Sei sulla pagina 1di 47

Review

pubs.acs.org/CR

Selectivity between Oxygen and Chlorine Evolution in the Chlor-


Alkali and Chlorate Processes
Rasmus K. B. Karlsson and Ann Cornell*
Applied Electrochemistry, School of Chemical Science and Engineering, KTH Royal Institute of Technology, SE-100 44 Stockholm,
Sweden

ABSTRACT: Chlorine gas and sodium chlorate are two base chemicals produced
through electrolysis of sodium chloride brine which find uses in many areas of industrial
chemistry. Although the industrial production of these chemicals started over 100 years
ago, there are still factors that limit the energy efficiencies of the processes. This review
focuses on the unwanted production of oxygen gas, which decreases the charge yield by
up to 5%. Understanding the factors that control the rate of oxygen production requires
understanding of both chemical reactions occurring in the electrolyte, as well as surface
reactions occurring on the anodes. The dominant anode material used in chlorate and
chlor-alkali production is the dimensionally stable anode (DSA), Ti coated by a mixed
oxide of RuO2 and TiO2. Although the selectivity for chlorine evolution on DSA is high,
the fundamental reasons for this high selectivity are just now becoming elucidated. This
review summarizes the research, since the early 1900s until today, concerning the
selectivity between chlorine and oxygen evolution in chlorate and chlor-alkali
production. It covers experimental as well as theoretical studies and highlights the relationships between process conditions,
electrolyte composition, the material properties of the anode, and the selectivity for oxygen formation.

CONTENTS 3.2. Extent of Oxygen Evolution from Different


Reactions 3019
1. Introduction 2982 3.3. Connection between Anode Stability and
1.1. Industrial Chlor-Alkali Production 2983 Selectivity 3020
1.2. Industrial Sodium Chlorate Production 2985 4. Conclusions 3021
1.3. Selectivity Concept in Chlor-Alkali and Author Information 3021
Sodium Chlorate Production 2986 Corresponding Author 3021
2. Selectivity for Oxygen and Chlorine Evolution in Notes 3021
Chlor-Alkali and Chlorate Processes 2986 Biographies 3021
2.1. Studies on Oxygen Formation during De- Acknowledgments 3022
composition of Hypochlorous Acid Species 2986 Nomenclature 3022
2.1.1. Uncatalyzed Decomposition 2986 References 3022
2.1.2. Catalyzed Decomposition 2987
2.2. Studies on Selectivity during Electrolysis 2988
2.2.1. Academic Research from 1900 to 1969 2988
2.2.2. Academic Research during the 1970s 2990 1. INTRODUCTION
2.2.3. Research during the 1980s 2991
Chlorine gas and sodium chlorate are base chemicals, with yearly
2.2.4. Research during the 1990s 2997
production rates of 50−60 million tonnes1,2 and over three
2.2.5. Research during the 2000s 3000
million tonnes,3 respectively. Chlorine gas is used in a wide
2.3. Electrochemical Studies Concerning the
variety of applications, including in production of construction
Relationship between Selectivity and Stabil-
materials such as polyvinyl chloride (PVC), in organic synthesis,
ity 3009
in metallurgy, and in water treatment.4 Sodium chlorate finds its
3. Discussion 3015
use almost exclusively for bleaching of pulp and paper products.3
3.1. Factors Affecting the Selectivity between
Both products are produced predominantly through electrolysis
Chlorine and Oxygen Evolution in the
of sodium chloride brine. Although the charge yields of both the
Production of Chlorine and Chlorate 3015
chlor-alkali process, in which chlorine gas and sodium hydroxide
3.1.1. Influence of Process Conditions 3015
are produced, and chlorate processes are quite high, ≥ 95%,5,6 the
3.1.2. Influence of Additions and Contami-
fact that both chlorine gas and chlorate are bulk chemicals,
nants in the Electrolyte 3017
3.1.3. Influence of the Anode Structure and
Composition 3017 Received: July 6, 2015
Published: February 16, 2016

© 2016 American Chemical Society 2982 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

produced in large quantities, means that losses of even a few ature, affect reactions both in the bulk electrolyte and on the
percent are costly. The yearly electricity consumption in chlor- anode surface. This review therefore discusses all three aspects:
alkali production alone is about 150TWh,1 which makes it clear electrolyte composition, anode composition, and the process
that improvements in energy efficiency of these processes could conditions. The focus of this review is on the selectivity issue in
yield significant reductions in world energy consumption. The general, without limitation with regard to the anode material
main side reaction in industrial electrolytic cells is the production used. Since there are connections between some mechanisms of
of oxygen. In the chlorate process, the unwanted oxygen electrode decomposition and oxygen production, the stability of
combines with hydrogen in the cell gas and may form explosive DSA will also be discussed.
gas mixtures, which also makes oxygen a safety hazard. Further The following section of the introduction constitutes a brief
study of the factors deciding the selectivity between chlorine or overview of the chlor-alkali and chlorate processes, as well as the
chlorate and oxygen production in industrial chlor-alkali and most important and well-known main and side reactions. It gives
chlorate cells is therefore warranted. However, there is no review the necessary basic information to describe effects of various
of the scientific literature that focuses specifically on the process, electrolyte, and anode properties on the selectivity. The
selectivity issue. This review aims to bridge that gap. selectivity in the chlor-alkali and chlorate processes is not a
The electrolytic production of chlor-alkali and sodium chlorate simple concept, as it involves both direct formation of oxygen on
is performed in either divided, for production of chlor-alkali, or the anode as well as oxygen resulting from bulk reactions.
undivided, for production of sodium chlorate, electrochemical Therefore, the selectivity concept as applied for industrial chlor-
cells using, for example, steel or activated Ni cathodes (Ni alkali and sodium chlorate production, as well as some of the
cathodes are only used in chlor-alkali production) and methods that have been used to explore it, will be described in
dimensionally stable anodes (DSA, dimensionally stable anode, section 1.3. Afterward, in section 2, the literature relevant to the
is a trademark of Industrie De Nora S.p.A., Italy.).1,7 DSA are selectivity is described, in chronological order from the early
mixed ruthenium−titanium oxide (RTO) coatings of rutile RuO2 1900s until today. The literature is described in three main
and TiO2 deposited on Ti. The usage of DSA to describe anodes sections, covering research on oxygen-evolving reactions relevant
in the present review is primarily used to refer to anodes of the to the electrolyte processes, electrochemical studies (both
typical industrial “Beer 2” (30% RuO2 and 70% TiO2 coated on experimental and theoretical), and studies on the link between
Ti) formulation.8 Commercial DSA electrodes most often stability and selectivity. Finally, a discussion of the results and
contain one or more other additional dopant materials. RTO is trends found in literature is given, as well as some conclusions
used in the present review as a more general term for anodes with regarding future research needs.
coatings containing varying amounts of RuO2 and TiO2. In these 1.1. Industrial Chlor-Alkali Production
coatings, the RuO2 and TiO2 components form solid solutions, Chlorine gas and sodium hydroxide are produced in chlor-alkali
where RuIV and TiIV are part of the same rutile lattice.9−11 While production according to the following overall reaction1
rutile RuO2 has a high electronic conductivity,12 pure rutile TiO2
is a semiconductor with a band gap of 3 eV.13 Nevertheless, the 2NaCl + 2H 2O → Cl 2 + H 2 + 2NaOH (1)
mixed oxide has a high electronic conductivity, enabling its use as
an electrode, as the doping with RuIV introduces new electronic The electrochemical formation of chlorine gas at the anode
states in the region of the TiO2 band gap.14,15 The preparation (DSA) occurs according to the following reaction
and usage of dimensionally stable anodes was patented by Beer in 2Cl− → Cl 2 + 2e− (2)
a series of patents in the 1960s (Britain) and 1970s (United
States).8,16−23 The discovery of DSA has been called “one of the with an associated equilibrium potential of E° = 1.3583 V versus
greatest technological breakthroughs of the past 50 years of SHE, where the standard potentials used are those of Bard et
electrochemistry”.24 Since then, the usage of DSAs in these al.,37 unless otherwise indicated. The evolved chlorine gas is
processes has led to significant energy savings due to their lower collected and processed in a number of steps. As chlorine is being
potentials at industrial current densities.24,25 However, as the evolved at the anode, hydrogen (except in the mercury cell
name implies, their most important advantage over previous process) forms at the cathode
graphite electrodes is their stability, with modern DSAs being 2H 2O + 2e− → H 2 + 2OH− (3)
able to operate at industrial current densities for more than 10
years.24 The literature concerning DSAs and RTOs in general is a reaction with E° = −0.8277 V. The anode compartment is kept
summarized in several reviews.23−33 The recent reviews of Over separated from the cathode compartment, and the way in which
concerning the chemistry and catalysis of pure RuO212 and they are separated constitutes the main characteristic of the
concerning correlations and contrasts between liquid phase different processes used to produce chlor-alkali. The mercury cell
electrochemical and gas phase chemical reactions on RuO2 and process, the diaphragm process, and the membrane process
related metallic oxide materials,34 as well as the review of dominate industrial production of chlor-alkali. In a membrane
Diebold35 concerning the surface chemistry of TiO2, should also cell, a cation-selective membrane separates the anode and
be mentioned. cathode compartments, whereas a microporous diaphragm is
The oxygen evolution side reaction (OER) in chlor-alkali and used in the diaphragm process. The first industrial process was
sodium chlorate production is connected to catalytic processes the mercury cell process, in which a mercury amalgam is
ongoing both in the electrolyte and on the anode surface, where produced at the cathode instead of hydrogen. The amalgam is
oxygen might be evolved both electrochemically and chemically then later decomposed in a separate reactor to yield hydrogen gas
through water or hypochlorous acid (or hypochlorite) and sodium hydroxide. Today, the mercury cell and diaphragm
decomposition. Therefore, to get a complete picture of the processes are being replaced by the membrane process for both
selectivity issue in these industrial processes, compositions of environmental and technical reasons.38,39 By the end of year
both the electrolyte and the anode are important. Furthermore, 2013, the total installed chlorine capacity in Europe consisted of
other process conditions, such as current density and temper- 59% membrane technology, 25% mercury cells, and 14%
2983 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

diaphragm cells.39 The usage of mercury cells in the European Table 2. Some Important Dissolved and Gaseous Chlorine
Union is scheduled to be discontinued in 2017.39 Typical Species and Their Cl Oxidation States.42
conditions applied in the membrane process are shown in Table
species name Cl oxidation state
1, and more detailed descriptions of the three technologies can be
Cl− (aq) chloride (ion) −1
Table 1. Typical Conditions for Chlor-Alkali Production in Cl−3 (aq) trichloride ion −1, 0
Membrane Processes1,2,36 HCl (g) or (aq) hydrochloric acid −1
Cl2 (g) or (aq) chlorine gas 0
cell voltage (V) 2.4−2.7 HOCl (aq) hypochlorous acid +1
current density (kA m−2) 1.5−7 [or chloric (I) acid]
temperature (°C) 90 ClO− (aq) hypochlorite (ion) +1
NaCl concentration in the anolyte (g dm−3) 200 [or chloric (I) ion]
anolyte pH 2−4 Cl2O (g) chlorine monoxide +1
NaOH concentration in the catholyte (wt %) 32 HClO2 (aq) chlorous acid +3
ClO−2 (aq) chlorite (ion) +3
found in, for example, Wendt et al.40 or Schmittinger et al.1 A ClO2 (g) chlorine dioxide +4
small percentage of the total chlorine production is based on ClO−3 (aq) chlorate (ion) +5
electrolysis of 17 wt % HCl in either diaphragm or membrane ClO−4 (aq) perchlorate (ion) +7
cells.1 The energy cost for the production is usually given in
terms of kWh t−1 NaOH produced, with 886 kg Cl2 being
produced per tonne of NaOH. In membrane cells, the specific
electrical energy requirement is approximately 2100 kWh t−1
NaOH.1
Dissolved chlorine may hydrolyze according to the dynamic
equilibrium in reaction 4. Formed hypochlorous acid is involved
in another equilibrium with water, reaction 5. The impact of the
pH value on the relative concentrations of Cl2, HOCl, and ClO−
is significant, as is indicated in Figure 1.
Cl 2 + H 2O ⇌ H+ + HOCl + Cl− (4)

HOCl + H 2O ⇌ ClO− + H3O+ (5)


Hypochlorous acid and hypochlorite react to form chlorate, Figure 2. Pourbaix diagram for aqueous species in the chlorine-water
ClO−3 , according to reaction 6. system, indicating the stable equilibria at 25 °C and unit activity of the
dissolved species. Figure prepared using pymatgen45−47 and thermody-
2HOCl + ClO− → ClO−3 + 2H+ + 2Cl− (6) namic data from Wagman et al.48

Figure 1. Percentages of active chlorine species Cl2, HOCl, and ClO− as


a function of anolyte pH at 90 °C and 200 g NaCl dm−3. The
percentages are calculated using equilibrium constants for Cl2 hydrolysis Figure 3. Pourbaix diagram for the aqueous metastable chlorates
and HOCl deprotonation. Reprinted with permission from ref 41. (excluding ClO−4 ) in the chlorine-water system, indicating the equilibria
Copyright 1990 Springer Science and Business Media. at 25 °C and unit activity of the dissolved species. Figure prepared using
pymatgen45−47 and thermodynamic data from Wagman et al.48
Cl−, HOCl, ClO−, and ClO−3 are the main species of interest
during chlor-alkali and chlorate production, but the chlorine- which is in equilibrium with Cl2 (aq) under chlor-alkali
water equilibrium in aqueous solution is significantly more conditions (chloride activities between 2 and 6),44 is not seen
complex. Table 2 shows some of the known dissolved and in the figures, as its window of stability is very small.
gaseous species that might occur during electrolysis.42,43 The In addition to chlorate ions, hypochlorite (or hypochlorous
stability of these species at different electrode potentials and pH acid) may also form oxygen as will be described in the next
values is indicated in Figures 2 (stable species), 3 (metastable section. Both chlorate ions and oxygen gas are highly undesired
chlorate species), 4 (metastable chlorite species), and 5 in chlor-alkali cells, and thus, the formation of the reactant
(metastable hypochlorite species). The trichloride ion, Cl−3 , hypochlorite species should be suppressed. In the membrane and
2984 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

NaCl + 3H 2O → NaClO3 + 3H 2 (7)


The energy requirement in kWh per tonne sodium chlorate
produced in the industrial process can be estimated using (with a
typical cell voltage of 3.1 V)51
Pe = 1511 Ah t−1 × 3.1V ≈ 4700 kWh t−1 (8)
The same electrode reactions that occur in the chlor-alkali
process, reactions 2 and 3, are also believed to be the first steps in
chlorate production.51 Dimensionally stable anodes and
cathodes of either low carbon steel or titanium are used in the
electrolysis cell.6 The chlorate cells are undivided, and the bulk
electrolyte pH is 6−7. However, the ongoing anode and cathode
reactions make the pH at the anode more acidic, while the pH at
Figure 4. Pourbaix diagram for the aqueous metastable chlorites the cathode is alkaline. Thus, a pH gradient is set up across the
(excluding ClO−3 and ClO−4 ) in the chlorine-water system, indicating the cell, and this means that the hydroxide ions that form in the
equilibria at 25 °C and unit activity of the dissolved species. Figure cathode reaction 3 can hydrolyze the chlorine that forms in the
prepared using pymatgen45−47 and thermodynamic data from Wagman anode reaction 2.6,7,51,52 This gives the following dynamic
et al.48 equilibrium
Cl 2 + 2OH− ⇌ ClO− + H 2O + Cl− (9)
Alternatively, water can hydrolyze the chlorine gas according to
reaction 4.6
Hypochlorite is formed through reactions 4 and 5, or through
reaction 9, and then chlorate is formed according to reaction 6.51
This reaction is a purely homogeneous chemical reaction, which
has a maximum rate around pH 6.1−6.4 under industrial
conditions and 80 to 90 °C.7 The pH that is used in the chlorate
process is chosen to prevent chlorine gas formation, which
increases in rate at lower pH, and oxygen gas formation, which
increases in rate at higher pH. Chlorate cells usually have a large
volume tank through which the reaction solution flows to allow
reaction 6 to reach high conversion.6,40,53 The process
Figure 5. Pourbaix diagram for the aqueous metastable hypochlorites temperature is kept high to increase the rate of reaction 6
(excluding ClO−3 , ClO−4 , HClO2, and ClO−2 ) in the chlorine-water (enabling the usage of a smaller reaction vessel) and of the
system, indicating the equilibria at 25 °C and unit activity of the electrode reactions (lowering the cell voltage).6,53−56 However,
dissolved species. Figure prepared using pymatgen45−47 and thermody- the rate of parasitic oxygen evolution also increases with
namic data from Wagman et al.48 increasing temperature, meaning that the temperature chosen
is a balance between minimizing the rate of oxygen evolution and
maximizing the reaction rates of the desired reactions.57
diaphragm cells, small amounts of hydroxide ions may pass A number of anodic reactions limit the charge yield,
through the separator, dividing the acidic anolyte and the alkaline particularly for chlorate production but also for chlor-alkali
catholyte, and increase the pH close to the separator in the anode production. One of the more well-known is the following anodic
chamber, with subsequent oxygen and chlorate generation. side-reaction in which both oxygen and chlorate are formed:58,59
Careful acidification of the anolyte can counteract the increase in 6ClO− + 3H 2O → 2ClO−3 + 4Cl− + 6H+ + 1.5O2 + 6e−
pH related to the hydroxide leakage, resulting in oxygen levels in
(10)
the chlorine gas of <0.5%. Oxygen levels in the chlorine gas are 54
typically up to 4% in a diaphragm cell49 and 1.5−2% in with E° = 1.14 V. The reaction is generally called anodic
membrane cells with no anolyte acidification.49 Mercury cells, chlorate formation and must necessarily be a sum of several
with no separator but where the alkali is formed in a secondary substeps.
reactor, do not have this problem and the chlorine gas contains Apart from reaction 10, the oxidation of water, hydroxide, and
<0.5% oxygen.49 Bilayered membranes, with perfluorosulfonate chlorate can also occur.
facing the anode and perfluorocarboxylate facing the cathode 2H 2O → O2 + 4H+ + 4e−, E° = 1.229 V (11)
side, are state-of-the-art today in membrane cells.1 The loss in
charge yield for alkali is 3−7% due to the migration of hydroxide 1
ions through the membrane, while the loss of chloride ions to the 2OH− → O2 + H 2O + 2e−, E° = 0.401 V
2 (12)
catholyte is extremely low. About 0.14 kg ClO−3 produced per
tonne NaOH is to be compared to 3.5 to 5.2 kg ClO−3 per tonne ClO−3 + H 2O → ClO−4 + 2H+ + 2e−, E° = 1.189 V
NaOH in the absence of acidification.50 (13)
1.2. Industrial Sodium Chlorate Production By comparing the standard potentials of reactions 11 and 2
Sodium chlorate is also produced using electrolysis of sodium (chlorine evolution), it is seen that the OER is thermodynami-
chloride solutions. The following overall reaction describes the cally favored versus the chlorine evolution reaction. However,
formation: high chlorine evolution reaction (ClER) selectivity is still
2985 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

