci 2006;119:97.<br />

[7] Xu Tongwen. J Membr Sci 2005;263:1.<br />

[8] Heitner-Wirguin C. J Membr Sci 1996;120:1.<br />

[9] Kedem O, Perry M. J Membr Sci 1983;14:249.<br />

[10] Paterson R, Gardner C. J Chem Soc A 1971:2254.<br />

[11] Pusch W, Walch A. Angew Chem 1982;94:670.<br />

[12] Buvet R, Buvet D, Hafsi M. Actes du symposium international sur les<br />

membranes et procedes a membranes denses fonctionnalisees. France:<br />

Pont-a-Mousson; 1991. p. P.3*1.<br />

[13] Auclair B, Nikonenko V, Larchet C, Métayer M, Dammak L. J Membr<br />

Sci 2002;195:89.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

24 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

[14] Koter S, Piotrowski P, Kerres I. J Membr Sci 1999;153:83.<br />

[15] Yeo RS, Yeager HL. Structural and transport properties of perfluorinated<br />

ion exchange membranes. In: Conway BE, et al, editor. Modern aspects<br />

of electrochemistry. Butterworth, L.; 1985<br />

[16] Foley T, Klinowski J, Meares P. Proc R Soc 1974;A336:327.<br />

[17] Zabolotsky VI, Nikonenko VV. Transport of ions in membranes. Nauka;<br />

1996 [in Rus.].<br />

[18] Yaroslavtsev AB, Nikonenko VV, Zabolotsky VI. Uspekhi khimii (Russ<br />

Chem Rev) 2003;72(5):438.<br />

[19] Gnusin NP, Berezina NP, Dyomina OA, Kononenko NA. Elektrokhimiya<br />

(Russ J Electrochem) 1996;32:173.<br />

[20] Zawodzinski ThA, Springer TE, Uribe F, Gottesfeld S. Solid State Ionics<br />

1993;60:199.<br />

[21] Gregor HP. J Pure Appl Chem 1968;16:329.<br />

[22] Komkova EN, Wessling M, Krol J, Strathmann H, Berezina NP.<br />

Vysokomolekularnye soedineniya (High Molec Compd) 2001;43:486 (А).<br />

[23] Kirsh YuE. Vysokomolekularnye soedineniya (High Molec Compd) A<br />

1993;35:163.<br />

[24] Electrochemistry of polymers. In: Tarasevich MR, Orlov SB, Shkolnikov<br />

EI, et al, editors. Nauka; 1990. [in Rus.].<br />

[25] Nagarale RK, Gohil GS, Shahi VinodK, Trivedi GS, Rangarajan R.<br />

J Colloid Interface Sci 2004;277:162.<br />

[26] Tan S, Bélanger D. J Phys Chem 2005;109:23480.<br />

[27] Tan S, Viau V, Cugnod D, Belanger D. Electrochem Solid-State Lett<br />

2002;5:E55.<br />

[28] Amado FDR, Gondran E, Ferreira JZ, Rodrigues MAS, Ferreira CA.<br />

J Membr Sci 2004;234:139.<br />

[29] Neosepta. Ion-exchange membranes. Tokuyama Soda Co. Ltd; 1979.<br />

[30] Aldrich. Germany catalogue. Handbook of Fine Chemicals; 1999–2000.<br />

[31] Ionics. Bulletins. Cation-Transfer Membranes. Anion-Transfer Membranes.<br />

USA, Watertown: Ionics, Incorporated; 1990.<br />

[32] Selemion. Ion-exchange membranes. Japan: Asahi Glass Co.; 1984.<br />

[33] Ion-exchange membranes. Granulates. Powders, Catalogue. NIITEKHIM;<br />

1977 [in Rus.].<br />

[34] Data Sheet Membrane. RALEX MEGA; 2005.<br />

[35] Kopilova VD, Mekvabishvili TV, Gefter EL. Phosphorsoderzhaschie<br />

ioniti, Voronezsh; 1992 (in Rus.).<br />

[36] Dyomina OA, Berezina NP, Сата Т, Dyomin AV. Elektrokhimiya (Russ<br />

J Electrochem) 2002;38:1002.<br />

[37] Pozniak G, Trochimczuk W. Angew Makromol Chem 1984;127:171.<br />

[38] Grot WG. Nafion as a separator in electrolytic cells. Presented at The<br />