obtained on modern electrocatalysts due to the significantly 1.3. Selectivity Concept in Chlor-Alkali and Sodium Chlorate
more favorable kinetics of the ClER. Oxidation of water Production
dominates if the anode potential is below that required for The desired reactions in chlor-alkali and sodium chlorate
chlorine formation, and if there is no hypochlorite in the production is first the anodic ClER (reaction 2, the first and
solution.54,60 The reaction is also dependent on the anode only desired anodic reaction in chlor-alkali production), followed
material, chloride concentration, and the pH of the electrolyte by, in the case of chlorate production, chemical conversion of Cl2
and can largely be controlled by adjusting these factors.7 When into ClO−3 (reactions 9 or 4 followed by reaction 6). In other
discussing the effects of pH, it is necessary to keep in mind that words, in chlor-alkali production, the current efficiency for
the conditions in the bulk of the electrolyte might differ chlorine formation is the primary measure of selectivity, while the
significantly from those at the electrode surface, where the charge yield for chlorate production is the measure of selectivity
electrochemical reactions occur.44,53,61 Due to the parasitic in the chlorate process. The selectivity between anodic chlorine
formation of oxygen gas (reaction 11) and the hydrolysis of Cl2 evolution (reaction 2) and oxygen evolution (reaction 11) on the
(reaction 4, less important in chlor-alkali production), the pH anode surface is related to the intrinsic selectivity of the anode
close to the anode surface will be more acidic than in the material used. Here, the properties (surface structure and
electrolyte bulk.44 For this reason, reaction 12 is very minor. composition) of the anode are the deciding factors, and its
importance is reflected in the large number of studies that have
Additionally, as described above, migration of OH− through the
been focused on the relationship between the anode and the
separator in chlor-alkali cells results in locally increased pH
selectivity. However, especially in the case of the chlorate
values in the anolyte close to the membrane. In turn, this process, it is only a part of the overall selectivity issue, for which
increased pH results in local formation of hypochlorous acid and secondary reactions in the bulk (decomposition of hypochlorous
hypochlorite, enabling subsequent homogeneous chlorate acid species, reaction 14) and reactions involving decomposition
formation and possibly also decomposition of hypochlorous of hypochlorous acid species on the anode are of key importance.
acid species to form oxygen. Chlorate oxidation, by reaction 13, The study of the selectivity in these processes thus requires
will occur under conditions of high chlorate concentration and methods that can account for both anode reactions and bulk
low chloride concentration, and although DSA yields very small reactions and reaction products in both liquid and gas phase. The
amounts of perchlorate, significant concentrations may accumu- anode reactions have mostly been explored using conventional
late over time in closed mode operation.5,7 electrochemical measurements (e.g., voltammetry and polar-
Apart from reactions occurring electrochemically at the ization curves), but theoretical methods have also been applied to
electrodes, there is also the possibility that hypochlorite is study the intrinsic selectivity of anodes.63−65 The composition of
decomposed in the bulk of the electrolyte, or heterogeneously on evolved gases are usually determined using, for example, gas
the electrodes, according to reaction 14:6,54 chromatography (GC)66−69 or absorption of O2 in, for example,
pyrogallol.36,41,54,70−74 Considering that decreases in anodic
2HOCl → 2HCl + O2 (14) selectivity primarily occur either through direct formation of O2
on the anode, or by formation of O2 by decomposition of HOCl
Reaction 14 is slow62 but may be catalyzed by metal ions and or OCl− in the electrolyte, it is reasonable to quantify the anodic
light. The literature on catalysis of this reaction in hypochlorite current efficiency (the percentage of the anodic current used for
solutions is discussed in section 2.1 of the present review. production of the desired products) based on oxygen in the
The reactions above dictate the design and operating evolved gases. Since the 1990s, differential electrochemical mass
conditions used in chlorate production. To maximize the spectroscopy (DEMS) has been used in several studies.75−84
chemical formation rate of chlorate (reaction 6), the volume of This method enables a highly sensitive determination of reaction
electrolyte in the electrolysis compartment or in the interelec- products formed on an anode. In the electrolyte, the
trode gap is minimized, and the formation of chlorate then concentrations of hypochlorite, chloride, and chlorate can be
mostly occurs in a large either internal or external reactor.7 The determined using titration methods.
electrolyte is circulated, either by gas lift in the electrolysis cell or
by pumping, and the temperature of the electrolyte is kept 2. SELECTIVITY FOR OXYGEN AND CHLORINE
constant by use of heat exchangers.6 The composition, pH, and EVOLUTION IN CHLOR-ALKALI AND CHLORATE
temperature of the electrolyte as well as the current density and PROCESSES
cell voltage need to be optimized for high current efficiency. 2.1. Studies on Oxygen Formation during Decomposition of
Some typical operating conditions for sodium chlorate cells are Hypochlorous Acid Species
shown in Table 3. Before considering the selectivity in the electrochemical
environment, the literature concerning reactions in the bulk
Table 3. Typical Operating Conditions in Sodium Chlorate will be presented. Both the uncatalyzed and catalyzed
Production6,7,51 decomposition of hypochlorous acid or hypochlorite to form
oxygen, reaction 14, will be discussed in this section.
cell voltage (V) 2.9−3.7 2.1.1. Uncatalyzed Decomposition. The uncatalyzed
current density (kA m−2) 1.5−4 decomposition of hypochlorous acid species to form oxygen
temperature (°C) 65−90 has been discussed less than the main reaction (6) forming
NaCl concentration (g dm−3) 70−150 sodium chlorate. It appears that this reaction has long been
NaClO3 concentration (g dm−3) 450−650 known,85,86 although which is the earliest discussion of the
NaOCl concentration (g dm−3) 1−5 reaction is unclear. Indeed, the reaction of
Na2Cr2O7 concentration (g dm−3) 1−6
electrolyte pH 5.5−7 2NaOCl ⇌ 2NaCl + O2 (15)

2986 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

occurring together with homogeneous sodium chlorate for- Slow rates of O2 production were noted for Sn, Cu, and As. Al
mation, is mentioned already in 1897, in the work of Bhaduri87 and Mg were instead found to yield hydrogen production. Sb, Pb,
who studied the decomposition of NaOCl in darkness at 100 °C. Zn, and Cd were found to be inactive. The effect of sunlight on
Returning to research from the 20th century, it seems that hypochlorite solutions was also studied, and a correlation
Lister88 is the first to consider the side reaction in much detail, in between sunlight and formation of small amounts of oxygen
a paper from 1952. It was suggested that the oxygen evolution was noted.
reaction is first order in HOCl but independent of the OCl− In 1914, Hofmann and Ritter95 compared the effects of a
concentration. The first-order rate constant was found to be 4.65 number of oxides on calcium hypochlorite solutions and sodium
× 10−4 min−1 at 30 °C, increasing to 12.9 × 10−4 min−1 at 40 °C. hypochlorite solutions. They found oxides of Co, Ni, and Ir to be
The associated activation energy for oxygen formation was 80 kJ active for the decomposition of both solutions. Salts of Ru and Rh
mol−1. As a comparison, the rate constant for the slow first step in were found to decompose the bleaching powder solution but not
the formation of chlorate, found to occur in two steps via ClO−2 , the sodium chlorite solution. Oxides (the specific type of each
was 0.95 M−1 s−1 at 40 °C. Two more papers were later oxide was not mentioned) of Cr, Fe, Mn, U, W, Pd, Os, Tl, and V
published86,89 by the same author and a co-worker. Both of these were found to be inactive in both solutions.
papers were focused on attempts to completely remove any In 1923, Howell93 studied the effect of Co peroxides (prepared
impurities, such as of copper, that might catalyze the formation of by precipitating CoSO4 with NaOH and NaOCl) on sodium
oxygen. Indeed, in the later paper of Lister and Petterson,89 it was hypochlorite solutions and found that they catalyze the
concluded that the uncatalyzed decomposition of hypochlorite is decomposition. Addition of NaCl was found to accelerate the
of second order in OCl−, with a rate constant of 7.5 × 10−6 M−1 decomposition.
min−1 at an ionic strength of 3.5 and 60 °C. On the basis of an In 1926, Chirnoaga96 again studied the effect of Co peroxide,
evaluation of activation energies, and comparison with the Ni peroxide, CuO, Fe2O3, MnO2, HgO, and alumina on sodium
chlorate formation reaction, it was suggested that the two hypochlorite solutions. Co and Ni peroxides (both prepared
reactions had activated complexes with the same stoichiometry, using precipitation of CoCl3 or NiCl2 with NaOH and NaOCl)
90
Cl2O2−2 , but with different arrangements. In 1994, Church again had the highest activity for decomposition, while activities of
considered the uncatalyzed oxygen evolution side reaction CuO, Fe2O3, MnO2, or HgO were lower, and alumina was
during hypochlorite decomposition, finding results in line with inactive. Alumina was found to promote the action of CoII
those of Lister and Petterson.89 In 1999, Adam and Gordon91 hydroxide.
measured a second-order rate constant in OCl− of 8.9 × 10−10 In two articles published in 1927 and 1930,97,98 Lewis studied
M−1 s−1 for uncatalyzed hypochlorite decomposition. Sandin et the catalyzed decomposition of NaClO by solutions of CuSO4,
al.92 studied the decomposition of hypochlorite in 2015, FeSO4, CoSO4, and suspensions of cobalt peroxide (prepared in
determining the rate dependence on pH in dilute hypochlorite a similar way as in the studies of Howell93 and Chirnoaga96),
solutions. The study combined mass spectrometry with ionic CuO, and Fe2O3. In the first article, it was shown that the
chromatography to study both gas phase and liquid phase peroxide suspension and sulfate solution of Co were most active
concentrations during decomposition. It was found that the for hypochlorite decomposition, while the corresponding CuO
uncatalyzed O2 formation reaction has a pH dependence that is suspensions and solutions were less active. Data for the iron
very similar to that of chlorate formation, with a maximum rate at oxide suspension alone was not reported. Instead, a mixed
about pH 6.5. This could further support the suggestion of Lister suspension of Fe2O3 and CuO was used, which was reported to
and Petterson,89 that the two pathways of chlorate formation be more active than the CuII suspension and solution. This was
(main reaction) and oxygen formation (side reaction) share an said to be due to a promoter effect by Fe2O3. In the second
intermediate. The oxygen evolution rate was found to be third article,98 a promoting effect of magnesium hydroxide on the
order in total hypochlorite concentration (cHOCl + cOCl−) at pH hypochlorite decomposition activity of CuO was reported. It was
6.5, a higher order than that found previously in the alkaline argued that the promotional effect was caused by Mg hydroxide
range.89,91 The third-order rate constant for uncatalyzed oxygen particles preventing the CuO particles from agglomerating.
formation by hypochlorite decomposition in this pH range was In 1952, as a part of a study on the properties of Ru in its + VI,
found to be 0.046 M−2 s−1, compared with that for chlorate + VII, and + VIII oxidation states published, Connick and
formation of about 0.73 M−2 s−1. Hurley99 presented results indicating a catalytic effect of the

2.1.2. Catalyzed Decomposition. Compounds that are ruthenate ion RuO2− 4 on decomposition of ClO
able to decompose hypochlorous acid species have been known
H 2O + 2RuO24 − + ClO− ⇌ 2RuO−4 + Cl− + 2OH−
ever since the early 19th century. Howell93 lists six studies from
the 1800s and notes that earlier researchers have found that silver (16)
oxide, powdered pyrolusite (a MnO2-containing mineral), oxides RuO2−
4 was then thought to be reformed through a reaction with
of Co, Ni, Cu, and Fe, and barium peroxide are able to, OH−:
catalytically or otherwise, decompose hypochlorous acid species 1
in solution. White94 quotes a study from 1834 in which it was 2RuO−4 + 2OH− ⇌ 2RuO24 − + O2 (g) + H 2O
2 (17)
found that irradiance could decompose hypochlorous acid
100
species to form oxygen. The same study also found that both In 1954, Patel and Mankad presented experimental
Sn and Cu would be converted to oxychloride form, with measurements on oxygen evolution due to hypochlorite
concomitant production of small amounts of oxygen. decomposition and studied the effect of CoII, NiII, and FeII
In 1903, White94 studied the effect of metallic Fe, Sn, Cu, As, sulfates and chlorides, as well as of CoII and NiII nitrates and ferric
Ni, Co, Al, Mg, Sb, Pb, Zn, and Cd on alkaline bleaching powder ammonium citrate on the decomposition rate. The rate of oxygen
(calcium hypochlorite) solutions at “ordinary temperatures”. Co, evolution was not determined directly and was instead found by
Ni, and Fe were found to give high production rates of oxygen. thiosulfate titration after 3 h of reaction to determine amount of
These metals were all found to be converted into hydroxide form. reacted hypochlorite. This means that the formation of chlorate
2987 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

due to decomposition of HOCl was disregarded in the analysis. detected, whereas no chlorate was formed at the higher pH
In any case, CoII sulfate, chloride, and nitrate were found to values. A first-order rate constant for oxygen evolution by
catalyze the decomposition, while the analogous FeII and NiII decomposition of NaOCl was calculated and found to exhibit a
species were either completely inactive (sulfates) or only maximum at around pH 5. This calculated rate constant agreed
increased the decomposition rate slightly versus the experiment with experimental data for pH > 7, below which the production
conducted without catalyst. Ferric ammonium citrate also of chlorine gas made comparisons impossible.
significantly increased the decomposition rate of HOCl. In 2015, Sandin et al.92 reported results of the catalyzed
In two papers by Ayres and Booth101,102 published in 1955, the decomposition of hypochlorite. The pH value was controlled
decomposition of 2 M NaCl-containing hypochlorite solutions during the experiments, allowing the effect of pH to be
by dissolved Ir metal, in the form of Na2IrCl6, was studied. The decoupled from the effect of the catalysts. The potential catalysts,
effect of initial pH on the catalyzed decomposition was reported which were tested at pH = 6.5, were the salts AgCl, CeCl3, CoCl2,
extensively in the first paper. In the second paper, the kinetics of FeCl3, IrCl3, and RuCl3, and the oxides Fe3O4, Al2O3, Na2Cr2O7,
the catalyzed decomposition was investigated. The reaction rate Na2MoO4, and RuO2. IrCl3, Na2Cr2O7, and RuO2 were also
was found to be first order with respect to hypochlorite tested at pH ≈ 11. Of these, only IrCl3 and CoCl2 were found to
concentration. Rate constants at temperatures between 20 and catalyze the formation of oxygen. Additionally, IrCl3 was found to
40 °C were found, increasing from 3.92 × 10−4 to 2.35 × 10−3 s−1, catalyze the formation of sodium chlorate.
and an energy of activation of 69 kJ mol−1 was calculated. As 2.2. Studies on Selectivity during Electrolysis
active catalysts only need to be present in small quantities, it was
problematic that the Ir metal that was used as the raw material to 2.2.1. Academic Research from 1900 to 1969. Chlorine
prepare Na2IrCl6 in the studies101,102 contained trace amounts of gas was first produced by Scheele in 1774, using chemical
several other metals (Pt, Rh, Cu, Ag, Mg, and Ca). methods. Industrial electrolytic chlorine production was first
In 1956, Lister103 reported the catalytic effect of salts of MnII, performed in Germany in 1890,107 when both the diaphragm and
FeIII, CoII, NiII, and CuII on the decomposition of sodium mercury cell technologies were introduced.1 The earliest
hypochlorite. Apart from FeIII, which was added as a nitrate salt, research concerning chlorine production was summarized long
chloride salts were used. CoII, NiII, and CuII were found to ago by Kershaw,108 Müller,109 and Baldwin.107 Industrial chlorate
catalyze the formation of oxygen, while MnII and FeIII were production has been performed ever since 1886, with research
inactive. None of the materials catalyzed the formation of starting even earlier than that, as was reviewed by Ibl and Vogt.62
chlorate. Reaction orders and activation energies were Focusing presently on the chlorate production process, it is
well-known that, among others, Oettel110−112 and Foerster and
determined. In the case of NiII and CoII, the reaction order in
co-workers58,113−124 studied the fundamentals of chlorate
hypochlorite concentration was found to be zero, while in the
formation during electrolysis of chloride solutions at the turn
case of CuII, the reaction was first order in hypochlorite
of the 20th century. This work might be traced to early German
concentration. For both NiII, CoII, and CuII, the reaction rate was
investigations into industrial-scale chlor-alkali electrolysis for
proportional to the added amount of catalyst.
production of, for example, caustic soda, which is said to have
In 1977, Gray et al.104 reported the catalytic effect of CuII
started in 1884.108 This early research found that formation of
hydroxide complexes on the decomposition of hypochlorite and
Cl2 is the first step in production of sodium chlorate, since a solid
hypobromide in alkaline solution at 25 °C. Since only high
product of Cl2, chlorine hydrate, is formed at the anode if
NaOH concentrations of above 2 M were considered, the results
electrolysis is performed at low temperature.58 It was also found
are less relevant for industrial chlorate production.
that hypochlorite is the intermediate compound in the
In the same year, Hamano and Mori105 presented results on
production of chlorate from chlorides and that the formation
the catalytic effect of CuO, NiO, and Co3O4 on NaOCl and H2O2
of chlorate probably occurs through reaction 6.114 Furthermore,
decomposition. The catalyst materials were calcined at three
Oettel110 and Foerster et al.114 found that the production of
different temperatures (300, 500, and 700 °C), and Brunauer−
chlorate occurs not only through reaction 6 but also through
Emmett−Teller measurements were performed to enable the
anodic oxidation of hypochlorite. Foerster and Müller also tried
activity of the different materials to be normalized by the specific
to find the overall reaction formula for anodic chlorate formation,
surface area. It was found that the reaction rate constant for
identifying eq 10 as the most fitting with their experimental
NaOCl decomposition, normalized by the specific area, was
data.125 Both reactions 6 and 10 are referred to as Foerster’s
highest for NiO and lowest for CuO. Whether the NiO catalyst
reactions in the literature. During this time, it was also suggested
was stable under the reaction conditions, or was oxidized to a
that both reaction 10 and direct oxidation of ClO−,
higher oxide, does not appear to have been studied.
In 1994, Church90 reexamined the catalytic effect of CuII on ClO− + 2O → ClO−3 (18)
the decomposition of NaClO solutions. Rate constants for both
the uncatalyzed and catalyzed decomposition of NaClO were the so-called primary chlorate formation, with no accompanying
determined. The kinetics for the catalyzed decomposition was chloride formation, occurred simultaneously in alkaline solution,
found to be well-described as first order in both CuII and ClO− both reactions forming chlorate (the reaction is written as given
concentration. by Foerster59). This early research was conducted mainly using
In 1998, Hamano et al.106 presented results of the pH effect on polished Pt electrodes, but experiments were also done with
the catalytic decomposition of NaOCl solutions by Co3O4 at 25 graphite electrodes,123,126 rough Pt electrodes,121 and Pt−Ir alloy
°C. In the absence of a catalyst, no oxygen was evolved at pH electrodes.117,123 Foerster123 recommended the Pt−Ir electrode,
values in the range of 7−10, in opposition to what Lister and rather than graphite, in chlorate and chlor-alkali production,
Petterson,89 Church,90 and Sandin et al.92 had found. However, based on its superior stability. These early studies were often
when cobalt oxide was added, oxygen was evolved even at pH 12. careful and detailed studies of the chemistry, often including
Increased decomposition rates were noted as the pH was measurements of both evolved oxygen gas and liquid phase
decreased down to 7. At pH 7−9, both O2 and chlorate were hypochlorite, chlorate, and perchlorate. Perchlorate formation
2988 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

was studied in detail by Oechsli.127 Furthermore, the effects of formation through an electrochemical oxidation of hypochlorite.
temperature, pH,121 chloride concentration,121,128 current Oxygen evolution during electrolytic chlorate formation is, in the
density, and current concentration129 (the ratio between the model of de Valera,133 a consequence of decomposition of H2O2
total current and the volume of the electrolyzer cell) were formed through oxidation of OH−.
reported in some detail in several of these publications. Even at Hammar and Wranglén54 examined the oxygen evolution side
this time, the possibility that oxygen produced during chlorate reaction in chlorate production in an early experimental study
formation might not only be the result of anodic chlorate from 1964, in which graphite anodes were used. The effects of
formation (reaction 10) or water oxidation but also due to current density, temperature, concentrations of chloride,
oxidation of HOCl or OCl− was discussed.116,121 Much of this hypochlorite, and chlorate, pH, and cathode type (steel, iron,
early research is summarized in English by Foerster59 and Ibl and or graphite) on both oxygen evolution and on cathodic losses
Vogt.62 were reported.
Building on these early German studies, Knibbs and It was found that the oxidation loss (determined as the
Palfreeman130 published an article in 1920 concerning the percentage of oxygen in the evolved gas) increased linearly with
chlorate and perchlorate formation reactions in industrial cells hypochlorite concentration from 0 to 250 mM (see Figure 6).
using platinum electrodes. The main part of the work was
focused on finding the proper process parameters for chlorate
and perchlorate formation. However, the different electro-
chemical pathways to oxygen formation in chlorate cells were
also examined. Experiments with addition of NaOH to an
operating chlorate cell were performed to see if the oxygen
evolution rate followed the addition of hydroxide or if it followed
the concentration of NaOCl. It was found that increasing
concentrations of HOCl led to a linear increase in the percent of
oxygen in the cell gas. Furthermore, it was found that the oxygen
evolution rate was diffusion limited at higher current densities.
The conclusion was that the main source of oxygen evolution is
the discharge of the OCl−.
In 1936, Joffe131 published results on the selectivity between
the OER and the ClER on graphite and magnetite electrodes in
concentrated chloride solutions. A way of calculating approx-
imate selectivities directly from the polarization curve was
presented, where the current densities at a certain potential in
H2SO4 (to obtain the polarization curve for OER) and in
concentrated NaCl were used to obtain the selectivity for
conditions under which both reactions occur. The selectivity for
Cl2 was seen to increase with current density during electrolysis
Figure 6. Dependence of the oxygen losses (%O2 in the off-gas) on the
of neutral 5 M NaCl solutions.
hypochlorite concentration, as found and reported by Hammar and
In 1945, further results were published by Rius and Llopis,132 Wranglén54 using graphite anodes. ● represent data for solutions
who performed a number of experiments, using Pt electrodes and without sodium chlorate, and × represent data for solutions with 273 g/
volumetric analysis of the formed gases as well as analyses of dm3 NaClO3. The temperature was 25 °C, cNaCl = 150 g/dm3, and j = 0.4
chloride, chlorate, and hypochlorite in the electrolyte solution, to kA/m2. Reprinted with permission from ref 54. Copyright 1964 Elsevier.
increase the understanding of the reaction mechanisms of
chlorate formation. They claimed results contrary to those of This suggests that the oxygen evolution rate due to hypochlorite
Foerster, and proposed a new overall stoichiometry for anodic decomposition is mass transfer controlled. It also increased
chlorate formation slowly with temperature (about 2%/K), which was related to an
increase in the diffusion coefficient for hypochlorite. Additions of
5ClO− + 8H 2O → 3ClO−3 + 2Cl− + 2O2 + 16H+ chlorate (273 g NaClO3/dm3) decreased the oxygen evolution
+ 16e− (19) rate somewhat. In a similar way as with the effect of temperature,
this was thought to be related to an increase in electrolyte
accompanied by viscosity that decreased the rate of mass transfer of hypochlorite
1 to the surface of the anode. Additionally, the oxidation loss
2OH− → H 2O + O2 + 2e− decreased with increasing current densities. This dependence
2 (20)
was approximately linear if the volumetric percentage of oxygen
In 1956, de Valera133 presented a model for sodium chlorate was corrected for the varying thickness of a Nernst diffusion layer
formation, based on the idea that the stoichiometry for during gas evolution at the anode. On the basis of these results
electrochemical chlorate formation according to Foerster (eq and an analysis of possible alternative reactions to 10, Hammar
10) could be replaced with simpler reactions, if the processes and Wranglén54 concluded that reaction 10, anodic chlorate
occurring in the anode diffusion layer are taken into account. de production according to Foerster, was the mass transfer
Valera proposed that the only electrolytic processes occurring are controlled oxygen evolving reaction.
oxidation of Cl− and OH− and that chlorate and hypochlorite In 1969, Jaksić et al.55 published a study where the Foerster
forms chemically in the diffusion layer. It was found that a model reactions were used to model the overall current efficiency of the
based on these considerations would also agree with chlorate process. Experiments were carried out in a stirred flow
experimental observations but without the need for chlorate reactor where pH, total available chlorine (the sum of
2989 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

hypochlorous acid and hypochlorite) in the inlet and outlet, as


well as chlorine and oxygen in the off-gas were measured
simultaneously. A good agreement between the model and the
results at different flow rates, current densities, and temperatures
was obtained, in turn supporting the Foerster reactions as
suitable for describing electrochemical chlorate formation. The
mathematical treatment by Jaksić et al.55 was further developed
by Despić et al.134 to extend its applicability to concentrated
electrolytes. In the same year, Jaksić et al.135 also suggested that
the current efficiency of the homogeneous chlorate formation
reaction should be dependent on the pH, through the activity of
the hypochlorous acid/hypochlorite pair. They thought that
Hammar and Wranglén54 had neglected the effect of CO2
absorption in the electrolyte when they reached their conclusion
that pH does not influence the efficiency of the chlorate process
in the pH = 6.5−11 range.
The most important development of the 1960s related to the
chlor-alkali and sodium chlorate processes was the invention by
Henri Beer of the DSA. However, Beer’s patents do not give any
detailed description of selectivity between chlorine and oxygen
evolution on DSAs.8,16−22 A relevant comment from the patents
is that other film-forming metal supports (Ta, Zr, W, or alloys of
these) than Ti or Nb are preferable for oxygen evolution. One of
the example compositions, that of (TiO2−IrO2−PtO2) on Ta, is
said to be “operated excellently in electrolytic processes for the
preparation of chlorine, oxygen, oxidation of organic compounds
and in galvanic baths” (“Example III” in U.S. patent 3,933,616).22
2.2.2. Academic Research during the 1970s. To our
knowledge, in 1972 Kuhn and Mortimer70 published the first
study of current efficiency for the ClER using metal oxide Figure 7. Gas-tight rotating disc electrode setup used by Kuhn and
Mortimer:70 “Diagram of the rotating disc apparatus: 1, carbon brush; 2,
electrodes, sodium chloride solutions of varying concentrations
drive pulley; 3, collar fixed to shaft; 4, spring; 5, teflon bearing, sliding fit
and at varying temperatures. The electrode materials used were on shaft; 6, teflon bearing, fixed to glass stem (8); 7, gas inlet to balance
RuO2 and IrO2 on Ti. A flow-through cell was used to keep the hydrostatic head of electrolyte; 8, glass stem; 9, titanium shaft; 10, teflon
electrolyte concentration constant. Current efficiency measure- bearing fixed to glass stem; 11, teflon cylinder screwed to the shaft end;
ments were performed at T = 20 °C and t = 70 °C, pH ≈ 2, with 12, electrode slug, secured to shaft by (11); 13, brine inlet”. Reprinted
stationary and rotating electrodes. A gastight rotating disc with permission from ref 70. Copyright 1972 Springer Science and
electrode (RDE) setup was used, which is shown in Figure 7. The Business Media.
current efficiency was calculated based on assuming that only
reactions 10 (anodic chlorate formation) and 11 (water example, gas composition. What is worth noticing from the
oxidation) occurred. Both these reactions yield the same amount point of view of selectivity is that bubbles of what was thought to
of oxygen per electron, allowing the current efficiency to be be oxygen were noticed at the rest potential when pure RuO2
calculated based on analysis of the amount of oxygen formed. electrodes were inserted into a chlorine-saturated NaCl solution
The results of the experiments showed that the current with low Cl− concentration. This was thought to result from a
efficiency for chlorine evolution at t = 20 °C on a nonrotating combined reduction of Cl2 and oxidation of hypochlorite species
electrode was increased from 97% to 99% as the chloride to form chlorate by the electrochemical reaction 10.
concentration was increased from 1 to 5 M at controlled current Again in 1972, Jaksić et al.137 published data about the effects
− −
densities. At both concentrations, increasing current density of Cr2O2− 3− 2−
7 , PO4 , NO3 , SO4 , and F on the anodic part of the
from about 3 to 20 kA/m2 had no effect on the current efficiency. chlorate process, using Pt and DSA anodes in an electrolyte of
At concentrations below 1 M, the current efficiency was lower. At 310 g/dm3 NaCl and 1 g/dm3Na2Cr2O7 at 25 °C, using the same
the lowest concentration of 0.13 M, the current efficiency system as in their previous work.55 Additions of anions were
decreased with increasing current density. When instead of the made at concentrations of 0.01 M ≤ canion ≤ 0.2 M. It was seen

stationary electrode a rotating disc electrode was used, it was that increasing concentrations of Cr2O2− 3−
7 , PO4 , and NO3 all
found that increased current densities, from 3 to 50 kA/m2, yielded higher percentages of oxygen in the off-gas. This
increased the current efficiency, up to a certain point at which the dependence seemed to level off at higher anion concentrations.
mass transport of chloride to the surface of the electrode became The effect was highest with dichromate addition and lowest with
limiting. However, it was noted that the rotation of the electrode nitrate addition. No effect on the oxygen evolution was seen from
drew evolved gases to the axis of the electrode, which would addition of sulfate or fluoride. Furthermore, since NO−3 has a no
result in blockage of the electrode area and possibly negatively pH buffering effect and phosphates have a higher buffering
affect the quality of the results at high current densities. In capacity than dichromate species, the effect on the oxygen
comparison with RuO2-coated Ti electrodes, IrO2-coated Ti evolution was thought to relate to adsorption of the anions,
electrodes exhibited higher efficiencies in dilute 0.5 M NaCl. rather than to a buffering effect on the hydrolysis of Cl2 formed at
In 1973, Kuhn and Mortimer136 reported further results from a the anode. Other factors, such as the similarity to an adsorption
purely electrochemical study without measurement of, for isotherm of the increase in oxygen percentage with increasing
2990 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

concentration of anion, and the higher effect on oxygen that HSO−4 and H2PO−4 adsorbed preferentially, displacing Cl−
percentage of Pt (which operates at a higher anode potential on RTO electrodes, elucidating the reason for the reduction in
than DSA), were also suggested as supporting the hypothesis of current efficiency upon addition of sulfate that Bondar et al.138
adsorption of anions. It was thought that the adsorption of observed. However, their measurements only covered the range
dichromate, phosphate, and nitrate increased the oxygen of potentials from 0 to 0.7 V versus SHE, lower than anode
evolution rate since they might increase the availability of the potentials expected during ClER, and were performed in dilute
surface for oxidation of hypochlorite species and that they also solutions. Additionally, some of the experiments had been
might catalyze the electron transfer reactions involved in anodic performed at pH below 0, where formation of Ru(OH)2+ 2 is
perchlorate production. predicted,47 possibly complicating the analysis.
In 1974, Jaksić56 reported the effects of several process In 1979, Fukuda et al.144 studied the effect of F− addition on
parameters on the current efficiency of the chlorate process. The the anodic behavior of DSAs (Ti/RuO2, Ti/IrO2, Ti/RuO2−
paper discussed the optimal current density, temperature, and TiO2, Ti/PtO2, Ti/Pd−PdO, Ti/Co3O4, and Ti/Rh2O3) in an
pH of the chlorate process without a stated specific focus on the aqueous solution of H2SO4 and (NH4)2SO4. The behavior was
selectivity between chlorine or chlorate and oxygen. studied at current densities in the range from 0.1 to 1000 A/m2
In the same year, Bondar et al.138 reported on the electrolysis (potentials between 1.5 to 2.5 V vs SHE). The addition of
of chloride-sulfate solutions using RTO anodes. Their most increasing amounts of F− (1 × 10−3 to 1 M NH4F) decreased the
relevant finding for the selectivity between oxygen and chlorine current efficiency of O2 formation on the Ti/RuO2, Ti/IrO2, and
was that the presence of H2SO4 decreased the current efficiency Ti/RuO2−TiO2 electrodes, instead promoting S2O2− 8 formation.
for chlorine evolution. This points to the importance of It was shown by a photographic method that F− adsorbed on the
competitive adsorption of other ions than Cl− on RTO. surface of these electrodes. Fukuda et al. rationalized the results
However, the result disagrees with that of Jaksić et al.,137 who by suggesting that F− deactivated the surface sites used for
found no effect from addition of sulfate. In the same year, oxygen evolution but not the sites used for S2O2− 8 formation.
Veselovskaya et al.139 showed that a decrease in RuO2 in the However, these results regarding the selectivity for S2O82−
RTO coating decreased the current efficiency for oxygen formation versus oxygen formation cannot be considered a
evolution. direct analog to the selectivity between chlorine and oxygen
In 1976, Agapova and Kokhanov140 reported on electrolysis of evolution, as the standard potential for S2O2−8 formation is 2.08 V
concentrated brine using Co3O4-coated Ti, finding a high versus SHE,37 a significantly higher anode potential than that
selectivity for chlorine evolution. At 2 kA/m2, low pH and 90 °C, applied for industrial chlorine evolution. Earlier studies by Jaksić
the current efficiency for chlorine evolution was above 99.8%. et al.137 found no effect of F− on the current efficiency for the
Also in 1976, Shlyapnikov141 found that titanium anodes coated chlorate process. Furthermore, Kelsall and Robbins145 have
with Pt or RuO2 exhibited higher chlorate production efficiencies predicted, based on thermodynamic considerations, that F−
than graphite or Ti coated with MnO2 or PbO2. The higher addition should result in dissolution of TiO2, which might
chlorate current efficiencies of Pt or RuO2 was associated with a complicate the analysis of the work of Fukuda et al.
lower oxygen evolution current efficiency. The optimal pH for In the same year, Iwakura et al.146 presented anodic
chlorate production was similar for all anode materials. polarization curves using a number of mixed metallic oxides,
In 1977, Buné et al.66 reported the selectivity between oxygen coated on Ti, in 1 M KOH and 3 M NaCl. They compared the
and chlorine evolution on titanium−ruthenium oxide anodes in characteristics of the RuO2 coating with coatings of SnO2−RuO2,
solutions as a function of applied current density. Buné et al. used SnO2−PtOx, and SnO2−PdOx. Relatively low current densities
one Tafel expression for the current density for each reaction. between 0.01 and 100 Am2 were applied. The mixed SnO2−
These expressions could then give the ratio of Tafel slopes for RuO2 coating exhibited a higher overpotential for oxygen
two different reactions. By measurement of polarization curves at evolution in 1 M KOH than the pure RuO2-coating, while
a high current density, when one of the two reactions is showing a similar overpotential for chlorine evolution. This same
dominating, the Tafel slope of that reaction could be decided. trend was also observed for the SnO2−PtOx and SnO2−PdOx
With the use of the ratio between Tafel slopes, the Tafel slope of coatings, although the SnO2−PtOx coating exhibited a steep
the other reaction could then be determined. Experimental increase in chlorine evolution overpotential as the current
measurements were performed using anodes composed of 30% density exceeded 10 A/m2. These results indicate that Sn-doped
RuO2 and 70% TiO2, in acidic (pH ≈ 2) NaCl solutions with RuO2 coatings and mixed Sn−Pd and Sn−Pt oxide coatings
different concentrations (300, 100, and 10 g/dm3), at 87 °C. The exhibit a higher selectivity for chlorine evolution than pure RuO2.
oxygen and chlorine content of the gas evolved was measured The stability of the coatings were not discussed, although
using GC and then converted into partial current density. It was Pourbaix diagrams45−47 predict instability of Ru at high pH and
found that the oxygen content of the evolved gas decreased as the anodic potentials in chloride solutions.
applied current density was increased from 25 A/m2 to 25 kA/m2 2.2.3. Research during the 1980s. In 1980, Caldwell and
and that the current efficiency for chlorine evolution increased as Hazelrigg147 reported on the properties of cobalt spinel anodes
the concentration of NaCl was increased. It was suggested that for chlorine evolution. The properties of pure cobalt oxide,
the reason for the decrease in oxygen evolution was due to a Co3O4, were contrasted with those of ZnII- and ZnII−ZrIV-doped
decreased surface coverage by intermediates in the OER as the cobalt oxide. The Co−Zn−Zr oxide exhibited a significantly
concentration of adsorbable Cl− increased (competitive higher surface area than the pure cobalt oxide or the Zn-doped
adsorption). cobalt oxide. The Zn−Zr−Co oxide exhibited a higher chlorine
One year later, Uzbekov et al.142 reported on the possibility of gas current efficiency (more than 99.5%) than either a Ru−Sn−
decreasing the active surface area of electrode coatings to Ti oxide coating (about 99%) or a RTO coating (about 98%) in a
increase the local current density on the active sites of the laboratory H-cell with concentrated NaCl on the anode side.
electrode, thereby increasing the selectivity for chlorine In the same year, Saito148 presented results of work on the
evolution. In the same year, Kazarinov and Andreev143 reported properties of a PdO-based anode coating (no details were given
2991 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

about the composition) for chlorine evolution. It was found that argon stream containing low concentrations of oxygen. The last
the PdO anodes had higher OER overpotentials, and lower ClER type was prepared with the same galvanic ruthenium deposit and
overpotentials, than conventional RTO electrodes. The off-gas heat treatment but with an added galvanostatic anodic
from a laboratory cell operated with a PdO anode and 5.4 M polarization in a sulfate solution. The polarization was stopped
NaCl showed significantly lower oxygen concentrations (less when a sharp increase in the electrode potential was detected.
than 0.5% at a concentration of 50 mM active chlorine) than an Auger electron spectroscopy showed the presence of titanium on
RTO (about 2% at 50 mM active chlorine) or pure RuO2 (about the surface of electrodes of the third type, while on the second
2.5% at 50 mM active chlorine). The lifetime of the PdO coating type Ti was only found deeper inside the bulk of the coating.
exceeded that of an RTO in an accelerated lifetime test. This is Pecherskii et al. thought that electrodes of the third type had a
unexpected, considering the low stability of PdO in chloride- Ru−Ti mixed oxide surface composition similar to that of DSA
containing electrolytes at pHs below neutral.149 produced by conventional means. The results of measurements
Again in 1980, Bergner150 presented work on the current in chloride solutions was that the corrosion rate was highest on
efficiency for chlorate and oxygen formation during chlorine type one and lowest on type three electrodes. The oxygen
evolution in a membrane cell. It was observed that the sum of the evolution rate followed the same trend, being lowest on
current efficiency for chlorine, oxygen, and sodium chlorate in electrodes of type three. Both the rate of oxygen evolution and
the membrane cell equals 100%. If anodes which exhibit the rate of corrosion increased with increasing applied current
decreased oxygen evolution selectivity are used in membrane densities. Increased pH increased the rate of oxygen evolution,
cells, they seem not only to decrease the oxygen evolution but the slope of the increase was lower on electrodes of type
selectivity but also to yield an increase in chlorate production. three. The decreased rate of oxygen evolution on type three
This agrees with the earlier study of Shlyapnikov.141 (RTO-like) electrodes was rationalized by the increased Ti oxide
In 1981, Ibl and Vogt62 devoted a chapter in their overview of amount in the surface separating the active ruthenium sites. The
inorganic electrosynthesis to chlorate production. Apart from a authors stated that chlorine evolution takes place from a single
detailed analysis of the overall efficiency of chlorate processes site, through an Erenburg mechanism,152 while oxygen evolution
based on process parameters such as reactor volume, current requires two “paired” sites (something that was suggested by
density, and temperature, there is also a discussion of the sources O’Grady et al.153 and Buné et al.66). Therefore, the rate of oxygen
of oxygen in the chlorate process in the chapter. This discussion evolution would be decreased significantly as the surface
is based on previously described sources, such as Hammar and coverage of Ru decreased, while the chlorine evolution reaction
Wranglén,54 and also on several articles by Jaksić et al.55 (such as rate would not be decreased as much.
refs 55, 56, 135, and 137). A general discussion about the In 1983, Denton et al.154 presented an overview of research
influence of different process parameters on the overall efficiency concerning chlor-alkali anodes performed at Imperial Chemical
of chlorate production is also given. A higher temperature, Industries (ICI), now Ineos Chlor-Chemical Ltd. Data for the
current density, and volume of the reactor for the homogeneous relationship between the Ru content of an RTO coating and the
reaction was reported to increase the current efficiency, based on oxygen content of the cell gas, as well as the chlorate
expressions similar to those Jaksić et al.55 developed. Increasing concentration in the anolyte, in an industrial diaphragm cell
flow rate through the system had both a positive effect and a were presented. The amount of oxygen in the off-gas increased,
negative effect. The positive effect was due to an increased and the chlorate concentration decreased, with increasing Ru
volumetric flow rate and thus a decreased concentration of concentration in the anode. The increase in oxygen percentage
hypochlorite species in the cell, increasing the current efficiency. was associated with a decrease in overpotential for OER with
The negative effect was due to the higher flow rate increasing the increasing Ru concentration. The overpotential also decreased
rate of mass transfer to the anode and thereby increasing the rate with increasing coating loading.
of anodic chlorate production. Ibl and Vogt62 believed the In 1984, three articles describing the development of a
negative effect is more important than the positive effect in selective titanium anode were published.71−73 The electrodes
industrial systems. were developed for increased selectivity for chlorine evolution,
In the same year, Gorodetskii et al.151 discussed the selectivity for applications in chlorate cells, and seawater electrolysis. The
between chlorine and sodium chlorate in a study that was mainly electrode was a mixed coating of RuO2−TiO2−PdSn2, forming a
focused on stability. The paper will be discussed more fully in compound of Sn, which has a high oxygen evolution over-
section 2.3, but it should be noted here that Gorodetskii et al.151 potential, and Pd, which was thought to increase the stability of
found that the amount of oxygen formed during chlor-alkali Sn in hypochlorite-containing solutions. Mixing the more
electrolysis increased with pH in the range from 0 to 5. This effect expensive Pd with RuO2−TiO2 was done to increase the stability
might be related to a decrease in the equilibrium potential for the further, at a lower cost (Ti and Ru being less expensive than Pd).
OER as the pH is increased. Furthermore, they noted that the The preparation of the anodes was done by standard thermal
chlorate formation also increased significantly. The authors decomposition of chloride precursors mixed in suitable ratios. X-
suggested that the chlor-alkali anolyte should be kept at low pH ray diffraction (XRD) was used to find ratios where more stable
(below 3) to maximize selectivity and minimize the formation of PdSn2 was formed in one phase, with solid solution RuO2−TiO2
sodium chlorate. making up the other phase in the surface.73 There was apparently
In 1982, Pecherskii et al.67 continued the work of Buné et al.,66 no reference data for the PdSn2 phase identified by XRD, and the
examining the parallel between the rate of oxygen evolution and stable form of such a phase according to Pourbaix diagrams,
the corrosion of electrodes during electrolysis through combined during both the preparation and during ClER or OER
gas phase oxygen and chlorine measurements and radiometric conditions, should be that of a mixed oxide rather than a Pd−
corrosion rate measurement. Three different types of electrodes, Sn alloy.45−47
all on Ti supports, were studied. The first type only had Electrodes with varying compositions were tested in 300 g/
ruthenium galvanically deposited on the surface. The next type dm3 NaCl at pH = 2 and t = 80 °C at current densities up to 1 kA/
had the same deposit but was subsequently heat treated in an m2. It was found that, to keep the same high activity as standard
2992 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Figure 8. Oxygen content of the cell gas as a function of cNaClO, corrected for cathodic inefficiencies, from Spasojević et al.71 Open circles represent
standard DSA anodes, and filled circles the modified coating. The dilute solution contained 30 g/dm3 NaCl, was kept at 25 °C, pH ≈ 7−9 and
electrolyzed at a current density of 1 kA/m2. Reprinted with permission from ref 71. Copyright 1984 Catalysis Research Center, Hokkaido University.