Electrochemical Society Meeting, Boston, Massachusetts; 1986.<br />

[39] Grot WC. The 1987 International congress on membranes and membrane<br />

processes. Du Pont Company; 1987.<br />

[40] Mazanko AF, Kamaryan GM, Romashin OP. Industrial membrane<br />

electrolysis. Khimiya; 1989 [in Rus.].<br />

[41] Berezina NP, Timofeev SV, Dyomina OA, Ozerin AN, Rebrov AV.<br />

Elektrokhimiya. Russ J Electrochem 1992;28:1050.<br />

[42] Berezina NP, Timofeyev SV, Rollet. A-L, Fedorovich NV, Duran-Vidal S.<br />

Elektrokhimiya (Russ J Electrochem) 2002;38:1009.<br />

[43] Berezina NP, Komkova ЕN. Kolloid Zh (Russ Colloid J) 2003;65:5.<br />

[44] Balster J, Krupenko O, Pünt I, Stamatialis DF, Wessling M. J Membr Sci<br />

2005;263:137.<br />

[45] Berezina NP, Karpenko LV, Dyomina OA, Kirsh YuE, Komkova ЕN.<br />

Elektrokhimiya (Russ J Electrochem) 1997;33:596.<br />

[46] Saldadze KM, Pashkov AB, Titov VS. Ion-exchange high-molecular<br />

compounds. Gosrhimizdat; 1960 [in Rus.].<br />

[47] Berezina NP, Timofeev SV, Kononenko NA. J Membr Sci 2002;209:509.<br />

[48] GOST 17552-72. Ion-exchange membranes. Techniques to meter the total<br />

and equilibrium exchange capacity. Izdatelstvo standartov; 1972 [in Rus.].<br />

[49] GOST 19180-73. Ion-exchange membranes. Techniques to determine<br />

sizes change at swelling. Izdatelstvo standartov; 1973 [in Rus.].<br />

[50] Sidorova MP, Ermakova LE, Cavina IA, Bogdanova NF, Timofeev CV.<br />

Kolloid Zh (Russ Colloid J) 1999;61:829.<br />

[51] Ermakova LE, Sidorova MP, Kiprianova AA, Savina IA. Kolloid Zh<br />

(Russ Colloid J) 2001;63:43.<br />

[52] Pourcelly G, Oikonomou A, Gavach C, Hurwitz HD. J Electroanal Chem<br />

1990;287:43.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002


[53] Timashev SF. Physical chemistry of membrane processes. Khimiya; 1988<br />

[in Rus.].<br />

[54] Halim J, Buchi FN, Haas O, Stamm M, Scherer GG. Electrochim Acta<br />

1994;39:1303.<br />

[55] Mizutani Y. Structure of ion-exchange membranes. J Membr Sci<br />

1990;49:121.<br />

[56] Yeager HL, Steck A. J Electrochem Soc 1981;128:1880.<br />

[57] Yeo SC, Eisenberg A. J Appl Polym Sci 1977;21:875.<br />

[58] Yeo RS, Yeager HL. In: Conway BE, et al, editor. Modern aspects of<br />

electrochemistry; 1985. L.: Butterworth.<br />

[59] Volfkovich YuM, Bagotzky VS, Sosenkin VE, Blinov IA. Colloids Surf<br />

A Physicochem Eng Asp 2001;187–188:349.<br />

[60] Divisek J, Mazin M EikeV, Schmitz H, Stimming U, Volfkovich YuM.<br />

J Electrochem Soc 1998;145:2677.<br />

[61] Kononenko NA, Berezina NP, Volfkovich YuM, Shkolnikov EI, Blinov<br />

IA. Zh Prikl Khim 1985;58:2199.<br />