RuO2−TiO2, at least 30 atm % noble metals were needed in the are added to the solution. Only at low chloride concentrations,
coating. The activity of a mixed coating of 14 atm % Ru, 35 atm % where water oxidation is the dominating source of oxygen, did
Ti, 17 atm % Pd, and 35 atm% Sn showed the same activity as that the modified coating in itself yield a decreased selectivity for
of a 40 mol % RuO2−60 mol % TiO2 anode.73 oxygen evolution.
Selectivity measurements on both electrodes of the new In 1984, Kokoulina and Bunakova155 published results on
composition and of standard composition were performed. The oxygen evolution currents in chloride-containing solutions at
result for oxygen content of the gas evolved at j = 1 kA/m2 in applied current densities between 0.1 and 1000 kA/m2 and at
dilute brine is shown in Figure 8, which indicates the existence of pHs between 1.4 and 2.1. The electrodes used were RuO2 and
multiple mechanisms for oxygen formation. The first occurs at RuO2 + TiO2, both with loadings of 5.5 g/m2. The electrolyte
low concentrations of hypochlorite, when mass transport is used was either 1 M NaCl or 4 M NaCl at 70 °C. Kokoulina and
limiting the rate of electrochemical oxidation of hypochlorite Bunakova first determined whether any oxygen evolution
(reaction 10). Under these conditions, electrochemical water occurred without electrode polarization. The electrolyte was
oxidation is the dominating source of oxygen. The second saturated by Cl2, and under such conditions (i.e., when ECl2/Cl− >
oxygen-evolving mechanism is said to be hypochlorite oxidation. E > EO2/H2O), only the following two reactions were assumed
This reaction is mass transfer limited, which explains its gradual likely to occur on the electrode:
rise with the hypochlorite concentration. Figure 8 shows that the
new anode coating decreased the volumetric percentage of 2H 2O → O2 + 4H+ + 4e− (21)
oxygen in the formed gas by about 4% in dilute NaCl.71,72
The measurements under chlorate conditions (sodium 2Cl 2 + 4e− → 4Cl− (22)
dichromate-buffered concentrated NaCl) showed that the
oxygen evolution rate was mass transfer limited for both the As these two reactions occur, 1 mol of gaseous oxygen and 4 mol
new coating and the standard coating.71,72 In concentrated brine, of HCl forms. Thus, titration of the electrolyte was performed to
the new electrode coating did not exhibit a decrease in the rate of measure the change in concentration of HCl a certain time after
oxygen evolution. However, addition of phosphates to the the RTO electrode was immersed in the solution. However, it
electrolyte decreased the oxygen concentration in the gas when was found that the change in pH corresponded directly to the
the new coating was used, explained by strong adsorption of amount of Cl2 that had been decomposed and that only a very
phosphates on Pd.71 The authors presented further results small amount of ClO−3 had formed. The change in pH did not
concerning phosphate addition in their third paper.72 In this support spontaneous evolution of O2 under these conditions.
paper, they reported the influence of pH on the current efficiency This contrasts with the results of Kuhn and Mortimer,136 who
of the new coating and a standard coating, with dichromate noted oxygen evolution from NaCl solutions even without
buffering or phosphate buffering. It was found that a high activity polarization, when RuO2 electrodes were used.
was maintained in a broader pH range around pH 6.5 when the Kokoulina and Bunakova then applied two methods to study
phosphate buffer was used with the new coating. A high current the oxygen evolution under electrode polarization.155 First,
efficiency could be maintained up to pH ≈ 7.5. They also titration was used to determine the amount of HCl formed
commented that while normal DSA is destabilized by phosphate during electrolysis, using this amount to then calculate the
addition, the modified coating with PdSn2 was not. Both the new amount of oxygen formed. It was concluded that the oxygen
and the “standard” coating had similar long-term stability and evolution rate decreased with increasing Cl− concentration. The
efficiency. The improvement in selectivity at chlorate process second method used potential compensation to obtain jO2, the
conditions by using the new coating only appears if phosphates partial current density for oxygen evolution, at different electrode
2993 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Figure 9. Voltammograms of DSA measured at increasing temperature from Kotowski and Busse.159 Reprinted from ref 159. Copyright 1986 John
Wiley & Sons Ltd. All Rights Reserved. The figure may not be reproduced, stored in a retrieval system, or transmitted, in any form or by any means,
electronic, mechanical, photocopying, recording or otherwise, without the permission of Ellis Horwood Limited, Market Cross House, Cooper Street,
Chichester, West Sussex, England.

potentials. The method is based on the assumption that if the In the same year, Buné et al.157 published results on the effect
electrolyte is continuously saturated with Cl2-gas at a certain of the electrolyte composition on the kinetics and selectivity for
partial pressure, the electrode potential can be maintained at the OER and ClER on RTO anodes of the typical 30% RuO2-70%
reversible potential versus a RHE in the same solution. However, TiO2 composition. The measurements were carried out at a
electrochemical oxygen evolution will cause the electrode current density of 2000 kA/m2 in either 4.27 or 0.71 M NaCl, at
potential to change from the reversible value. If a current is 87 °C and pH 1.6. Effects of addition of either phosphate, sulfate,
applied to maintain the electrode potential at Eanode = E0,Cl2, the or perchlorate to the electrolyte were reported, continuing the
current that is applied will be the same as that used for the side work of Bondar et al.138 and Kazarinov and Andreev.143 After
reaction. Under such conditions, the OER current is addition of phosphate, even at low concentrations of 0.004 to 0.1
compensated by an equal and opposite current. To measure M NaH2PO4, significant increases in anode potential were noted.
the OER current at different electrode potentials, the chlorine The addition of 0.1 M of phosphate resulted in an increase in the
pressure of the gas above the electrolyte was decreased to lower selectivity for oxygen evolution, as measured by GC. Smaller
and lower values, decreasing the equilibrium chlorine potential. effects were noted upon addition of similar amounts of sulfate,
and no effect on the anode potential was noted upon addition of
The current that is applied is said to be the jO2(Eanode), yielding a
perchlorate. It was therefore concluded that phosphates and
polarization curve for oxygen evolution. The results showed 40 sulfates adsorb competitively, limiting the number of electrode
mV lower overpotentials for oxygen evolution at RuO2 than on surface sites accessible for chloride. The addition of increasing
RuO2+TiO2, indicating an increase in Cl2 selectivity for the concentrations of phosphates also increased the electrode
mixed oxide. Effects on the overpotential of changes in pH were potential during oxygen evolution from solutions containing
also noted. The results from both the titration and the potential 4.27 M NaClO4, indicating the effect of competitive adsorption
compensation methods compared favorably. also on the oxygen evolution reaction.
Also in 1984, Trasatti156 reported results of several studies of In 1986, Cairns et al.158 presented further work on ICI anode
the oxygen and chlorine evolution on oxide materials, primarily coatings for use in chlor-alkali membrane cells. It was noted that
pure oxides such as RuO2 or Co3O4. The bulk of the article the “classical” method to prevent oxygen evolution was brine
discussed the interplay between surface area (q*) and point of acidification, while modification of anodes was less frequently
zero charge in determining the activity of oxides in the relevant used. However, data for a modified “low oxygen” coating was
reactions. In the concluding part of the article, Trasatti correlated presented, although without any details regarding the
oxygen evolution overpotentials and chlorine evolution over- composition of this anode. It is probable that the coating was
potentials for several oxide materials. He found that the slope similar to those discussed earlier by Denton et al.154
between the two was close to unity. On materials, such as RuO2, In the same year, Kotowski and Busse159 presented research
where the overpotential for chlorine evolution is low, the on the oxygen evolution side reaction in membrane cells. The
overpotential for oxygen evolution is also low. He thought that results were based on measurements in both a diaphragm cell and
this indicated that the selectivity between chlorine and oxygen in a membrane flow cell. Kotowski and Busse discussed the
evolution largely was not determined by the oxide material itself. relative influence of reactions 10 (anodic chlorate production),
While the activity for both oxygen evolution and chlorine 11 (water oxidation), 12 (hydroxide oxidation), 14 (bulk
evolution shows differences between different materials, the decomposition of hypochlorite), and also the reaction (first
selectivity does not. proposed by Wranglén160)
2994 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

HOCl + H 2O → 3H+ + Cl− + O2 + 2e− (23) Based on these results, Kotowski and Busse159 concluded that
in industrial cells, oxygen is evolved in reactions 10, 11, and 23.
which is a combination of HOCl reduction (E° = 1.482 V) and Two of the three reactions occur when hypochlorite reaches the
H2O oxidation, making it relevant in the potential range 1.23 to anode. All three reactions involve water, leading the authors to
1.48 V versus SHE. suggest that the oxygen evolution overpotential measured in
Reactions 10 and 11 were found to make up between 20 and sulfuric acid could be used to predict the oxygen selectivity of the
60% of the total oxygen evolved in a membrane cell. Reaction 12 anode materials. The idea was supported by measurements on
was not seen as important, since the pH at the anode surface is RTO anodes with decreasing amounts of RuO2 in the coating.
likely quite acidic when the electrode is anodically polarized. The For these electrodes, a linear relation between the logarithm of
amount of oxygen evolved due to reaction 11 was estimated the oxygen content in the membrane cell gas and the
based on data from amalgam cells, in which very little overpotential for oxygen evolution, measured in sulfuric acid
hypochlorite is formed. The amount of oxygen evolved due to solution, was found (see Figure 10).
reaction 10 was estimated based on the amount of chlorate that
formed and the assumption that all chlorate formed in the
membrane cell was formed due to oxidation of hypochlorite at
the anode.
Based on these initial experiments, Kotowski and Busse
reached the conclusion that a large percentage (between 40 and
80%) of the total oxygen formed had to be formed in reactions
involving hypochlorite, reactions 14 and 23. All other reactants
for oxygen formation, such as Cl2O, ClO−2 , or HClO2, were
excluded based on spectrophotometric data, leaving only
hypochlorite species as reactants for oxygen evolution.
Voltammograms of DSA were then measured in 200 g/dm3
NaCl and 3 to 5 g/dm3 NaClO. The voltammograms exhibited
peaks at 400 to 500 mV and around 1000 mV (see Figure 9), with
increasing intensity at increasing temperatures, which led to the
hypothesis of a coupled redox reaction occurring at the anode.
Kotowski and Busse thought it likely that concurrent oxidation,
according to reaction 23, and reduction of hypochlorite
2e− + H+ + HOCl ⇌ Cl− + H 2O (24)
occurred on the anode. The net result would then be production
of oxygen according to
2HOCl → O2 + 2H+ + 2Cl− (25)
To quantitatively test this hypothesis, further experiments
were then performed at room temperature and high pH (pH ≈ Figure 10. Relationship between the overpotential for the OER and the
10, to prevent chlorine evolution), at which only reactions 10 and volume percent of O2 in the off-gas, as found by Kotowski and Busse.159
23 were thought to occur. The electrolyte contained 30 g/dm3 The slope of the relationship between ηOER and log %O2 was found to be
NaCl and approximately 5 g/dm3 NaClO. Since the number of −169 mV per decade. Reprinted from ref 159. Copyright 1986 John
electrons released per mole of O2 formed is different for the two Wiley & Sons Ltd. All Rights Reserved. The figure may not be
reactions (2 electrons for reaction 23 and 4 electrons for reaction reproduced, stored in a retrieval system, or transmitted, in any form or
by any means, electronic, mechanical, photocopying, recording or
10), the percentage of the charge due to each of the reactions
otherwise, without the permission of Ellis Horwood Limited, Market
could be calculated. Three different anode coatings were used, a Cross House, Cooper Street, Chichester, West Sussex, England.
“normal”, a “special” (only said to be containing another
platinum-group metal together with Ru and Ti), and a Pt/Ir
coating. In industrial membrane cells, the gas evolved from the
cell using the normal coating contained 0.45 vol % oxygen, the Again in 1986, Buné et al.68 published results using GC to
special coating gave a gas containing 0.17 vol % oxygen, and the measure the selectivity for oxygen formation using DSA. This
Pt/Ir coating gave a gas containing 0.09 vol % oxygen. For the time, the composition of the anode was changed; previous
standard coating, the percentage of the oxygen evolved due to selectivity studies66,67 had examined the effect of current density,
reaction 10 constituted 21% of the overall oxygen formation. For chloride concentration, and pH. The molar percentage of Ru in
the special and the Pt/Ir coatings, the percentages of oxygen the coating was varied from 10% to 100%, similar to what was
generated from the same reaction were 0% and 69%, respectively. done in the study of Kotowski and Busse. The experiments were
In each case, the rest of the oxygen was accounted for by reaction done in a similar way as in previously described articles by the
23, the coupled hypochlorite reduction/water oxidation reaction. same group.66,67 The conclusion from the measurements was
Reaction 14, homogeneous decomposition of hypochlorite, was that the selectivity for chlorine formation increased as the
not considered to be important. Kotowski and Busse159 had done amount of Ru in the coating decreased, especially as the amount
previous experiments that confirmed that several heavy metal of Ru was decreased below 30%. The rationale was similar to that
ions promote the rate of this reaction. However, they did not find of Pecherskii et al.,67 that oxygen evolution is dependent on
that solids, such as DSA or solid Pt/Ir, increased the rate of adjacent Ru sites, while chlorine evolution is not. Their results
reaction. are summarized in Figure 11.
2995 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

oxygen evolution is around 130 mV. This means that ϵ > 0 (eq
28), so that an increase in current density yields an increase in
current efficiency for chlorine evolution (eq 26). That this is the
case for chlorine evolution on DSA has been shown many times
(refs 54, 56, 57, 62, 66, 69, 70, 75, 131, 154, and 162−164).
Tilak et al. then went on to perform experiments to test these
equations. The experiments were performed at T ≈ 95 °C using
an electrolyte of 5 M NaCl and 2 M NaClO4 saturated with
chlorine at pH ≈ 4. Coatings with decreased j0,oxygen evolution
(RuO2 coatings doped with different amounts of HfO2 or SnO2)
were tested using polarization measurements, showing that the
Figure 11. Dependence of the oxygen content (vol %) of the cell gas on
overpotential for chlorine evolution was usually higher on such
RTO anode composition obtained by Buné et al.68 (plotted using data electrodes than on standard DSA. On the basis of galvanostatic
from the article68). The NaCl concentration was 4.27 M, pH ≈ 1.6, and selectivity measurements at 2.32 kA/m2, the selectivity for
the current density was 1.2 kA/m2. Adapted with permission from ref 68. chlorine evolution was decreased for most of the Hf- or Sn-
Copyright 1986 Springer Science+Business Media. containing coatings. It was also found that the current efficiency
for chlorine evolution increased with increasing current density.
In 1987, Spasojević et al.161 presented the optimization of a The percentage of oxygen in the off-gas using anode materials
RuO2−TiO2 anode for use in chloride electrolysis. Optimal Ru− with increasing roughness factor (increasing Areal/Ageometric) was
Ti ratios, loadings, and calcination temperatures to achieve an then measured at the same current density. The results are shown
active, selective, and stable anode were presented. Of interest to in Figure 12. The roughness factor was changed by using
the question of selectivity is that results (very similar to those of
Buné et al.68) indicating a clear decrease in partial current density
for OER for coatings with Ru content below 40% were given. The
lifetime of the anode during electrolysis of a dilute chloride
solution was also maximized at around 40% Ru.
Two years later, Tilak et al.69 published an article examining
the selectivity in the chlorate process. Tilak et al.69 derived a
mathematical expression, based on Tafel slopes, for the current
efficiency of a main reaction when two competing reactions are
occurring simultaneously. The equations are valid when both
reactions are activation controlled. Tilak et al. believe that the
relevant oxygen-evolving side reaction in the chlorate process is
anodic oxidation of water (reaction 11). They disregard the
reaction discussed by Kotowski and Busse,159 reaction 23, since
they did not think it was “satisfactorily demonstrated”. The
approach in developing the equations was similar to that of Buné
et al.66 However, a more extensive set of equations was
developed. The central one was
Figure 12. Relationship between surface roughness and the percentage
1 of O2 in the gas phase, from the study of Tilak et al.69 Reproduced with
em =
1 + δem−ϵI −ϵ (26) permission from ref 69. Copyright 1988 The Electrochemical Society.

where I is the overall current density, em the current efficiency for


the main reaction, different coating deposition methods (brushing or spraying),
different drying and calcination temperatures, and by varying the
j0, p
δ= × e(k ′×ΔE) × j0,m A ·ϵ number of coating layers applied. The measurements showed
j0,m (27)
that an increase in roughness factor gives an increase in the
oxygen percentage in the off-gas. This seems to support the idea
and expressed by Uzbekov et al.142 and Trasatti and Lodi,165 that an
k − k′ increase in surface area yields a decrease in local current density
ϵ= and thereby a decrease in selectivity for chlorine evolution.
k (28)
In 1989, Hardee and Mitchell162 published an article regarding
k and k′ are the inverse of the Tafel slopes for the main reaction the selectivity for oxygen evolution at chlorate conditions,
(chlorine evolution) and the “parasitic” reaction (oxygen examining the effect of process conditions on the percentage of
evolution), respectively. ΔE is the difference in reversible oxygen in the gas phase, using an undivided cell with a DSA. The
potential between the main reaction and the parasitic reaction. electrolyte contained NaCl (either 50 or 200 g/dm3) and 2 g/
A is the effective surface area of the electrode. The main dm3 sodium dichromate at pH = 6.5 and 60 °C.
implication of the expressions is that the selectivity for the main Hardee and Mitchell found that hypochlorite decomposition
reaction (chlorine evolution in the case of chlorate and chlor- was the main source of oxygen evolved. The percentage of
alkali production) could be increased if materials with a lower j0,p oxygen in the evolved gas was dependent on the hypochlorite
(j0,oxygen evolution) were used (eq 27). Furthermore, for chlor-alkali concentration at both sodium chloride concentrations. At the
and chlorate processes, k > k′ since the Tafel slope for chlorine higher NaCl concentration, the dependence was approximately
evolution on DSA is around 40 mV while the Tafel slope for linear (see Figure 13). Extrapolation to cNaClO = 0 indicated that
2996 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

evolution rate at increasing concentrations of dichromate.


Looking at the data presented by Hardee and Mitchell, a slight
increase in the percentage of oxygen with increasing concen-
tration of dichromate can be seen. Finally, long-term data for two
anodes were reported: one of standard DSA and one with a secret
modified composition. The modified coating, which had been
adjusted based on the importance of the hypochlorite oxidation
at the anode, showed a significantly lower percentage of oxygen
in the off-gas.
In the same year, Zhinkin et al.163 contributed a study mainly
focused on stability of DSA. The work will be discussed in more
detail in section 2.3, but it will be noted here that the authors
Figure 13. Dependence of hypochlorite concentration on the percent
showed that during chlor-alkali electrolysis using the conven-
oxygen evolved as found by Hardee and Mitchell.162 The concentration tional DSA and concentrated NaCl solution, the amount of
of sodium dichromate was 2 g/dm3, the temperature was 60 °C, and the oxygen in the off-gas increased with pH in the range of 2−5
current density was 2.5 kA/m2. DSA were used. Reproduced with (possibly related to the pH dependence of the O2/H2O
permission from ref 162. Copyright 1989 The Electrochemical Society. equilibrium) and decreased with increasing current density in
the range from 2 to 10 kA/m2.
the background oxygen production level, due to water oxidation, 2.2.4. Research during the 1990s. In 1990, Boodts and
would be equivalent to 0.3% oxygen in the gas phase, for the 200 Trasatti166 published an article investigating the activity of
g/dm3 NaCl electrolyte. To determine the rate of homogeneous Ru0. 3Ti(0.7−x)SnxO2 coatings for oxygen evolution from 1 M
oxygen evolution, the cell was also operated in the same way but HClO4. The selectivity was not investigated per se but rather the
without the electrodes inside. A nitrogen flow, of similar activity and surface charge of this ternary oxide. Boodts and
magnitude to the hydrogen evolved in normal operation, was Trasatti found that the activity, measured as j(1.43 V vs RHE)/q*
led through the cell to remove the formed oxygen. It was seen (current normalized by the surface charge), for oxygen evolution
that the oxygen evolution was dependent on the NaClO is increased by increasing Sn addition. In contrast to the
concentration also in this case. The sum of the oxygen suggestion of Krstajic et al.,73 Boodts and Trasatti believed
percentages from water oxidation and homogeneous hypochlor- addition of Sn to the coating should decrease the selectivity for
ite oxidation was not enough to cover the total percentage of chlorine evolution, and they proposed that it was PdSn2 rather
oxygen in the gas phase. Thus, the remaining part of the oxygen than Sn itself that promoted the selectivity for chlorine evolution
evolved was thought to be produced in reaction 23, as proposed seen in the study of Krstajic et al.
by Kotowski and Busse.159 Also in 1990, Couper et al.167 reported further data for an ICI
Hardee and Mitchell162 also examined the effect of some other membrane chlor-alkali process anode coating. The coating, with
process parameters. Increasing NaCl concentrations, from 50 to unnamed composition, was an improvement of the previous low
300 g/dm3, were seen to reduce the oxygen evolution rate. The oxygen coatings154,158 and had been developed for a high oxygen
oxygen percentage in the off-gas decreased linearly with evolution overpotential. Experimental data comparing the new
increasing NaClO3 concentrations between 0 and 400 g/dm3. low oxygen anodes with previous coatings indicated significantly
It was thought that this might be due to competitive adsorption lower oxygen percentages in the off-gas when using the new
of chlorate onto the anode. There was no consideration of the anode coating, at least up to pH ≈ 4.5. The results were explained
explanation of Hammar and Wranglén,54 who showed that the using eq 29,
increased viscosity of the electrolyte with increasing chlorate log(oxygen content) = −Aη(O2 ) + B pH (29)
concentrations was a likely reason for the decreased oxygen
evolution rate. Higher temperatures were seen to yield higher which describes the oxygen evolution in membrane cells where
rates of oxygen evolution. This was partly explained by improved the oxygen content of the cell gas is related partially to the pH
kinetics for homogeneous hypochlorite oxidation and partly by and partially to the overpotential for oxygen evolution. A and B
hypochlorite oxidation having a higher increase in rate with are introduced simply as fitting parameters without any physical
temperature than chloride oxidation. pH apparently had no effect interpretation.
in the 5−7 range, but an increased oxygen evolution rate was In the same year, Bergner et al.41 reported data for HCl
noticed at higher pH. This was thought to be related to the addition to membrane chlor-alkali cells. This addition is done to
increasing prevalence of the deprotonated hypochlorite species neutralize any OH− that might migrate from the cathode side of
at higher pH. The selectivity for oxygen evolution was seen to the membrane. Bergner attempted to obtain data on suitable acid
decrease with increasing current density. This was thought to be additions to the brine inlet of an industrial membrane chlor-alkali
related either to difference in Tafel slope between the chlorine- cell. The study did not attempt to decide the magnitude of
and the oxygen-evolving reactions or due to a limiting current different oxygen-producing reactions but presented suggestions
density for hypochlorite oxidation being reached. Hardee and for HCl addition that were shown to reduce the oxygen levels
Mitchell also examined the effect of dichromate concentrations significantly, from just below 2% to 0.4% (see Figure 14). These
between 0 and 5 g/dm3. At 0 g/dm3, the oxygen percentage results agreed with those previously obtained by Gorodetskii et
increased significantly. However, this was due to the large al.151
decrease in cathodic current efficiency that is associated with Also in 1990, Wanngård57 reported results concerning the
complete removal of dichromate from the electrolyte. When this oxygen forming reactions in chlorate electrolysis in a pilot scale
effect was corrected for, the dichromate amount was said to have system, with industrially relevant current densities and temper-
no effect on the oxygen evolution rate. This contrasts with the atures and using a RTO anode and a steel cathode. Wanngård
results of the work of Jaksić et al.,137 who saw increase in oxygen found, in agreement with other studies,54,162 that the oxygen
2997 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