[62] Berezina NP, Volfkovich YuM, Kononenko NA, Blinov IA. Elektrokhimiya<br />

(Russ J Electrochem) 1987;23:912.<br />

[63] Berezina NP, Kononenko NA, Volfkovich YuM. Elektrokhimiya (Russ J<br />

Electrochem) 1994;30:366.<br />

[64] Meshechkov AI, Dyomina OA, Gnusin NP. Elektrokhimiya (Russ J<br />

Electrochem) 1987;23:1452.<br />

[65] Gnusin NP, Berezina NP. Zh Phyz Khim (Russ J Phys Chem)<br />

1995;69:2129.<br />

[66] Karpenko LV, Dyomina OA, Dvorkina GA, Parshikov SB, Larchet C,<br />

Auclair B, Berezina NP. Elektrokhimiya (Russ J Electrochem)<br />

2001;37:328.<br />

[67] Shaposhnik VA. Kinetika elektrodializa, Voronezsh; 1989 (in Rus.).<br />

[68] Gnusin NP, Dyomina OA, Meshechkov AI, Tur'yan IYa. Elektrokhimiya<br />

(Russ J Electrochem) 1985;21:1525.<br />

[69] Nikolaev NI. Diffuziya v membranach. Khimiya; 1980 [in Rus.].<br />

[70] Gnusin NP, Berezina NP, Shudrenko AA, Ivina OP. Zh Phys Khim (Russ<br />

J Phys Chem) 1994;68:565.<br />

[71] Berezina N, Gnusin N, Dyomina O, Timofeyev S. J Membr Sci<br />

1994;86:207.<br />

[72] Loza NV, Kononenko NA, Shkirskaya SA, Berezina NP. Elektrokhimiya<br />

(Russ J Electrochem) 2006;42:907.<br />

[73] Water in polymers. In: Rowland S, editor. Mir; 1984. [in Rus.].<br />

[74] Mauritz KA. JMS — Rev Macromol Chem Phys 1988;C28(1):65.<br />

[75] Mauritz KA, Hopfinger A. Mod Asp Electrochem 1982;14:485.<br />

[76] Falk M. Can J Chem 1980;58:1495.<br />

[77] Chizmadzhev YuA, editor. Membrani: ionnie kanali. Mir; 1981. [in Rus.].<br />

[78] Eisenberg A, editor. Ions in polymers. Advances in chemistry series.<br />

Washington: ACS; 1980. N 187.<br />

[79] Zabolotsky VI, Berezina NP, Kononenko NA, Komkova EN, Shudrenko<br />

AA, Skuridin MA. Zh Prikl Khim 1997;70:1619.<br />

[80] Rubatat L, Rollet A-L, Gebel G, Diat O. Abstracts of International<br />

Congress on Membranes and Membrane Processes, ICOM Toulouse;<br />

2002. p. 152.<br />

[81] Berezina NP, Kononenko NA, Gnusin NP. Synthetic polymeric<br />

membranes. In: Sedlacek B, Kahovec J, editors. Berlin-New York:<br />

Walter de Gruyter; 1987. p. 71.<br />

[82] Dyakonova OV, Kotov VV, Voitshev VS, Bobreshova OV, Aristov IV.<br />

Elektrokhimiya (Russ J Electrochem) 1999;35:500.<br />

[83] Kotov VV, Dyakonova OV, Sokolova SA, Volkov VI. Elektrokhimiya<br />

(Russ J Electrochem) 2002;38:994.<br />

[84] Gnusin NP, Dyomina OA, Berezina NP, Meshechkov AI. Elektrokhimiya<br />

(Russ J Electrochem) 1988;24:364.<br />

[85] Ivina OP, Shokhman MYa, Berezina NP, et al. Zhurn Phys Khim (Russ J<br />

Phys Chem) 1992;66:2758–62.<br />

[86] Berezina N, Karpenko LV. Percolation effects in ion exchange materials.<br />