of the oxygen evolved originated from homogeneous hypochlor-


ite decomposition (reaction 14).
In 1992, Wanngård168 published a report concerning impurity
effects in the chlorate process, where it was noted that electrolyte
concentrations of Pb, Mn, Sn, F, Al, Si, and BaSO4 should be
minimized to prevent deactivation (due, for example, to
blockage) of the anode. Tests of hypochlorite decomposition
catalysts under chlorate conditions at 70 °C had also been carried
out, finding that Ni, Cu, Co, and Ir increased the oxygen
percentage in the cell gas upon addition, while Pt, Ru, and Fe had
no effects on the oxygen percentage (the form of the metal, i.e.,
salt, metallic, or oxidic, was not indicated). Wanngård also
pointed out some important areas for further studies of the
selectivity, such as studying the amounts of oxygen forming at the
anode versus in the bulk electrolyte, deciding criteria for
sufficient electrolyte purity and regarding demands on analysis
methods used to detect impurities, and finally further studies of
which current concentration gives the optimum selectivity.
Figure 14. Effect of anolyte acidification with HCl on current
efficiencies in chlor-alkali production, from Bergner.41 Original caption Also in 1992, Traini and Meneghini169 presented experimental
reads “Current efficiencies of products and by-products in brine pH data for the percentage of O2 in a chlor-alkali membrane cell off-
range from −0.8 to 2.0. The points (×) at pH 2 correspond to the means gas for a De Nora “Low-oxygen coating”, which decreased the
of Figure 5.”. Reprinted with permission from ref 41. Copyright 1990 percentage of O2 by about 0.4% at pH 4.0 in comparison with
Springer Science and Business Media. that of a conventional DSA coating. No information was given
about the composition of the low-oxygen coating, although it was
pointed out that an improved distribution of chlorides at the
surface of the anode could be used to prevent local low brine
concentration in the off-gas was related to the concentration of
concentrations leading increased rates of oxygen evolution.
hypochlorite species (the sum of [HOCl] and [ClO−]). In turn,
In the same year, Czarnetzki and Janssen170 presented a study
the concentration of hypochlorite species was found to be
of hypochlorite production by galvanostatic electrolysis (j = 2.76
dependent on pH, temperature, ionic strength, and the cubic
kAm−2 or j = 5.54 kAm−2) of dilute NaCl solutions (6 to 90 g/
root of the current concentration (total current divided by the
dm3) in a lab-scale membrane cell at alkaline pH (in the range 8−
total electrolyte volume in the system). Furthermore, Wanngård
10), using a commercial RTO anode. The study focused
found that the oxygen percentage was independent of the flow
specifically on the initial behavior of the cell, with measurements
velocity in the electrode gap. This contrasts with results obtained
covering about 1 h of electrolysis. It was found that increasing the
in studies at low temperature,53,54 where the oxygen evolution
hypochlorite concentration resulted in a linear increase in oxygen
rate was mass transport limited and indicates that the oxygen
formation rate, although an extrapolation of the trend to zero
evolution was kinetically limited at the higher temperatures used
hypochlorite concentration resulted in a significant background
industrially. Wanngård called
oxygen formation rate. Increasing the initial NaCl concentration
%O2,off − gas instead resulted in decreased oxygen formation rates. Fur-
= SR thermore, increasing the current density resulted in a slight
[HOCl + OCl−]electrodegap (30) increase in oxygen formation rate. It was concluded that the rate
a “selectivity ratio” (SR) when discussing the results. Linear of chloride oxidation was mass transport limited in the study, and
relationships (indicating that the rate of oxygen evolution is therefore, the effect of adding nonreactive anions was attributed
linearly proportional to the amount of hypochlorous acid species to their effect as supporting electrolytes. The addition of
in the cell) were found when plotting SR versus (1/t), where t is supporting electrolytes reduced the migration mass transfer of
the temperature in degrees centigrade. The same slope was found chloride to the electrode surface, thus further reduced the
in all cases, except for when the electrolyte was contaminated by limiting current density for chloride oxidation. However, it was
compounds active for hypochlorite decomposition. Wanngård also noted that the addition of ClO−3 , ClO−4 , or SO2−4 seemed to
claimed that this linear dependence implied reaction control of catalyze the oxygen formation, as a concomitant decrease in
the oxygen evolution under chlorate conditions. Neither anode potential was measured. The authors also estimated that
changing the electrode gap from 2 to 3 mm nor increasing the the pH at the anode during electrolysis was about 2.0, based on
height of the electrode blades (i.e., changing from rectangular the formation rate of oxygen and an assumed concomitant
blades with 0.333 m height and 0.379 m width to rectangular production of H+ ions. Based on this, they concluded that the
blades with 0.8 m height and 0.3 m width) altered the slope of the electrochemical oxidation of ClO− should be negligible and
linear dependence. However, the increase in height decreased the suggested that HOCl should be the dominating species at this
value for SR, implying that the electrode height (or possibly the pH. However, according to Figure 1, it is more probable that at a
total area, as the higher blades had a higher geometric surface pH of 2, HOCl will be converted to Cl2 and water. Finally, the
area) had an impact on the oxygen formation. Wanngård authors suggested that their chlorate formation results could be
cautioned that the effect on SR that higher electrodes exhibited explained if the formation of chlorate occurred through both the
could point to a difficulty in drawing conclusions relevant for full conventional chemical conversion of hypochlorite, as well as
scale cells from laboratory scale data. Wanngård further claimed through a direct electrochemical oxidation of Cl− to ClO−3 .
that experimental studies performed at the same company, at In 1993, Bergner and Hartmann36 published an article
similar temperatures (50 to 90 °C), had shown that only 5−10% concerning oxygen evolution in membrane chlor-alkali cells.
2998 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Results were presented for oxygen and chlorate production in in this review. Furthermore, Trasatti reviewed some data that
industrial membrane cells, operated at industrially relevant points to the importance of a connection between surface redox
process parameters over a period of several years. The cells were reactions of electrocatalyst materials (e.g., changes the oxidation
either equipped with “low oxygen” or “high oxygen” anodes, with state of Ru in the surface coating of DSA) and the redox
unnamed compositions. The results confirmed the previous potentials of reactants. Trasatti comments that the reason why
trends141,150,154 in that the anode designed for lower oxygen Tilak et al.69 and Spasojević et al.71 claim that Sn addition
production exhibited a lower oxygen production but also a higher increases the selectivity for chlorine evolution is that “...the redox
chlorate production. transition leading to oxygen evolution on RuO2 is pushed to
In the same year, the first research, to our knowledge, that more positive potentials by the presence of the inactive
employs mass spectrometry (MS) for detection of Cl2 and O2 component. This is also the case for RuO2+TiO2 electrodes”.
formed during electrolysis using DSA was published. Takasu et In 1994, further lab-scale research concerning the current
al.171 adapted DEMS for use with a titanium net coated with efficiency in the chlorate process was presented by Baolian and
different compositions of RuO2, TiO2, and IrO2. The potentials Wenhua.173 The main contribution of the study was a new
where chlorine and oxygen, respectively, were first detected by formula for calculating the current efficiency for a chlorate cell
the MS were termed the threshold electrode potentials. The utilizing an oxygen cathode instead of the conventional hydrogen
difference between the two threshold potentials was proposed as cathode. The study again reconfirmed previously seen trends
a measure of the selectivity for the two reactions (see Figure 15). regarding optimal pH for chlorate production (finding an
optimum at around pH 6.5−6.75), sodium chloride concen-
tration (increased efficiency with increasing NaCl concentra-
tion), and sodium chlorate concentration (a slight increase in
current efficiency with increasing NaClO3 concentration).
However, the study also examined the effect of additions of
Na2HPO4, Na2Cr2O7, or NaF to the electrolyte at two different
concentrations of NaCl (either 200 g/dm3 or 50 g/dm3).
Additions of increasing amounts, up to 7 g/dm3, of Na2HPO4 or
Na2Cr2O7 resulted in a slight decrease in current efficiency for
chlorate formation. However, the addition of NaF resulted in
increases in current efficiency of about 1% at the high NaCl
concentration and about 3% at the low NaCl concentration.
These results agree with the general conclusion of Fukuda et
al.,144 who found that F− addition could reduce the OER
Figure 15. Threshold electrode potentials of (a) oxygen and (b)
selectivity of DSAs, while they disagree with those of Jaksić et
chlorine evolution on RuO2−TiO2 electrodes of varying composition. al.,137 who found no effects on current efficiency for chlorate
Reprinted with permission from ref 171. Copyright 1993 Elsevier. formation from fluoride addition.
In 1997, Eberil et al.164 examined the selectivity issue in a
Also, the ratio between chlorine evolved and the total gas evolved chlorate system, calculating the efficiency for chlorate electrolysis
at a certain potential (above the threshold potential for both based on the equation
reactions) was proposed as a measure of selectivity. Based on the ΦeNaClO3 = 100 − 2V (O2 ) − V (Cl 2) − 2.5
results, the authors conclude that a 20% RuO2 + 80% TiO2 (31)
coating shows the highest selectivity for chlorine evolution. The where 2.5 was a previously determined efficiency for cathodic
applicability of the results to actual anodes in industrial reduction in the cell and V are volume fractions. Experiments
production is uncertain, since DEMS studies are always done
under chlorate conditions showed that ηNaClO3 increased with
at low current densities. In this case, the maximum current
density reached during the sweep was about 40 A/m2. potential, reaching a plateau at anode potentials E > 1.37 V versus
In the same year, Nakajima et al.172 presented a study on the SHE. The efficiency was constant at pH 6−7 at potentials above
properties of Ru−Ti−Sn oxide and Ru−Sn oxide electrodes used 1.365 V versus SHE. At potentials lower than the plateau value,
for chlorine evolution in a laboratory-scale membrane cell. XRD the efficiency was significantly higher for solutions with pH 6−
indicated that the Ru−Sn oxide was of the rutile type, but the 6.5. The maximum ΦeNaClO3 was found at potentials close to the
structure of the other electrodes was not clearly stated. The Ru− so-called “critical potential”, Ecr. This potential has been
Sn oxide electrodes exhibited lower anode potentials than both associated with a break in the Tafel line and an increase in the
Ru−Ir−Ti, Ru−Ti, and Ru−Ti−Sn electrodes. Similarly to rate of decomposition of the anode.165,174 Furthermore, they
conventional RTO electrodes, the volume percent O2 in the off- found that the efficiency of chlorate production decreased with
gas increased with the amount of Ru in the Ru−Sn oxide coating. increasing temperature but increased with increasing current
Lower Ru percentages in the mixed Ru−Sn coating also led to density. Thus, a high current efficiency can be maintained at a
longer coating lifetimes, which also matches the behavior of higher temperature if the current density is increased.
RTO. In 1998, Arikawa et al.75 published DEMS results on the
In 1994, Trasatti’s latest extensive review of DSA and related selectivity between oxygen and chlorine evolution on RTO and
conductive metal oxides was published.26 While the review RuO2. Several known effects, regarding NaCl concentration,
discussed both chlorine and oxygen evolution on DSA and other oxide loading, and RTO composition, were reconfirmed by their
materials, there is very little said about the selectivity between the experiments. The experiments using different concentrations of
two reactions. The main reference is to the studies of Krstajic et NaCl showed that the selectivity for chlorine evolution increased
al.73 and Spasojević et al.71 on a selective water electrolysis and with increasing concentration of NaCl. Additionally, the
chlorate production anode, which has been described previously selectivity increased with increasing anode potential (and current
2999 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

density) in solutions with high chloride concentrations (2 M). A increasing temperature. At Ea > Ecr, ΦeO2 increased with increasing
number of measurements were also carried out at a lower temperature. The efficiency for chlorate production was
chloride concentration of 0.3 M. At this concentration, the determined largely by ΦeO2. A higher temperature, a lower pH
selectivity for ClER in general decreased with increasing anode
and a higher concentration of NaCl increased ΦCl e
. The
potential. The decrease was more severe when the oxide loading 2

of RuO2 was increased from 0.1 to 0.9 mg/cm2. For a fixed oxide measured efficiencies were also compared to those obtained in
loading of RTO, it was found that the selectivity for ClER was the system164 with directed circulation and a holding tank for
improved over that of pure RuO2 when the RuO2 content in the chlorate formation. The system with the holding tank had higher
efficiencies and showed similar trends in efficiency at E < Ecr. No
coating was between 20% (the lowest concentration) and 60%.
experiments at E > Ecr had been done in the system with the
2.2.5. Research during the 2000s. This section is divided
holding tank and directed circulation. Furthermore, a difference
into one part describing experimental studies and one part
between the Ir-containing DSA and the standard DSA was noted.
devoted to a more detailed description of the first-principles
For the standard DSA, the maximum ΦeNaClO3 was moved to
modeling of chlorine and oxygen evolution on metallic oxides.
This latter field has seen a large growth during the past few years. higher anode potentials (higher applied current densities), and
In the present review, these studies are described after the ΦeO2 increased overall, when the chloride concentration was
experimental studies, although the approach used in some of the decreased. This could be interpreted as resulting from
experimental studies has been inspired by recent results from concentration limitations of Cl− at the anode surface. However,
first-principles calculations. when Ir-containing DSA were used, ΦeO2 did not increase as the
2.2.5.1. Experimental Studies. In 2000, Eberil et al.74 chloride concentration was decreased. The reason for the
published a thorough examination of the current efficiency for increase in oxygen evolution at standard ruthenium−titanium
chlorate production, by measuring the evolution of chlorine and DSA, once the critical potential is reached, was supposed to be
oxygen, using both DSA (30 mol % RuO2 + 70 mol % TiO2) and the formation of volatile RuO4. As IrO2 increases the stability of
Ir-containing DSA (15 mol % RuO2 + 15 mol % IrO2 + 70 mol % the coating, it was proposed that this process is hindered on Ir-
TiO2). The maximal ΦeNaClO3 was again found at potentials E ≈ DSA. The Ir-DSA coating showed other advantageous character-
1.4 V versus AgCl/Ag, close to Ecr (see Figure 16). This istics, such as somewhat lower ΦeCl2 and higher ΦeNaClO3, at most
coincided with the minimum in ΦeO2. Depending on whether the current densities.
In 2001, Byrne et al.176 published a tertiary model of current
anode potential was below or above the critical one, the distribution in the chlorate cell, including mass transport through
dependence of the efficiency for oxygen evolution, ΦeO2, on the diffusion, migration, and convection for relevant species, using
anode potential, Ea, was different. At Ea < Ecr, ΦeO2 decreased with the Nernst−Planck equation. Butler−Volmer expressions were
used to model reactions 2, 11, and 23 on the anode. The last
reaction was modeled for oxygen formation through both
electrochemical oxidation of hypochlorite and electrochemical
oxidation of hypochlorous acid. The possibility of different type
of heterogeneously catalyzed, nonelectrochemical, decomposi-
tion of hypochlorite on the surface of the electrode was not
considered. Byrne found that electrochemical water oxidation
was an insignificant source of oxygen when compared with that of
anodic decomposition of hypochlorous acid or hypochlorite. The
latter reactions were the dominating source of oxygen at the
anode. Two different trends were seen for the anodic
decomposition of hypochlorous acid and the hypochlorite ion.
While the decomposition of hypochlorous acid was highly
dependent on the flow rate, the rate of decomposition of
hypochlorite ions was found to be relatively independent of the
flow rate. This agreed well with the results of Wanngård,57
indicating that hypochlorite, rather than hypochlorous acid,
might be the main reactant in the anodic decomposition of
hypochlorite species in the chlorate process. The model was
experimentally validated using a cell with segmented electrodes
in a vertical column.177 It was found that the model corresponded
Figure 16. Current efficiencies for O2 and Cl2 formation as a function of well to the experimental data only at low flow velocities between
anode potential, determined under different conditions. Unfilled, half- 0.25 and 0.75 m/s and near the inlet inside the column.
filled, and filled symbols depict measurements at 60, 80, and 90 °C,
The same year, Jirkovský et al.76 published a study concerning
respectively. The series of symbols (circles, squares, triangles, and stars
with increasing number of points) depict measurements done at sol−gel-prepared RuO2 coatings on Ti meshes. The study was
different current densities from 0.15 to 1.7 A/cm2. Original caption focused on the selectivity for chlorine and oxygen evolution on
reads (Eberil et al.74 used the symbol η to signify current efficiency, coatings made up by crystals of different sizes. The particle sizes
rather than using the conventional Φe175): “Dependences of ηCl2 and ηO2 increased with the annealing temperature used during the
on Ea of ORTA at different ia and T in solution containing 150 g l−1 NaCl preparation. The DEMS technique was used to measure the
and 350 g l−1 NaClO3; pH 6.5.” ORTA is an abbreviation meaning oxygen and chlorine evolution during cyclic voltammetry (CV)
ruthenium−titanium oxide anode (RTO). Reprinted with permission measurements. The study focused on examining the activity of
from ref 74. Copyright 2000 Springer Science and Business Media. different crystal edges on RuO2 nanocrystals. The electrolyte was
3000 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

a dilute solution of 0.1 M HClO4 and 0.3 M NaCl. For the thought to be active for oxygen evolution, sites with one nickel
coatings with the smallest particles, Jirkovský et al. found that the atom were thought to be active for chlorine evolution, and sites
onset potential for chlorine evolution was more positive than that with two Ni atoms were imagined to be selective for oxygen
for oxygen evolution, as opposed to what was found in the study evolution. The third type was thought to be less active for oxygen
of Takasu et al.171 However, Arikawa et al.75 also made evolution than the first type. This model could describe some
measurements at cNaCl < 0.5 M, finding that the threshold trends observed in the experimental data but was unable to
potential (or onset potential) of oxygen evolution was lower than account for the low activity for chlorine evolution that coatings
that of chlorine evolution. Furthermore, Jirkovský et al.76 found with the highest Ni content exhibited. There was no comment on
that the Tafel slopes of both oxygen and chlorine evolution were the stability of the coating, but Pourbaix diagrams45−47 indicate
114 mV/dec. These values are, like in the previous studies using that Ni2+ might dissolve from the coating.
DEMS,75,171 much higher than what is normally reported for Also in 2008, Macounová et al.180 also presented results for Fe-
RuO2. For coatings prepared at high temperature, with larger doped RuO2 (Ru0.98Fe0.02Ox and Ru0.7Fe0.3Ox) prepared using
particles, the threshold potential for oxygen evolution shifted to a sol−gel synthesis. XRD indicated that the Fe dopants entered
higher potential than that for chlorine evolution. The selectivity into the overall Ru rutile lattice. Selectivity measurements using
for chlorine evolution was higher on coatings prepared at these DEMS indicated that the selectivity was affected strongly by the
higher temperatures and showed a linear dependence on the gas present in the 0.15 M NaCl−0.1 M HClO4 electrolyte. In the
annealing temperature. The authors rationalized the increased Ar-saturated electrolyte, the selectivity for Cl2 was high for both
selectivity for chlorine evolution based on the types of crystal electrode compositions, whereas in oxygen-saturated electrolyte,
edges that were more prevalent at the larger particles that formed oxygen evolution was the dominating process for the
at these temperatures. However, the specific surface area of the Ru0.7Fe0.3Ox electrode. The authors suggested that the trend
coating was about half as large on the electrodes prepared at 900 could be explained by oxidation of the Fe sites in oxygenated
°C. It is known from previous studies that an increase in current solutions.
density increases the selectivity for chlorine evolution. Whether The same group also published an article concerning the
this had any effect does not appear to be considered in the article. Ru0.8Co0.2O2−x material in 2008.78 The electrodes were prepared
Also in 2006, Santana and De Faria178 presented research on by sol−gel synthesis onto Ti meshes. The study used X-ray
the electrolytic production of Cl2 and O2 using DSA doped with photoelectron spectroscopy (XPS) and TEM to draw con-
CeO2 and Nb2O5. Their conclusions were that Nb2O5 had no clusions regarding the connection between the surface
effect on the activity or selectivity of the coatings. However, morphology and the selectivity between chlorine and oxygen
introduction of even 1 mol % CeO2 into DSA decreased the low evolution on this material. DEMS measurements were done in
Tafel slope for O2 evolution to 30 mV/dec from 40 mV/dec, but the same way as the previous articles from the same group, by
only at current densities lower than 10 A/m2. On the other hand, using CV and DEMS. This time, the solution not only contained
CeO2-doping did not seem to affect the Tafel slope for Cl2 0.1 M HClO4, but also (like in refs 76 and 77) 0.3 M NaCl. The
evolution. CeO2-doped DSA thus exhibited higher O2-selectivity addition of chlorides shifted the onset potential for oxygen
at low current densities. evolution to more anodic potentials. The particle size did not
In 2007, Panić et al.179 presented stability and selectivity data have any effect on the rate of oxygen evolution, but an increased
on RTO deposited, by use of a sol−gel procedure, on a ternary selectivity for chlorine evolution (due to an increased rate of
titanium carbide, Ti3SiC2. It was thought that this carbide would chlorine evolution) was noted on coatings with larger (d > 40
be much more stable as a support material than Ti. Instead of the nm) crystals. The onset potential for the chlorine evolution
more usual 30% RuO2 + 70% TiO2, a coating with equimolar reaction was about 1.1 V versus SCE, while the onset potential for
amounts of ruthenium oxide and titanium oxide was used. OER was approximately 1.2 V. At higher potentials, the chlorine
Experimental results suggested that the selectivity for chlorine evolution current was almost constant, while the oxygen
evolution was more favorable for anodes on the Ti3SiC2 support evolution current continued to increase. The Tafel slope for
than on Ti. However, the new support was shown to be much less the chlorine evolution was 120 mV/dec below 1.2 V and a very
stable in an accelerated lifetime test than the standard Ti support. high value of about 600 mV/dec at higher potentials. The change
In 2008, Macounová et al.77 published DEMS results on the in Tafel slope at 1.2 V was thought to be related to a change in the
selectivity between further DEMS research examining the number of active sites for chlorine evolution at that potential.
selectivity between oxygen and chlorine evolution on Makarová et al.78 claimed that this supported the Krishtalik
Ru1−xNixO2−y, where x was between 0.02−0.3, coated on Ti mechanism, in which an oxidation of the surface is the first step in
meshes. Similarly to the study by Jirkovský et al.,76 the coatings the chlorine evolution reaction. The oxygen evolution reaction
were characterized using XRD and energy-dispersive X-ray was thought to be retarded on the mixed Ru−Co oxide, due to
spectroscopy (EDX) and evaluated using DEMS and CV in stabilization of the surface oxygen intermediate by Co.
solutions containing 0.1 M HClO4 and 0.3 M NaCl. This time, Two papers concerning ZnII-substituted RuO2 were published
the coatings were all heat treated at 400 °C for 20 min. The in 2010 and 2011.181,182 The coating was prepared by use of
DEMS study indicated a complicated dependence of the freeze-drying, with subsequent high temperature annealing and
selectivity on the Ni content of the coating. The selectivity was deposition onto Ti meshes. The coating was characterized by use
dependent both on the applied potential and the Ni content. The of DEMS (in 0.1 M HClO4 and 0.15 M NaCl), X-ray absorption
maximum selectivity for chlorine evolution was found at x = 0.1. spectroscopy (XAS), X-ray absorption near-edge spectroscopy
The Tafel slope for chlorine evolution was found to be around 35 (XANES), extended X-ray absorption fine-structure spectrosco-
to 40 mV/dec, while that for oxygen evolution assumed its lowest py (EXAFS), EDX, scanning electron microscopy (SEM), and
value of 40 mV/dec at x = 0.02 and then had a maximum of 100 transmission electron microscopy (TEM). XAS and EXAFS data
mV/dec at x = 0.1. A model was presented to explain the activity, showed that addition of Zn into the structure obstructed the
based on there being four types of active sites, with increasing stacking of Ru coordinatively unsaturated sites (CUSs). On the
numbers of adjacent Ni atoms. Sites with no Ni atoms were basis of refined EXAFS data, it was proposed that Zn gathered in
3001 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

ilmenite defects on the surface, similarly to how Fe impurities in solutions could then be explained by concurrent oxygen
rutile TiO2 gathers. A graphical representation of rutile and formation via both the peroxo route suggested by Rossmeisl et
ilmenite RuO2 and Ru0.9Zn0.1O2 from the article by Petrykin et al.184 and by recombination of oxygen intermediates on crystal
al.182 can be seen in Figure 17. The EXAFS function for ilmenite edges in the ilmenite structures on the surface of the coating. It
was proposed that at x < 0.1, the recombination mechanism
could make up for the lost rutile structure peroxo species
formation.
In two articles published in 2011 and 2012, Chen and co-
workers studied the seawater electrolysis performance of RuO2−
Sb2O5−SnO2 electrodes with and without IrO2 in the coat-
ings.185,186 The Ir-doped coatings were found to be smoother,
and their stability in an accelerated life test was found to be
higher. Current efficiencies were similar for both electrodes.
In 2012, Neodo et al.187 compared the electrolytic behavior of
RuO2−SnO2-coated Ti electrodes with platinized Ti electrodes.
The study was performed in dilute NaCl solutions (0.07 M).
Current efficiencies for active chlorine, O2, chlorate, and
perchlorate were determined. It was found that the current
efficiency for oxygen formation increased with increasing
temperature (tests were performed at 10, 25, and 65 °C),
consistent with a change in activation energy for oxygen-evolving
Figure 17. Graphical representations of rutile and ilmenite RuO2 and reactions. Furthermore, the current efficiency for oxygen
Ru1−xZnxO2 from Petrykin et al.182 It is seen that the ordered formation was slightly higher for the platinized Ti electrode.
arrangement of Ru CUSs (indicated by adsorbed O in the figure) is In the same year, Spasojević et al.60 presented current
obstructed in the ilmenite structure. Reprinted from ref 182. Copyright efficiency and stability studies for a RTO electrode with an
2011 American Chemical Society. overlayer of Pt and rutile IrO2 nanoparticles. The electrode was
prepared on Ti, by first thermally depositing a 10 g/m2 40%
structure Ru1−xZnxO2, based on the structure of ZnTiO3 ilmenite RuO2-60% TiO2 layer, and then thermally depositing a second
but with Ti atoms replaced with Ru atoms in the optimization of layer with a loading of 6 g/m2 from 60% H2PtCl6 and 40% IrCl3
the structure, fit the measured Ru-EXAFS function well. Since no solution. This electrode was compared with an electrode with
ilmenite structures were indicated in XRD or electron diffraction only the first RTO layer. The double-layer electrode was found to
patterns, it was thought that they should be present in small exhibit 50 mV higher overpotentials for oxidation of dilute pH 7
domains distributed in the catalyst crystallites. No information NaCl solutions than the more standard RTO electrode.
was given regarding the stability of Zn in the coating, but Furthermore, the double-layer electrode had a much smoother
Pourbaix diagrams45−47 indicate that ZnII might leach from RuO2 surface and a significantly lower electrochemically active surface
at low pHs. area. The anodic current efficiency for active chlorine formation
DEMS data showed that in solutions without chlorides, the (evaluated based on the gas composition) was some 4% higher
activity of Ru0.9Zn0.1O2 was higher than that of RuO2, based on a for the double-layer coating than for the single-layer RTO
lower onset potential and a higher current density at three coating (see Figure 18). By extrapolating to zero hypochlorite
potentials above the onset potential. However, at higher levels of concentration, it could be concluded that OER is the dominating
Zn addition, the activity started to decrease while staying slightly source for oxygen evolution for both coatings. Furthermore, the
higher than that of pure RuO2. In DEMS measurements in stability, as evaluated by the time before a steep increase in
solutions containing chlorides, the activity for concurrent oxygen electrode potential after galvanostatic electrolysis at j = 10 kAm−2
and chlorine evolution was lower than that of pure RuO2. While a of a 0.5 M H2SO4 solution, was found to be significantly higher
similar maximum of activity for 10% substitution was observed for the double-layer coating than for the single-layer RTO
also in this case, the current density was lower than that of RuO2 coating and several times higher than for a DSA electrode of
for all degrees of substitution and at all applied potentials. conventional composition.
However, the ZnII-substituted coating had a much improved In 2013, Xiong et al.188 presented research on DSA coatings
selectivity for oxygen evolution and the selectivity increased with and Sn−Sb-substituted DSA coatings deposited either on Ti or
increasing substitution by Zn. The increased selectivity was on TiO2 nanotubes (TNT). In the substituted DSA coating, half
rationalized based on the presence of ilmenite structures in the of the Ti content had been replaced by Sn and Sb to obtain
surface of Ru1−xZnxO2. The formation of peroxo groups, found coatings of the nominal structure Ru 0.3Ti0.34Sn0.3Sb0.06O2
to precede the formation of Cl2 on RuO2 according to the work (Ru0.33Ti0.4Sn0.18Sb0.08O2 according to EDX). SEM imaging
by Hansen et al.183 (see the following section), is not favorable showed that the coatings deposited on TNT were smooth, in
since CUSs and bridge sites are distributed randomly on the contrast to the usual mud crack structure of coatings on Ti.
ilmenite structures (see Figure 17). However, this blocks the Determination of the voltammetric surface charge for the
formation of precursor intermediates for both chlorine and coatings, including one coating of overall composition
oxygen evolution. This would conflict with the observation that Ru0.3Ti0.34Sn0.36O2, showed that the Sn−Sb-substituted coating
the activity for oxygen evolution was improved by substitution deposited on TNT had the highest electrochemically active
with Zn in chloride-free solutions. Therefore, Petrykin et al.182 surface area, slightly higher than that of both RTO deposited on
thought that oxygen evolution could take place via crystal edge TNT and Ru0.3Ti0.34Sn0.36O2 deposited on TNT. It was
recombination of surface oxygen intermediates on the ilmenite suggested that the introduction of Sb possibly retards crystallite
structures. The higher activity of Ru0.9Zn0.1O2) in chloride free growth, yielding a higher electrochemically active surface area.
3002 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

reaction barrier for OER, offering a potential explanation for the


trend in selectivity observed for Co- and Mn-doped RuO2.
Later the same year, Krtil and co-workers also examined Co,
Ni, and Zn-doped IrO2 in an analogous way.189 On the basis of
XRD and EXAFS measurements, it was concluded that Zn and
Co were doped into the cationic positions of the rutile lattice,
while Ni formed clusters embedded in the overall structure. The
concomitant evolution of oxygen and chlorine during electrolysis
was characterized using DEMS in experiments where either the
concentration of NaCl or the anode potential was varied.
Increasing the chloride concentration yielded the expected
increase in selectivity for chlorine evolution. The pure IrO2
material showed a higher selectivity for chlorine evolution than
any of the three-doped materials. Whereas the doped materials
showed decreasing Cl2 selectivity at potentials between 1.55 to
1.65 V versus RHE and the highest NaCl concentration of 0.05
Figure 18. Anodic current efficiency for active chlorine (hypochlorous M, the pure material showed 100% Cl2 selectivity up to the
acid and hypochlorite) production, based on measurements of O2 in the maximum 1.65 V versus RHE potential. At 0.01 M NaCl,
cell gas, as a function of hypochlorite concentration, from Spasojević et however, all materials showed similar and decreasing Cl2
al.60 ■ are a conventional 40 mol % RuO2-60 mol % TiO2 coating, while selectivities of around 10−20% at potentials above approximately
● represent the Pt-IrO2 double-layer coating. Experiments were carried 1.6 V versus RHE. At potentials between 1.4 to 1.6 V versus RHE,
out in a 30 g/dm3 NaCl solution at 25 °C and a current density of 1 the Cl2 selectivity of the pure IrO2 material was significantly
kAm−2. Reprinted from ref 60. Copyright 2012 Elsevier. higher than that of the doped materials.
In 2015, Le Luu et al.190 studied the selectivity between Cl2
and O2 evolution in acidic 5 M NaCl solution, comparing IrO2
Linear sweep voltammetry indicated that the onset potential and RuO2 electrodes. The electrochemically active surface areas
difference between Cl2 and O2 evolution was highest for the Sn− of the electrodes, which were prepared by conventional thermal
Sb-substituted coating deposited on TNT, followed by the Sn− decomposition, were determined using CV. The IrO2 electrodes
Sb-substituted coating deposited on Ti, indicating an enhanced were found to exhibit a 40% larger active surface area.
Cl2/O2 selectivity. The increased Cl2 selectivity is thus a result of Electrochemical measurements at pH 2, 4, and 6 showed that
the introduction of the tin−antimony component into the the current efficiency for chlorine evolution was higher for the
coating, rather than an effect of the smoother electrode surface. RuO2 electrodes. The current efficiency for both electrodes
Since previous studies have found Sn-doping to increase the Cl2 decreased as the NaCl concentration was decreased, but the
selectivity, it is possible that this component (and not the Sb RuO2 electrode in all cases exhibited a higher current efficiency.
component) is the main origin of the selectivity increase. The authors suggested surface oxygen deficiency, XPS d-electron
The work of Krtil and co-workers continued in 2013 and 2014, binding energies, differences in point-of-zero charge, electrical
now in cooperation with Rossmeisl and co-workers.82,83 The conductivity, and hydrophilic properties as reasons behind the
focus was now on RuO2 doped with small quantities (up to 20 increased selectivity of the RuO2 electrodes. However, it is
mol %) of Co, Ni, or Zn. In the initial experimental study of Co- perhaps more likely that the difference in selectivity was the result
doped RuO2, the structural and electrochemical properties of this of the significantly lower electrochemically active surface area
mixed oxide were tested using XRD, XAS, EXAFS, and DEMS which the RuO2 electrodes were found to exhibit (see section
experiments examining both chlorine and oxygen evolution in 3.1.3 for further discussion).
0.1 M HClO4 with 0.01 to 0.3 M NaCl. Refinement of the EXAFS In 2014, Spasojevic et al.191 presented a new chlorate
functions indicated that the rutile structure was maintained, with production current efficiency model. The chlorate cell was
Co occupying sites diagonally across the unit cell (i.e., with no modeled as either an ideal tubular reactor or an ideal stirred tank
neighboring Co CUS). The selectivity for oxygen evolution was reactor. The model only included losses due to water oxidation
found to increase with increasing concentration of Co in the (reaction 11) and anodic chlorate formation (reaction 10). A
coating. The relationship between the selectivity trend and the DSA and a Ti plate cathode were used as electrodes. Experiments
structure of the mixed oxide electrocatalyst was examined in the were performed in a pumped laboratory chlorate setup, using a
following study by Halck et al.83 In this study, electrochemical conventional small-volume electrolysis cell connected to a larger
selectivity measurements on Co- and Ni-doped RuO2 electro- holding tank, to get experimental data for the model. The
catalysts were combined with first-principles DFT calculations conditions for the experiments were close to those used in
(see also the following section) based on EXAFS-refined industrial cells. The model was verified by comparing predicted
structures. The calculations showed that the presence of Ni or and measured inlet and outlet active chlorine concentrations, for
Co dopants serve to activate the bridge site oxygen, allowing it to two different inlet concentrations and for electrodes of different
accept a hydrogen adsorbate. This modification of the OER sizes, yielding good agreement. This means that only the loss
mechanism, which on rutile materials involves the formation of reactions of water oxidation and of anodic chlorate formation
an *OOH intermediate adsorbed on, for example, Ru alone, were enough to account for current efficiency losses in a
stabilizes the *OOH intermediate. The binding energy of the laboratory chlorate production setup.
*OOH intermediate is determined by, for example, Ru alone on In the same year, Zeradjanin et al.192 presented experimental
pure rutile materials, but in these doped materials, it can be tuned results for the selectivity between chlorine and oxygen evolution
by both the main constituent cation and by the dopant cation. on Ti−Ru−Ir mixed metal oxide electrodes. DEMS was used to
This stabilization of the *OOH species decreases the overall determine the evolved gas composition. The DEMS measure-
3003 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

ments were performed at 4 kA/m2, 25 °C, pH = 3 and NaCl 2.2.5.2. Research Based on First-Principles Modeling
concentrations varied between 1 and 4 M. Scanning electro- Concerning Chlorine and Oxygen Evolution on Metal Oxides.
chemical microscopy (SECM) was used to study the impact of The number of articles using first-principles approaches to
the microstructure of the electrode on the activity and selectivity understand electrochemical processes has increased substantially
for the two reactions. DEMS indicated that oxygen was evolved since the turn of the century. The majority of studies performed
even at NaCl concentrations of 3 M, over which the current so far fall into the group of computational free energy (or ab
efficiency for chlorine evolution reached almost 100%. EDX initio thermodynamics) studies,194 where only the overall
measurements found local Ru enrichment after electrochemical thermodynamics of the elementary steps of surface reactions
testing. SECM measurements, sensitive only to the evolved Cl2, are characterized. The two articles which laid the groundwork for
found the current distribution over the electrode surface was the first theoretical works on the selectivity between oxygen and
inhomogeneous on the micrometer scale. The SECM tip current chlorine evolution on metal oxides are those of Rossmeisl et al.184
(which is due to the reduction of Cl2 formed at the electrode and Hansen et al.,183 published in 2007 and 2010, respectively.
surface) varied from −500 to −3000 nA, and it was estimated The first article models oxygen evolution on a few metallic
that the local electrode potential at the most highly active site was oxides, while the second one continues the work by also
1.41 V versus AgCl/Ag and at the least active site 1.63 V versus including chlorine evolution in the model. Both studies
AgCl/Ag. It was thus deemed possible that Cl2 could be formed employed the “computational hydrogen electrode” (CHE)
at certain active parts of the electrode, only to be reduced at method of Nørskov et al.195 This method allows results from
other, less active, parts. That this was possible despite the high first-principles calculations in which only adsorption energies of
anodic potentials was explained based on an increase in the Cl2/ intermediates on, for example, metal oxide slabs in vacuum are
Cl− redox potential due to the combination of a high calculated to be used to describe electrochemical reactions in
concentration of formed Cl2 and a low concentration of Cl− solution, by correcting the adsorption energies in vacuum using
(due to concentration limitations). standard thermodynamic relationships and reference data. This
In 2015, a purely experimental study on Mg-doped RuO2 was significantly simplifies the computational expense and complex-
also published.193 From the group of Krtil, the work employed ity of the modeling, as neither the charge transfer process nor the
many of the same techniques as previously (XRD and EXAFS effect of polarization on the binding energies of intermediates
measurements for structural studies and DEMS for electro- have to be modeled explicitly. However, this also means that the
chemical studies). The authors concluded that Mg was unevenly full kinetics of electrochemical processes cannot be captured in
distributed in the overall RuO2 rutile structure, and that the local the model. For this to be possible, methods including explicit
environment surrounding the Mg dopants was ilmenite-like, polarization of the model surface, such as those of Filhol and
similar to what was found for Zn-doped RuO2,182 except for the Neurock,196 Otani and Sugino,197 or Rossmeisl et al.198 have to
lowest dopant concentration. The Mg-doped material was found be used. Nevertheless, the existence of Brønsted−Evans−
to exhibit improved Cl2 selectivity. This effect was most Polanyi (BEP) relationships in heterogeneous catalysis and
significant for the lowest dopant concentration of 5 mol %. electrocatalysis,199−201 which show that the activation energy for
The authors believed that the structure of this Ru0.95Mg0.05O2 an elementary reaction step in general is correlated with the
electrode could be most similar to that of undoped RuO2, reaction free energy of the same step, does indicate that the
indicating that a significant improvement in selectivity could conclusions from computational free-energy studies give a
occur when Ru is replaced by Mg in the rutile lattice. Also higher qualitatively correct description of the full thermodynamics
concentrations of Mg dopant resulted in higher chlorine and kinetics of surface reactions. Indeed, the successful
selectivity, but the rationale given was then the broken CUS predictions of more active electrocatalysts for a variety of
stacking in the ilmenite structure.182 reactions that have resulted from computational free-energy
In the same year, Macounová et al.84 studied anodic oxidation studies support this conclusion.202,203
of hypochlorite in 0.1 M NaClO4, at pH values of at least 9.5, The computational free-energy study of Rossmeisl et al.184
using DEMS and a Pt mesh electrode. It was suggested that, at an used DFT calculations to calculate the stability of intermediates
anodically polarized Pt electrode, hypochlorite is decomposed in in the oxygen evolution reaction on the (1 1 0) surface of RuO2,
a radical reaction involving formation of ClO•. Once this species IrO2, and TiO2. The (110) surface of rutile oxides has two
has formed, it was suggested that it should react with OH− to different sites on which reactions can take place, bridge sites or
form the superoxide radical •OO−, which could either react with CUSs. The nature of the (1 1 0) surface of RuO2 was studied
water to form H2O2 and O2 or with HOCl to form O2 and OH·. previously by Over et al.,204 and the importance of CUS for the
The latter radical was suggested to regenerate ClO−, closing a activity of oxide materials was conjectured already by, among
catalytic loop resulting in conversion of water into O2. On the others, Mars and van Krevelen.205 A model of the (1 1 0) surface
basis of the electrical charge transferred per mol of O2 generated of RuO2 is found in Figure 19. Rossmeisl et al.184 calculated the
on the Pt electrode, which was lower than expected during OER, adsorption free energies of all intermediates that could result in
it was suggested that the reaction via ClO•, •OO−, and OH• was oxygen evolution. The calculations showed that oxygen
occurring on the electrode. This pathway does not involve evolution on rutile oxides occurs via the following steps, where
formation of H2O2 and results in accumulation of Cl− in the HOc, Oc, and HOOc are adsorbates bound to the CUS:
system. However, H2O2 was detected, suggesting that also the
H 2O → HOc + H+ + e− (32)
reaction between •OO− and water was ongoing. Furthermore,
measurements of chloride concentrations before and after an
experiment did not indicate accumulation of Cl− and none of the HOc → Oc + H+ + e− (33)
suggested radical species were detected experimentally. Whether
the suggested radical mechanism is relevant for industrial chlor- Oc + H 2O → HOOc + H+ + e− (34)
alkali or sodium chlorate production using DSAs remains an
open question. HOOc → O2 + H+ + e− (35)

3004 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Figure 19. A model of the (1 1 0) surface of RuO2. Ruthenium cations


are green, while oxygen anions are red. The Ru CUSs and the O bridge
sites (binding to Ru bridge sites) are indicated with arrows.