Kolloid Zh (Rus Colloid J) 2000;62:749.<br />

[87] Broadbent SR, Hammersley JM. Proc Camb Phie Soc 1957;53:629.<br />

[88] Kirkpatrick S. Phys Rev Lett 1971;27:1722.<br />

[89] Kononenko NA, Zabolotsky VI, Berezina NP, Kolodkina NV. Kondensirovannye<br />

sredy i mezhfaznye granitsy 2003;5:85 (in Rus.).<br />

[90] Volfkovich YuM, Khozyainova NS, Elkin VV, Berezina NP, Ivina OP,<br />

Mazin VM. Elektrokhimiya (Russ J Electrochem) 1988;24:344.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

[91] Gavach C, Pamboutzglou G, Nedyalkov N, Pourcelly G. J Membr Sci<br />

1989;45:37.<br />

[92] Pechenkina ES, Shishkina SV, Berezina NP, Medvedchikova AV. Book<br />

of Abstracts of Russian Meeting on Galvanic. Izdatelstvo Kirov<br />

universitet; 2003. p. 79. [in Rus.].<br />

[93] Hsu WI, Berzins I. J Polym Sci Polym Physics Ed 1985;23:933.<br />

[94] McLachian DS, Blaszkiewicz M, Newnham RE. J Am Ceram Soc<br />

1990;73:2187.<br />

[95] Kononenko NA, Berezin NP, Loza NV. Colloids Surf A Physicochem<br />

Eng Asp 2004;239:59.<br />

[96] Kononenko NA, Berezina NP, Shkirskaya SA. Kolloid Zh (Russ Colloid<br />

J) 2005;67:485.<br />

[97] Berezina NP, Dyomina OA, Gnusin NP, Timofeev SV. Elektrokhimiya<br />

(Russ J Electrochem) 1989;25:1467.<br />

[98] Kitchener DA. New problem of modern electrochemistry. In: Bockris J,<br />

editor. Inostrannaya literature; 1963. [in Rus.].<br />

[99] Xie G, Okada T. Electrochimica Acta 1996;41:1569.<br />

[100] Reshetnikova AK, Rozhkova MV, Kotov VV, Akimenko IB. Elektrokhimiya<br />

(Russ J Electrochem) 1996;32:200.<br />

[101] Kulintsov PI, Bobreshova OV, Aristov IV, Novikova IV, Khrikina LA.<br />

Elektrokhimiya (Russ J Electrochem) 2000;36:365.<br />

[102] Novikova IV, Kulintsov PI, Bobreshova OV, Bobilkina OV. Elektrokhimiya<br />

(Russ J Electrochem) 2002;38:1016.<br />

[103] Helferich F. Ionenaustauscher. Verlag Chemie; 1959.<br />

[104] Ion exchange. In: Marinsky LA, editor. Mir; 1968. [in Rus.].<br />

[105] Spiegler KS. Trans Faraday Soc 1958;54:1408.<br />

[106] Narebska A, Kujawski W, Koter S. J Membr Sci 1987;30:125.<br />

[107] Gnusin NP, Dyomina OA, Berezina NP, Parshikov SB. Theoria i praktika<br />

sorbtsyonnych processov (Theory and Practice of Sorption Processes),<br />

vol. 25; 1999. p. 213. (in Rus.).<br />

[108] Hladik J, editor. Physics of electrolytes. London: Academic Press; 1972.<br />