Rossmeisl et al. found oxygen evolution through recombination


of two oxygen atoms (a Volmer−Tafelmechanism) to have a too Figure 20. Volcano plot for oxide materials obtained by Rossmeisl et
high activation energy to be plausible, based on previous DFT al.184 The different lines represent the negative of the free energy of
calculations by the same group.199 This means that the traditional reaction of each step. These lines each limit the reaction rate, based on
view of OER on rutile oxides as being dependent on O−O the binding energy of oxygen on the surface of the three oxides and the
recombination on neighboring Ru CUS67,68 is likely incorrect. In surface coverage. The most active type (based on surface coverage) of
the case of RuO2 and IrO2, the largest ΔG was found for the third each oxide is included as ● in the diagram. Reprinted with permission
step in the mechanism above, the decomposition of the second from ref 184. Copyright 2007 Elsevier.
water molecule. Therefore, for RuO2 or IrO2, the calculated ΔG
for this third step was used as the measure of the overpotential.
The overpotential (ΔGstep 3 − ΔG0 = ΔGstep 2 −1.23 V) was potentials above 1.4 V versus SHE. However, Fang and Liu206
determined to be 0.37 V for RuO2 and 0.56 V for IrO2. In the case found that a mixed HObr-Obr phase was slightly more stable than
of TiO2, the step associated with the highest change in free a fully deprotonated RuO2 surface at potentials up to about 1.6 V
energy was the second one, formation of O* from HO* on the versus SHE, enabling the HObr-stabilized pathway. Interestingly,
surface. The overpotential, based on ΔG for this step, was 1.19 V. in 2014, a type of Obr stabilization was suggested by Halck et al.83
The adsorption energies for the intermediates (Oc, HOc, and to explain the improved OER activity of Ni- and Co-doped RuO2.
HOOc) for different levels of surface coverage of Oc and HOc on In that case, it was suggested that the dopant cations
the surfaces of the oxides were also calculated. A linear predominantly occupied the cationic bridge sites, allowing the
relationship, a scaling relation, was found between the adsorption Obr sites to stabilize the HOObr intermediate in the mechanism of
energy for O on the surface and the adsorption energy of HO or Rossmeisl et al.184
HOO. In other words, surfaces that bound O strongly also bound In 2010, Hansen et al.183 applied DFT calculations to treat
HO and HOO strongly. This also implied that, for all three oxide chlorine evolution at rutile (1 1 0) surfaces. The transition metal
materials treated, the rate of the oxygen evolution reaction is oxide surfaces considered were IrO2, RuO2, PtO2, and TiO2. It
given by the adsorption energy of oxygen on the surface. The was found that the bridge sites were covered by oxygen for most
scaling relation was then applied in the calculation of the free conditions, so that the CUSs were active in OER while, for
energy of reaction for the different steps of the reaction. This example, in the case of RuO2, O-covered Ru CUSs were active for
yielded a volcano plot, as shown in Figure 20. The volcano plot ClER. The energy of adsorption of the different intermediates
qualitatively agreed with experimental data (mainly that (Clc, OHc, Oc, Oc2, Cl(Oc)2, and ClOc) were then calculated, and
described by Trasatti27,156). a number of Pourbaix diagrams were determined. The Pourbaix
The topic of the present review is the selectivity between diagram for RuO2 is reproduced in Figure 21. In each area of the
chlorine (or chlorate) formation and oxygen evolution, and thus Pourbaix diagram, the predominant species is the one which has
all studies on ways to improve the OER on metallic oxide the most negative free energy of adsorption versus a suitable
materials will not be included here. However, it is interesting to reference state (e.g., gas phase Cl2 in the case of the Cl
compare the results of the study of Rossmeisl et al.184 with the adsorbate). The Pourbaix diagrams indicated under which
later theoretical study of Fang and Liu.206 In 2010, Fang and conditions chlorine or oxygen evolution would be more
Liu206 studied the OER on RuO2 (110), now also calculating the probable. Furthermore, as was the case for oxygen evolution,184
kinetic barriers for both the Volmer−Tafel pathway (recombi- it was found that the binding energy of the intermediates
nation of two Oc) and a water dissociation pathway, similar to (including the Cl-containing species) all scaled with ΔE(Oc)
that suggested by Rossmeisl et al.184 The study represents a step (the adsorption energy of Oc). Application of the scaling
forward as the calculation of kinetic barriers allows for an explicit relations allowed a general phase diagram, where U and ΔE(Oc)
account for both reaction thermodynamics and kinetics. Fang decide which surface state is most stable, to be constructed (see
and Liu206 found, in agreement with the assumption of Rossmeisl Figure 22). This phase diagram is correct for all pure rutile oxide
et al.,184 that the Volmer−Tafel pathway was less energetically (110) surfaces, as long as all relevant intermediates have been
favorable than the water dissociation mechanism. A difference considered. The diagram was constructed for pH = 0 and aCl− = 1,
between the mechanisms of Rossmeisl et al.184 and Fang and to make the diagram relevant at conditions where chlorine gas
Liu206 resided in whether the Obr sites were protonated. In both forms, but it can be adapted for other conditions as well. Only the
cases, a protonated bridge site was found to reduce the areas which allow formation of stable chloride-containing
thermodynamic reaction barrier, but Rossmeisl et al.184 intermediates are likely as chlorine evolution sites, since their
concluded that bridge sites will be deprotonated at anodic stability likely makes the sites abundant on a real electrocatalyst.
3005 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Cl(Oc )2 + Cl−(aq) → 2Oc + Cl 2(g) + e− (41)


c c
for the Cl(O )2 intermediate. In these reactions, signifies a bare
CUS and Occ 2 is an adsorbate with molecular oxygen-like
geometry bound to two neighboring CUSs. For each reaction, an
expression for the electrode potential where all steps are
spontaneous, as a function of ΔE(Oc), can be written and
entered into the general Pourbaix diagram (Figure 22). The
resulting diagram is shown in Figure 23, which contains the

Figure 21. Phase diagram calculated by Hansen et al.183 The only area
(at the pH and potentials included in the diagram) where the surface has
a stable Cl-containing intermediate is at U > 1.55 V at pH < 1.3. Since the
only stable chlorine intermediate is Cl(Oc)2, Hansen et al. believe that
the reaction mechanism that yields Cl2 on real RuO2 surfaces includes Figure 23. General phase diagram, with thick lines indicating the
the Cl(Oc)2 intermediate. Reproduced with permission from ref 183. “volcano” for chlorine evolution activity, taken from Hansen et al.183
Copyright 2010 PCCP Owner Societies. Actual oxide materials are indicated as red dots with error bars. Dotted
lines are the lines for the chlorine evolution reaction involving each
chloride surface species (ClOc, Cl(Oc)2, or Clc). The volcano is limited
by the reaction that requires the lowest applied potential to occur, for
each ΔE(Oc). It is also limited by predominance areas where chloride
containing compounds are stable, and therefore common, on the
surface. The horizontal line is the equilibrium potential for chlorine
evolution. It is seen that RuO2 has an activity that is close to the optimal
one in the volcano plot. The blue dashed line is the line for oxygen
evolution. It is seen that the potential necessary for chlorine evolution is
always lower than the potential necessary for oxygen evolution.
Reproduced with permission from ref 183. Copyright 2010 PCCP
Owner Societies.

Figure 22. General phase diagram, with the dominant surface volcano plots. The volcanoes are limited both by the areas where
compounds indicated for each predominance area, of Hansen et al.183 chloride-containing intermediates are prevalent and where the
The background is a visual representation of the surface and the chlorine evolution reaction involving these intermediates is
compounds, where metal ions are blue, oxygen is red, hydrogen is white spontaneous. Hansen et al. conclude their work by supposing,
and chlorine is green. Reproduced with permission from ref 183. based on Figure 23, that there is room to find an oxide material
Copyright 2010 PCCP Owner Societies. which has an optimal ΔE(Oc), at which the ClER would progress
with a lower applied potential and with a higher selectivity for
Three reaction mechanisms, one for each of the stable surface chlorine evolution. No similar analysis at conditions for chlorate
chloride compounds (ClOc, Cl(Oc)2, or Clc), were proposed production has been published so far.
based on the calculated adsorption energies. The effects of In 2014, Exner et al.207 performed a complete reevaluation of
applied potential, pH, chloride concentration, and electrode the theoretical Pourbaix diagram for chlor-oxy species on the
material were considered, using the computational hydrogen (110) surface of RuO2. All relevant intermediates were
electrode (CHE) model,195 when determining under which considered. The Pourbaix diagram produced, shown in Figure
conditions each of the three reaction mechanisms would become 24, mostly agreed with that of Hansen et al.,183 even though
spontaneous. The reaction mechanisms are the following Exner et al. performed their study using the PBE exchange-
correlation functional instead of RPBE. However, a key
Oc + Cl−(aq) → ClOc + e− (36) difference was that Exner et al.207 did not find that the Occ 2
(oxygen molecule-like) adsorbate was the most stable adsorbate
ClOc + Cl−(aq) → Oc + Cl 2(g) + e− (37) on the surface of RuO2 at applied potentials and pH below those
c
for the ClO intermediate, required for ClER or OER, instead identifying 2 Oc as the most
stable configuration. Still, the intermediates present at pH and
Cl−(aq) +c → Clc + e− (38) potentials required for both ClER and OER on RuO2 were found
c − c − to be the same in both the study of Exner et al.207 and the study of
Cl + Cl (aq) → Cl 2(g) + +e (39) Hansen et al.183 Furthermore, the difference in stability between
c
for the Cl intermediate, and the Occ c
2 and 2 O was within the errors expected from GGA-level
− − DFT calculations. Nevertheless, the results of Exner et al.207
Occ c
2 + Cl (aq) → Cl(O )2 + e (40) agreed with the GGA-level DFT study of Wang et al.,208 who also
3006 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Exner et al.210 continued the theoretical study the year after, by


considering the effect on the chlorine evolution activity of
applying overlayers of other metal oxides on RuO2 (110). A
significant finding was that deposition of a monolayer of PtO2 on
RuO2 (110) should yield increased activity for ClER. Addition-
ally, in 2015, Exner et al.211 reported results on the effects of
doping another transition metal (either Cr, Ir, Mn, V, Au, Rh, Co,
or Pd) in either the active CUS or the cationic bridge site next to
the active site in RuO2. It was found that either doping Ir into the
active site, or doping Cr, Ir, Mn, Zr, or V into the neighboring
cationic site, could improve the activity for OER. However, in
neither study was the selectivity between OER and ClER
considered.
Later in 2014, two theoretical studies of chlorine evolution
Figure 24. Theoretical Pourbaix diagram for OER and ClER on RuO2 selectivity, a study focusing on RTO by Karlsson et al.64 and a
(110) according to Exner et al.207 (excluding solvation effects). The study focusing on monolayers of TiO2 on RuO2, by Exner et al.,63
areas in which ClER and OER can occur are marked in the figure. The were published. The study of Karlsson et al.64 was based on the
conditions (aCl− = 1 and T = 298 K) are the same as those in Figure 21, previous volcano plot of Hansen et al.183 The adsorption energy
the Pourbaix diagram according to Hansen et al.183 The diagram is
similar to that of Hansen et al.183 in most respects, but the key difference
of atomic oxygen on the CUS of TiO2, RuO2, and both Ru-doped
resides in Exner et al.207 finding that Occ TiO2 and Ti-doped RuO2, as well as on a structure with the same
2 is unstable on the surface and
that both CUSs are covered by Oc at potentials and pH values below overall stoichiometry as that of conventional DSA (30% RuO2
those where ClER and OER can occur. Reprinted with permission from and 70% TiO2) were studied (see Figure 25). The adsorption
ref 207. Copyright 2014 Elsevier. energy of oxygen had been found by Hansen et al.183 to serve as a
descriptor for both OER and ClER on rutile oxides. The
found that the Occ approach used by Karlsson et al.64 allowed the effect of the local
2 adsorbate on RuO2 should dissociate into 2
Oc. The authors also took a step further by including the effect of structure of an electrode (i.e., the position of the dopant relative
solvation, by use of a continuum solvent model. This had a clear to the surface-active CUS) on the electrocatalytic properties of
effect on the Pourbaix diagram and also made the 2 Oc the mixed oxide to be studied. It was found that while doping
configuration significantly more stable than the Occ 2 config-
RuO2 with Ti has a relatively minor effect, doping TiO2 with Ru
uration. However, the effects of solvation on the adsorption serves to activate Ti CUSs. These activated Ti sites had oxygen
energies of the different intermediates were found to be much adsorption descriptor values in the range optimal for highly
larger than what was found in the first-principles studies of both selective and active chlorine evolution. If the dopant Ru atom was
Rossmeisl et al.184 and a later more thorough study by situated in the CUS, the descriptor value was the same as for pure
Siahrostami and Vojvodic.209 The latter calculation, in which RuO2, indicating that the properties of Ru were maintained even
water was considered explicitly, found that the effect of up to 2.5 in the mixed oxide. These results are shown in Figure 26. The
monolayers water molecules on the adsorption energies of the study indicated that the improvement in ClER selectivity that
OER intermediates (HOc, Oc, and HOOc) is minor for RuO2 RTO electrodes exhibit, when compared with electrodes of pure
(110).209 However, the effects of water molecules on the RuO2, is related to activation of surface Ti atoms by nearby Ru
adsorption energies of chlorine species was not considered, dopant atoms. The study also indicated that even electrodes with
warranting future study of the effects of solvation on such species. monolayers of TiO2 on RuO2 should exhibit improved selectivity

Figure 25. Model structures considered by Karlsson et al.64 (a) is a side view of the structure of Ru-doped TiO2, (b) indicates is a side-view of the
structure of the DSA model system, and (c) is a three-dimensional image of the surface formed by the system indicated in (a). In (a and b), the surface
normal is in the z direction, upward in the figure. Reprinted with permission from ref 64. Copyright 2014 Elsevier.

3007 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Figure 26. Trends in oxygen adsorption energy descriptor values as the position of the dopant atom is changed, indicating an activation of Ti sites by
dopant Ru atoms. This could explain the improved selectivity for ClER that RTO electrodes exhibit. Reprinted with permission from ref 64. Copyright
2014 Elsevier.

Figure 27. Results of doping rutile TiO2 with other dopants, in either the CUS (left side) or the bridge site (right side). The vertical lines correspond to
descriptor values for pure TiO2 (thin green), optimal selectivity and activity for ClER (thick red), maximum OER activity (thick blue), and for pure
RuO2 (black). Reprinted with permission from ref 65. Copyright 2015 Elsevier.

for ClER, while maintaining a similarly high activity as pure Pourbaix diagrams and volcano plots and used two different
RuO2. Interestingly, the study of Exner et al.,63 building upon the exchange-correlation functionals and assumptions related to
results from their previous study,207 also concluded that a effect of the electrolyte on the reaction. The reason for the
monolayer of TiO2 on RuO2 should optimize the selectivity for improvement in selectivity is the decrease in Oc adsorption
ClER. Thus, the two studies were in basic agreement in spite of energy that occurs when applying a monolayer of TiO2 on
being based on two independent sets of calculated surface RuO2.63,64
3008 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Figure 28. Pourbaix diagram for Ru in concentrated chloride solution, from Loučka.214 Original caption reads: “Potential-pH diagram for the Ru−
H2O−Cl−, system at 25 °C in a solution of 264 g/dm3 NaCl. Activities of other ions are 1 × 10−6 M”. Reprinted with permission from ref 214. Copyright
1990 Springer Science and Business Media.

In a publication from 2015, Karlsson et al.65 expanded their works that have clarified the relationship between selectivity for
theoretical studies of doped rutile TiO2.65 The method was Cl2 or ClO−3 and O2 and the stability of the electrode material will
analogous to the one used in Karlsson et al.,64 but now included be considered. As will be seen, the decomposition of RTO
38 dopants, mainly transition metals. A general study of the effect electrodes is connected with the removal of Ru from the
of the dopant either as active site, or located close to a Ti site, was electrode in reactions related to the OER. For this reason, studies
performed, suggesting a number of possible dopants that would of both Ru metal, RuO2, and mixed Ru oxides will be considered
activate TiO2 either for OER or ClER. The results are shown in in this section. The discussion will start from the Pourbaix
Figure 27. Arsenic was suggested as a dopant that would result in diagram for Ru, which was published in 1966,213 and the review
maximal ClER activity regardless of position. The dopants Bi, Co, will then discuss papers published from 1966 and onward.
Ir, Mn, Pd, Ru, or V were predicted to result in a doped oxide The Pourbaix diagram for Ru213 indicates the regions of
with optimal ClER selectivity and activity, while doping with Mo stability of the pure metal (see Figure 28 for the Pourbaix
or Re was predicted to result in an oxide with high OER activity. diagram for Ru in chloride solutions of Loučka214). The
However, it was found that the local structure of the active site is conversion from Ru, to Ru(OH)3, RuO2, and finally volatile
key to achieving these effects. The properties of some ternary RuO4 at potentials above 1.5 V versus SHE at pH 0 is central to
oxides were also studied, finding that the combined effect of two the understanding of the stability of Ru and its oxides. The
dopants located next to a Ti-active site was close to the average potential required for conversion of RuO2 to form RuO4
effect of the dopant in their respective binary Ti oxides. decreases with increasing pH, until the final product of the
Furthermore, it was found that under OER conditions, but not oxidation of RuO2 becomes RuO−4 and then RuO2− 4 at pH > 9−
under ClER conditions, there is a driving force for most dopants 12.
to segregate toward the active CUS. One of the highest driving Around the same time, Llopis and Vázquez215 and Llopis et
forces was noted for Ru-doped TiO2, indicating that the al. studied the corrosion of Ru. In the first study, anodic and
216

deactivation of Ru-based DSAs might occur through segregation cathodic charging curves for Ru in HClO4 and HCl solutions
of Ru toward the active site, followed by formation of RuO4.212 were measured. On the basis of these results, the electrode
However, the process after segregation toward the active site potentials where oxidation of Ru to form Ru2O3, RuO2, and
(e.g., formation of RuO4) was not studied in the paper. RuO4 occurs were found. In Llopis et al.216 a radiochemical
technique, which allowed for the detection of very small amounts
2.3. Electrochemical Studies Concerning the Relationship of dissolved Ru, was applied to study the potentiostatic corrosion
between Selectivity and Stability
of Ru in 2 or 4 M HCl, at 50 °C. The technique was based on
There are many studies of Ru metal oxide degradation and detecting a radioactive Ru isotope, which is formed after neutron
decomposition. A complete discussion of this literature would irradiation of the electrodes in a nuclear reactor. The potential in
necessitate a full review. Therefore, in this section a selection of the electrochemical measurements was kept constant at a
3009 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