[109] Narebska A, Koter S, Kujawski W. Desalination 1984;51:3.<br />

[110] Narebska A, Koter S, Kujawski W. J Membr Sci 1986;25:153.<br />

[111] Balavadze EM, Bobreshova OV, Kulintsov PI. Uspekhi khimii (Russian<br />

Chem Rev) 1988;57:103.<br />

[112] Dyakonova OV, Kotov VV, Voitshev VS, Bobreshova OV, Aristov IV.<br />

Elektrokhimiya (Russ J Electrochem) 2000;36:81.<br />

[113] Kang M-S, Choi Y-J, Choi I-J, Yoon T-H, Moon S-H. J Membr Sci<br />

2003;216:39.<br />

[114] Choi J-H, Strathmann H, Park J-M, Moon S-H. J Membr Sci<br />

2006;268:165.<br />

[115] Chamoulaud G, Bélanger D. J Colloid Interface Sci 2005;281:179.<br />

[116] Park J-S, Choi J-H, Yeon K-H, Moon S-H. J Colloid Interface Sci<br />

2006;294:129.<br />

[117] Choi Y-J, Kang M-S, Kim S-H, Cho J, Moon S-H. J Membr Sci<br />

2003;223:201.<br />

[118] Rubinstein I, Warshawsky A, Schechtman L, Kedem O. Desalination<br />

1984;51:55.<br />

[119] Rubinstein I, Zaltzman B, Pretz J, Linder C. Elektrokhimiya (Russ J<br />

Electrochem) 2002;38:956.<br />

[120] Kedem O, Rubinstein I. Desalination 1983;46:185.<br />

[121] Krol JJ, Wessling M, Strathmann H. J Membr Sci 1999;162:145.<br />

[122] Kharkats YuI. Elektrokhimiya (Russ J Electrochem) 1985;21:974.<br />

[123] Zabolotsky VI, Sheldeshov NV, Gnusin NP. Uspekhi khimii (Russ Chem<br />

Rev) 1988;57:1403.<br />

[124] Shaposhnik VA, Vasil'eva VI, Grigorchuk OV. Transfer phenomena in<br />

ion-exchange membranes. MFTI; 2001 [in Rus.].<br />

[125] Zabolotsky VI, Nikonenko VV, Pismenskaya ND, Laktionov EV,<br />

Urtenov MKh, Strathmann H, Wessling M, Koops GH. Separation and<br />

Purification Technology 1998;14:255.<br />

[126] Zabolotsky VI, Berezina NP, Nikonenko VV, Shaposhnik VA, Tskhai<br />

AA. Membranes 1999:6 (in Rus.).<br />

[127] Nikonenko VV, Pis'menskaya ND, Volodina EI. Elektrokhimiya (Russ J<br />

Electrochem) 2005;41:1351.<br />

[128] Lee H-J, Strathmann H, Moon S-H. Desalination 2006;190:43.<br />

[129] Choi J-H, Lee, Moon S-H. J Colloid Interface Sci 2001;238:188.<br />

[130] Ellatar A, Elmidaoui A, Pismenskaia N, Gavach C, Pourcelly G. J Membr<br />

Sci 1998;143:249.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002<br />

25


[131] Berezina NP, Fedorovich NV, Kononenko NA, Botukhova GN,<br />

Komkova EN. Elektrokhimiya (Russ J Electrochem) 1999;35:103.<br />

[132] Lebedev KA, Kononenko NA, Berezina NP. Kolloid Zhurn (Russ Colloid<br />

J) 2003;65:232.<br />

[133] Berezina NP, Stenina EV, Fedorovich NV, Damaskin BB, Kononenko<br />

NA. Elektrokhimiya (Russ J Electrochem) 1987;23:811.<br />

[134] Rubinstein I, Zaltzman B. Surface chemistry and electrochemistry of<br />

membranes. In: Sorensen TS, editor. Dekker; 1999. p. 591.<br />

[135] Sata Toshikatsu, Sata Tomoaki, Yang W. J Membr Sci 2002;206:31.<br />

[136] Kotov VV, Peregonchaya OV, Selemenev VF. Elektrokhimiya (Russ J<br />

Electrochem) 2002;38:1034.<br />

[137] Zolotareva RI, Kotov VV, Romanov MN. Electrochemistry of ion<br />

exchanger. In: Gnusin NP, editor. Kubansky universitet; 1977. [in Rus.].<br />

[138] Berezina NP, Kononenko NA. Elektrokhimiya (Russ J Electrochem)<br />

1982;18:1396.<br />

[139] Berezina NP, Fedorovich NV, Kononenko NA, Komkova EN. Elektrokhimiya<br />

(Russ J Electrochem) 1993;29:1254.<br />

[140] Rubinstein I, Kedem O, Staude E. Desalination 1988;89(2):101.<br />

[141] Ibanez R, Stamatialis DF, Wessling M. J Membr Sci 2004;239:119.<br />

[142] Pivovarov NP. Heterogeneous ion exchange membranes in electrodialysis<br />

processes. Dalnauka; 2001 [in Rus.].<br />

[143] Volodina E, Pismenskaya N, Nikonenko V, Larchet C, Pourcelly G.<br />

J Colloid Interface Sci 2005;285:247.<br />

[144] Belova E, Lopatkova G, Pismenskaya N, Nikonenko V, Larchet C.<br />

Desalination 2006;199:59.<br />

[145] Lopatkova G, Volodina E, Pismenskaya N, Fedotov Yu, Kot D,<br />

Nikonenko V. Elektrokhimiya (Russ J Electrochem) 2006;42:942.<br />

[146] Berezina NP, Shaposhnik VA, Praslov DV, Ivina OP. Zh Phiz Khim<br />

1990;64:2790.<br />

[147] Frumkin AN, Fedorovich NV, Stenina EV. Itogi nauki i tekhniki.<br />

Elektrokhimiya. V<strong>IN</strong>ITI; 1978 [in Rus.].<br />

[148] Fedorovich NV, Stenina EV. Itogi nauki i tekhniki. Elektrokhimiya.<br />

V<strong>IN</strong>ITI; 1981 [in Rus.].<br />

[149] Gnusin NP, Kononenko NA, Parshikov SB. Elektrokhimiya (Russ J<br />

Electrochem) 1994;30:35.<br />

[150] Adhikari B, Majumdar S. Prog Polym Sci 2004;29:699.<br />

[151] Bullen RA, Arnot TC, Lakeman JB, Walsh FC. Biosens Bioelectron<br />

2006;21:2015.<br />

[152] Jurado JR, Colomer MT. Chem Ind 2002;56:264.<br />

[153] Loza NV, Berezina NP, Kononenko NA, Shkirskaya SA. Izvestiya Vuzov<br />

Severo-Kavkazskiy Region; 2006. p. 51. (in Rus.).<br />

[154] Berezina NP, Kubaisy AA-R, Alpatova NM, Andreev VN, Griga EI.<br />

Elektrokhimiya (Russ J Electrochem) 2004;40:333.<br />

[155] Berezina NP, Kubaisy AA-R. Elektrokhimiya (Russ J Electrochem)<br />

2006;42:91.<br />

[156] Berezina NP, Kubaisy AA-R, Timofeev SV, Karpenko LV. J Solid State<br />

Electrochem 2007;11:378.<br />

<strong>ARTICLE</strong> <strong>IN</strong> <strong>PRESS</strong><br />

26 N.P. Berezina et al. / Advances in Colloid and Interface Science xx (2008) xxx–xxx<br />

[157] Kedem O, Katchalsky A. Trans Faraday Soc 1963;59:1918.<br />

[158] Michaeli L, Kedem O. Trans Faraday Soc 1961;57:1185.<br />

[159] Meares P. Trans Faraday Soc 1959;55:1970.<br />

[160] Gardner CR, Ferguson H, Paterson R. Proc. 2nd conf. Appl. Phys.<br />

Chem. — Budapest; 1971. p. 575.<br />

[161] Gnusin NP, Zabolotsky ВИ VI, Nikonenko VV, Meshechkov AI. Zh Phiz<br />

Khimii 1980;54:1518.<br />

[162] Berezina NP, Gnusin NP, Meshechkov AI, Dyomina OA, Kononenko<br />

NA. Theoria i praktika sorbtsyonnych processov (Theory and Practice of<br />

Sorption Processes), vol. 20; 1989. p. 142.<br />

[163] Zabolotsky VI, Nikonenko VV. J Membr Sci 1993;79:181.<br />

[164] Gnusin NP. Elektrokhimiya (Russ J Electrochem) 1996;32:420.<br />

[165] Gnusin NP. Teor Osnovi Khim Tekhnol 2004;38:316.<br />

[166] Gnusin NP, Kononenko NA, Parshikov SB. Elektrokhimiya (Russ J<br />

Electrochem) 1993;29:757.<br />

[167] Kononenko NA, Gnusin NP, Berezina NP, Parshikov SB. Elektrokhimiya<br />

(Russ J Electrochem) 2002 Т. 38, № 8. С. 930.<br />

[168] Gnusin NP, Parshikov SB, Dyomina OA. Elektrokhimiya (Russ J<br />

Electrochem) 1998;34:1316.<br />

[169] Gierke TD, Munn GE, Wilson C. Polym Sci Polym Phys Ed 1981;19:1687.<br />

[170] Gnusin NP, Berezina NP, Dyomina OA, Dvorkina GA. Elektrokhimiya<br />

(Russ J Electrochem) 1997;33:1342.<br />

[171] Hsu WI, Gierke TD. Macromolecules 1982;15:101.<br />

[172] Narebska AR, Wodzki. Angew Makromol Chem 1979;80:105.<br />

[173] Gnusin NP, Dyomina OA, Berezina NP, Kononenko NA. Euromembrane'99.<br />

Book of Abstracts, vol. 2. Elsevier; 1999. p. 527.<br />

[174] Gnusin NP, Grebenyuk VD, Pevnitskaya MV. Elektrokhimiya ionitov.<br />

Nauka; 1972. [in Rus.].<br />

[175] Gnusin NP, Grebenyuk VD. Elektrokhimiya granulirovannykh ionitov.<br />

Naukova dumka; 1972. [in Rus.].<br />

[176] Gnusin NP, Ivina OP. Zhurn Phys Khim (Russ J Phys Chem)<br />

1991;65:2461.<br />

[177] Gnusin NP, Berezina NP, Kononenko NA, Dyomina OA. Zh Phys Khim<br />

(Russ J Phys Chem) 1999;73:1312.<br />

[178] Berezina NP, Kononenko NA, Dyomina OA, Gnusin NP. Vysokomolekularnye<br />

soedineniya (High Molecular Compounds) A., 46; 2004. p. 1071.<br />

[179] Gnusin NP, Berezina NP, Kononenko NA, Demina OA. J Membr Sci<br />

2004;243:301.<br />

[180] Gnusin NP, Berezina NP, Demina OA, Kononenko NA. Teor Osnovi<br />

Khim Tekhnol 2004;38:419.<br />

[181] Artemenko SE, Kardash MM, Berezina NP, Kononenko NA, Klachkova<br />

NYu, Shirokova SV. Khimicheskie Volokna 1997;5:40.<br />

[182] Shishkina SV, Kovyazina LI, Maslennikova IYu, Pechenkina ES.<br />

Elektrokhimiya (Russ J Electrochem) 2002;38:998.<br />

[183] Larchet C, Dammak L, Auclair B, Parchikov S, Nikonenko V. New J<br />

Chem 2004;28:1260.<br />

[184] Larchet C, Auclair B, Nikonenko V. Electrochim Acta 2004;49(11):1711.<br />

Please cite this article as: Berezina NP et al, Characterization of ion-exchange membrane materials: Properties vs structure, Adv Colloid Interface Sci (2008),<br />

doi:10.1016/j.cis.2008.01.002

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!