number of potentials in the range from 0.95 to 1.25 V versus SCE that Ru oxides form on the surface of the electrode during OER
during more than 10 h. The Coulombic corrosion yield (the and that ruthenates are released into the solution.
percentage of the current resulting in the corrosion of Ru) In 1978, Uzbekov et al.221 presented a radiochemical study of
decreased sharply as the potential was increased above 1.0 V decomposition of RTO electrodes prepared using thermal
versus SCE. At these potentials, a passivating Ru surface oxide decomposition to achieve the standard composition of 30%
grows on the metal, decreasing the rate of Ru cation transport RuO2 and 70% TiO2. The electrodes were polarized anodically in
through the film. At potentials approaching 1.4 V, formation of 5 M NaCl, kept at pH 7, at several potentials during 10 to 80 h.
RuO4 started.213,216 Initially, measurements were conducted at an anodic current
In 1973 and 1976, Gorodetskii et al.217 and Pecherskii et al.,218 density of 2 kA/m2 and an anode potential of 1.34 V versus SHE.
respectively, presented further work on the anodic behavior of The authors found that the corrosion rate, as measured by the
Ru. The studies examined the anodic dissolution behavior of Ru radioactivity of dissolved Ru in the solution measured during the
in the potential range from 0.8 to 1.4 V versus SHE. A experiment, decreased from more than 10 × 10−6 g/cm2 h to
radiochemical technique was again used to determine the around 10 × 10−7 g/cm2 h over the first few hours and then
amount of dissolved Ru. In the first study, the anodic behavior of steadily decreased further over time. Although the solution was
the electrode was studied in both 3 M HClO4, where oxygen continually stirred, the radioactivity measured when draining the
evolution is the dominating anodic process, and in 3 M HCl, cell after the end of the experiment was higher than that
where chlorine evolution is the dominating anodic process. In the measured in the circulating solution. This indicated to the
former solution, the total current density as well as the rate of Ru authors that a part (10 to 40%) of the degradation of the anode
dissolution increased as the potential was increased above 1.2 V took place by detachment of solid particles from the surface. Ru
versus SHE. The slopes for both processes were equal, at about was also dissolved in the electrolyte, indicated by the deposition
30 mV per decade in either corrosion rate, measured in g/cm2 s of Ru on the counter electrodes and released into the gas phase as
or current density. The authors concluded that the rate of the RuO4. It was estimated that 5−15% of the Ru was released into
OER was connected with the rate of Ru dissolution, possibly due solution and that 10% was released into the gas phase.
to the existence of shared surface intermediates. However, in 3 M Furthermore, it was noted that the corrosion rate increased 2−
HCl at the same potentials, although the rate of Ru dissolution 5 times upon current interruption of even a few seconds. After
did increase with increasing electrode potential, the slope of the several hours of electrolysis, the corrosion rate again decreased to
Ru dissolution rate was significantly lower than that of the total a low value. Finally, the authors examined corrosion rates at
current density. It appeared that during chlorine evolution, several different anodic potentials. A significant increase in the
increasing current density resulted in a decreasing current corrosion rate was noted at potentials above 1.3 V versus SHE.
efficiency for Ru dissolution. Furthermore, Gorodetskii et al. Increasing the anode potential increased the rate of corrosion,
studied the effect of increasing Cl− concentration from 0.01 to 1 but the specific rate (per mole Cl2 produced) decreased.
M at an anode potential of 1.4 V versus SHE, finding a 4-fold In the same year, Kokoulina et al.222 presented polarization
decrease in the Ru dissolution rate together with an increase in measurements on conventional RTO anodes measured in 1 N
overall current density by a factor of 300. The authors suggested H2SO4 at different pH in the range of 0.5−13 and temperatures
that electrodes, specifically RTO, with increased selectivity for between 20 and 80 °C. An increase in the Tafel slope was noted
the ClER thus should also exhibit improved corrosion resistance. at 1.5 V versus RHE at 20 °C and 1.4 V versus RHE at 80 °C. The
Pecherskii et al.218 performed a deeper study of the effect of pH potential at which this bend occurs was termed the critical
on the Ru dissolution rate. The overall anodic current density potential (a term first introduced by Veselovskaya et al.219) at
and the Ru dissolution rate were determined in solutions of which the change in Tafel slope might signal an increase in
HClO4, H2SO4, or HCl at a constant ionic strength of 3 M and corrosion rate due to formation of RuO4. Increasing the pH
acid concentrations from 0.01 to 3 M. In the sulfate and decreased the critical potential, which would agree with the
perchlorate solutions, both the overall current density and the Ru Pourbaix diagram.213
dissolution rate increased with increasing pH, again indicating Also in 1978, results based on anodic polarization measure-
the connection between Ru dissolution and the OER. In the HCl ments conducted in buffered sulfuric acid solutions using pure
solution, the overall current density decreased slightly with RuO2/Ti electrodes was presented by Bondar and Kalinovskii.223
increasing pH, and the Ru dissolution rate increased. The authors determined the critical current density (the current
In 1974, Veselovskaya et al.219 published a study of the anodic density at the critical potential,219,222 at which the change in Tafel
behavior of metallic Ru including measurements conducted in a slope of the polarization curve occurs) as a function of pH at 20
number of solutions, including 300 g/dm3 NaCl at pH 3. All and 60 °C. This indicated the current densities which should not
measurements were performed at 80 °C. The authors noted that be exceeded for RuO2 electrodes, at various pH values. As an
the regions of Ru dissolution at lower anodic potentials (below 1 example, the critical current density for oxygen evolution on
V vs SHE) are followed by regions of lower Ru dissolution rate RuO2 in sulfuric acid at pH 0 was approximately 1 kA/m2.
(due to the formation of hydrous RuO2) at potentials above 1 V In 1979, Gorodetskii et al.224 examined the effect of current
versus SHE. The Tafel slope for electrolysis in the chloride density on the corrosion rate of DSA. Again, electrodes with the
solution at these higher potentials coincided with the typical 40 standard 30:70 Ru:Ti composition were prepared by thermal
mV/dec Tafel slope for ClER on a DSA electrode, supporting the decomposition. Electrolysis was performed at 87 °C in 5 M NaCl
idea that a hydrous RuO2 layer covers the surface under these at a pH of around 2, and the corrosion rate of Ru was determined
conditions. The paper is perhaps most notable due to its using a radiochemical method. Results for one measurement
introduction of the “critical anode potential”, Ecr, here given as carried out at a current density of 2 kA/m2, giving an anode
approximately 1.45 V versus SHE, as the potential where potential of 1.33 V versus SHE, was shown. As in Uzbekov et
conversion of RuO2 into volatile RuO4 becomes favorable. al.,221 the corrosion rate decreased slowly over several hours and
In 1977, Iwakura et al.220 presented research on possible increased for a short time after shutdowns when the electrode
mechanisms for OER and for Ru dissolution from Ru, suggesting had been removed from the solution. It was estimated that 14%
3010 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

of the corrosion was due to detachment of particles and that 12% of particles, independent of the pH value of the solution. The
was due to formation of gas phase RuO4 (apparently being results indicated that the connection between the OER and the
produced at an anode potential below 1.4 V vs SHE). A rate of corrosion is not necessarily as simple as was previously
comparison between the corrosion rate of pure RuO2 and of the indicated. At pH below zero, the corrosion rate was said to
RTO electrodes indicated that the corrosion rate of the pure Ru decrease significantly when increasing the pH up to 0. In the pH
oxide was 1 order of magnitude higher than that of the RTO range from 0 to 3, the rate of corrosion of the DSA was essentially
material. The total corrosion rate of the RTO anode was also constant at a value of about 1 × 10−8 g/cm2 h in the concentrated
shown as a function of the current density, once more indicating NaCl solution. The corrosion rate was somewhat higher in the
that the rate of RTO degradation increases with increasing 100 g/dm3 NaCl solution, here showing a minimum at about pH
current density but that the specific (per mol Cl2 produced) 1, and then increased with increasing pH for both concentrations.
corrosion rate decreased with increasing current density. However, the OER rate showed a clear increase when increasing
In 1980, Barral et al.225 published results on the deactivation of the pH from 0 to 2, and then again when increasing the pH from
RuO2 and RTO during OER in 1 M H2SO4 at 40 °C. The 4 to 5, in the 300 g/dm3 NaCl solution. Increases in the OER rate
deactivation was determined indirectly, by measuring the time was also noted when increasing the pH from 2 to 4 in the 100 g/
required to reach an anodic potential of 4 V versus the saturated dm3 NaCl solution. The results thus indicate a region of maximal
mercury sulfate electrode. Interestingly, it was found that pure stability around pH 0 to 3 in the 300 g/dm3 NaCl solution at
RuO2 electrodes had a longer lifetime than RTO electrodes at 10 industrially relevant current densities. Furthermore, the total
kA/m2. However, the electrodes had been prepared by thermal percentage of Ru released as gaseous RuO4 was determined as a
oxidation at 500 °C for only 10 min with no final prolonged function of pH at both NaCl concentrations. In both cases, at a

heating. This is significantly shorter than the conventional time current density of 2 kAm 2 and anode potential of 1.33 V versus
of about 1 h for the final oxidation of the coating. Indeed, Gerrard SHE, the percentage released as RuO4 increased significantly
and Steele11 note that for pure RuO2, about 5 min is needed to when increasing the pH, from about 10% at pH 0−1 to about
achieve a stable value for the electrical conductivity, but an oxide 60% at pH 4 in the 100 g/dm3 NaCl solution and 40% at pH 5 in
coating of 30:70 Ru:Ti needed more than 1 h to reach a stable the 300 g/dm3 NaCl solution.
conductivity. Therefore, it is possible that the results of Barral et The work of Pecherskii et al.67 from 1982 has already been
al.225 disagree with the majority of other work since their mixed described in section 2.2.3, but the corrosion results will here be
oxide coatings were not sufficiently stabilized during preparation. discussed in more detail. Again, the study in question examined
Still, they noted a trend for the pH dependence of the the properties of three types of electrodeposited coatings on Ti
deactivation of pure RuO2 that seems to contradict that of substrates. The first type consisted of Ru only, the second of Ru
Bondar and Kalinovskii.223 Barral et al. found that the lifetime which was then heat-treated for several hours, and a third which
was essentially constant in the range of pH from 0 to 13 at a was treated the same way as the second but was then polarized
current density of 5 kA/m2. At higher pH, the lifetime increased anodically in sulfate solution until the electrode potential
significantly. The authors suggested that the reason for the increased sharply. The first electrode could thus be considered
constancy in lifetime is related to the difference between bulk and to consist of a Ru hydroxide or metal, the second of RuO2, and
anode surface pH. Because of the ongoing OER, the pH at the the third of RuO2 mixed with TiO2. Auger electron spectroscopy
anode is lower than in the bulk. Only at very large pH values (e.g., was performed, indicating that Ti was present in the surface layer

a bulk pH > 11.5 at a current density of 100 Am 2), does this of the third type of electrode but not in the second type. The
change, they argue. Nevertheless, the fact that the corrosion rate same trend as was noted for the oxygen production rate was seen
was not determined directly by measurement of the corrosion also for the corrosion rate. Electrodes of the third type not only
products limits the utility of the study. Determining the lifetime produced less oxygen but also exhibited significantly lower
based only on an increase in anode potential means that the corrosion rates than the second type. The corrosion rate at 1 kA/
deciding factor for the lifetime might be the formation of a m2 was about 1 × 10−5 g/cm2 h for the first type of electrode,
nonconductive interlayer between the Ti substrate and the about 1 × 10−6 g/cm2 h for the second type of electrode, and
coating, rather than decomposition of the coating itself. It is about 1 × 10−9 g/cm2 h for the third type of electrode. The pH
therefore difficult to disentangle the various mechanisms of dependence of the corrosion was also studied. Electrodes of the
deactivation when using only the potential as the deciding factor. third type exhibited the behavior seen by Gorodetskii et al.,151 of
Still, it will be indicated much later in the works of Tilak et al.226 essentially constant corrosion rate up to pH 3, followed by an
and Eberil’ et al.227 that the appearance of an insulating TiO2 increase in corrosion rate. Corrosion rates increased with anode
layer might not be an important mechanism for DSA potential at potentials positive of 1.3 V versus SHE for all three
deactivation. types of electrode. The slopes of the increase in corrosion rate
In 1981, Gorodetskii et al.151 published a study of the effect of with electrode potential was similar to the Tafel slope of the
pH in the range from 0 to 5 on both the corrosion rate and on the oxygen evolution reaction.
rate of the OER on DSA. Some corrosion measurements were In 1982, Burke et al.228 presented results for OER using RuO2.
also carried out at lower pH values. Again, standard DSA As in the study of Barral et al.,225 the lifetime was estimated based
electrodes were studied at 87 °C in 300 g/dm3 or 100 g/dm3 only on a steep increase in anode potential. Nevertheless, it was
NaCl. Once more, electrolysis was carried out at a current density again seen that increasing current density results in a significant
of 2 kA/m2. The same radiochemical method as used previously decrease in lifetime in both acid and base. The lifetime of the
was employed. However, the O2 formed was now quantified by anode decreased significantly with acid concentration in H2SO4
GC. Corrosion measurements were started after the system had solutions, while in alkaline solution of NaOH the lifetime
reached steady state (after about 200 h), and measurements with increased significantly with increase in concentration above 1 M
different pH and NaCl concentration were then performed. NaOH. Mechanisms for both OER and Ru corrosion were
Again, about 14% of the corrosion was the result of detachment suggested.
3011 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

In 1983, Denton et al.154 presented results from radiochemical current density of 2 kA/m2 at 60 °C. Corrosion rate
experiments using RTO electrodes used for chlorine evolution in measurements were conducted using the same radiochemical
either membrane or diaphragm cells, finding a direct relationship method as previously applied. The effect on the anode potential
between the rate of Ru loss from the coating and the oxygen and corrosion rate of changing the electrolyte concentration
percentage in the cell gas. It was also claimed that increased pH of from 5 M NaCl to 6 M NaClO3 was examined, finding a clear
the anolyte resulted in increased anode wear, that increased cell increase in anode potential as the NaCl concentration was
temperature or current density resulted in increased anode wear, decreased. At the same time, the corrosion rate increased. The
and that increased chloride concentration was associated with a mechanism involved is likely the same as has been seen in several
decrease in anode wear rate. other studies, that the rate of OER increases significantly when
In 1984, Kötz et al.212 contributed a detailed study of the the Cl− concentration decreases, resulting in an increased
corrosion of RuO2, in which CV was combined with in situ corrosion rate.
differential reflectance spectroscopy, finding that during OER in In 1986, Kötz and Stucki232 presented results on stability of
H2SO4, RuO4 is the only corrosion product. Furthermore, the mixed Ir−Ru rutile oxide coatings. While the corrosion rate was
release of RuO4 started at the same potential as the OER. This only determined based on the standard accelerated lifetime test
study clearly shows the mechanistic link between the OER and (as determined by an increase in the anode potential) in H2SO4,
the corrosion of RuO2. the authors indicated that mixing RuO2 and IrO2 results in
In 1985, Uzbekov and Klement’eva229 published a study stabilization of the material, although at the cost of a decrease in
employing radioactive isotopes of Ru and Ti in the salts used to activity for OER.
prepared a DSA coating, so that both Ru and Ti corrosion In the same year, Kolotyrkin233 presented a review of the
products (regardless of oxidation state) could be detected during relationship between corrosion and kinetics for anodes. For
electrolysis. The experiments were conventional chlorine RuO2 and DSA, the results essentially mirror the discussion thus
evolution experiments, conducted in 5 M NaCl and 90 °C, far: for RuO2 both the rate of OER and rate of corrosion decrease
with a current density of 2 kA/m2. The effect of pH, in the range in a similar way as a chloride solution is acidified. For DSA, the
from 0 to 11, was studied. The electrode potential was essentially dissolution rate exhibits a minimum in the pH range 1−3, as seen
constant at about 1.33 to 1.38 V versus SHE in all experiments, by Gorodetskii et al.151 and Uzbekov and Klement’eva.229
indicating the constancy of the activity of the electrode during In 1987, Lyons and Burke234 presented a study of the
corrosion, over time, and at all pH values. It was found that there stabilization of RuO2 by addition of SnO2 in which the lifetime
are clear differences in the dissolution of Ru and Ti from RTO was evaluated by a standard accelerated lifetime test during OER
electrodes. During a long-term electrolysis experiment at pH 2, in concentrated NaOH at 80 °C.
during the first 30 h, removal of Ru from the coating was the In 1989, Zhinkin et al.163 presented results on the corrosion
dominating process. Although the ratio between Ru and Ti were behavior of standard DSAs, focusing on current densities
0.45 to 1 in the coating, during the first 30 h, the selectivity for exceeding 2 kA/m2 and several temperatures in the range from
dissolution of Ru over Ti decreased from values between 4 and 2 70 to 90 °C. The pH range was 2 to 5. Measurements were
toward equal dissolution. After 30 h, the selectivity was close to 1, conducted in 300 g/dm3 NaCl. The same radiochemical method
indicating equal rates of Ru and Ti dissolution. This means that as had been used in several previous Soviet studies was again
during electrolysis, the surface layers of RTO lose Ru and applied. The authors analyzed the gas phase and solid phase
become more concentrated in Ti. Measurements of the steady particulate losses of Ru, finding that the gas phase loss made up
state corrosion rate of Ti and Ru as a function of pH were also 20−40% of the total loss and that the solid particulate losses were
performed. At low pH (around and just under 0), the corrosion less than 1% of the total corrosion. A clear increase in the
rates of Ru and Ti were both elevated, but the selectivity for Ru corrosion rate was noted as the pH was increased above 4. This
dissolution was less than 1, indicating a higher rate of Ti increase was concomitant with a clear increase in the percentage
dissolution. As the pH then was increased, the dissolution rate of of O2 in the off-gas as the pH was increased. The connection
Ti decreased, reaching stable values above 2 and up to a pH of between Ru corrosion and the rate of the OER was thus again
5.4. One measurement was apparently then made at pH 11.3, at seen. Increasing the current density from 5 to 10 kA/m2 also
which the rate of dissolution of Ti had increased significantly. increased the corrosion rate but less significantly. There was a
The dissolution rate of Ru showed a minimum at about pH 2 clear effect of temperature on the corrosion rate, with higher
then increased. The resulting trend in the selectivity for corrosion rates noted at 90 °C than at 70 °C. Still, the effect of
dissolution of RuIV versus TiIV is an increase from less than 1 increasing the pH from 2 to 5 was more significant than both the
at the lowest pH, increasing to almost 9 at pH 5.4. Due to the effect of current density and of temperature.
clear increase in Ti removal rate at pH 11, the selectivity In 1990, Manli and Yanxi235 presented results on corrosion of
decreased to 4. Aside from the lowest pH values below pH 2, the DSA, prepared using the conventional method, during
dissolution rate of Ru was up to 9 times higher than the electrolysis at 10 kA/m2 in saturated NaCl solution at 80 °C.
dissolution rate of Ti. It was proposed that both Ti and Ru The corrosion rate was evaluated based on when the electrode
dissolve due to formation of soluble metal chlorides at the lowest passivated, resulting in a high anode potential. Secondary ion
pH values, while Ru removal could also occur due to the mass spectrometry (SIMS) results indicated that Ru is being
electrochemical conversion of RuO2 to RuO4 at higher pH selectively removed from the surface layer of the electrode during
values. At high pH values, Ti would be converted into soluble Ti electrolysis. This agrees with the results found by Uzbekov and
oxyhydroxides, in agreement with the Pourbaix diagram.230 Klement’eva.229
Further work from the same group appeared in 1985, this time In 1990, Loučka214 presented a Pourbaix diagram for Ru in
focusing on the corrosion of RTO during sodium chlorate concentrated chloride solution, as relevant for chlor-alkali
production.231 DSA were used, prepared using the conventional production. The main detail important to notice from this
thermochemical method and with the standard 30:70 Ru:Ti Pourbaix diagram is the supposed existence of soluble Ru
composition. Measurements were conducted using an anodic chlorides at low pH and anodic potentials. Interestingly, the
3012 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

diagram of Loučka only includes a very small pH range of RuO2 In 2003, Cornell et al.174 studied the critical anode potential
stability above the Cl2/Cl− equilibrium potential, between pH using commercial DSA for chlorate electrolysis. It is known from
2.3 and 3.0. It was pointed out that this pH range roughly agrees industrial practice that during the first months of industrial
with that found to give the highest stability of DSA electrodes.151 operation in a chlorate cell, the anode potential decreases
In accordance with the Pourbaix diagram of Loučka,214 at lower simultaneously with an increase in chlorate selectivity. From
pH, formation of soluble Ru chlorides was predicted to become laboratory studies, it was concluded that a new, relatively
favorable, while at higher pH, formation of (per)ruthenates compact DSA operates at a potential higher than Ecr. When in
becomes favorable. production, the real surface area of the electrode increases
In 1998, Hardee and Kus236 reported on studies of passivation (probably due to corrosion of the coating) and the electrode
of DSAs through formation of an interlayer between the coating potential thus decreases to a stable value at, or below, Ecr. The
and the substrate. Electrodes deactivated either in the industrial critical current density, at which the anode operates at Ecr,
setting or by using accelerated aging regimes exhibited changes in depends on factors such as the chloride concentration. Later
impedance phase angle at low frequencies (around 1 Hz) as well work by Nylén and Cornell238 concluded that process conditions
as in high-frequency regions (around 1000 Hz). These changes in industrial chlorate production leads to a higher risk of
were attributed to formation of a TiO2 interlayer, although no exceeding Ecr than the conditions in the chlor-alkali process,
direct detection of this interlayer was presented. A DSA coating where the chloride concentration is higher, pH lower, operating
with a number of TiO2 layers deposited on the surface showed temperature higher, and there is no chromate in the electrolyte.
changes only in the low-frequency region. Using XPS and atomic In 2004, Gajić-Krstajić et al.239 presented the application of a
emission spectroscopy (AES), removal of Ru in the outer layers spectrophotometric method to determine low concentrations of
of DSA after deactivation was noted. This was not deemed to be dissolved Ru species on the corrosion of RuO2 deposited on Ti.
the reason for the deactivation, as such removal was also noted The method was based on the catalytic action of Ru on the As−
for electrodes that still functioned with lower anode potentials. Ce redox reaction in sulfuric acid. This method is of a comparable
sensitivity (1 × 10−9 M Ru) as the radiochemical methods
Furthermore, a pure IrO2 electrode was deactivated in a sulfate
employed in several studies mentioned previously in this section.
electrolyte, after which changes particularly at high frequencies
The experiments were conducted in 0.5 M H2SO4 solution, at a
were noted. Also on the basis of this observation, it was
current density of 5 kA/m2. The service life, determined in the
concluded that formation of an interlayer was the mechanism of
standard way, increased exponentially with decreasing applied
deactivation, as a significant amount of IrO2 still remained on the current density (and thus with decreasing rate of OER). By
electrode. performing measurements at several potentials in the range from
In 2001, Tilak et al.226 presented an electrochemical 1.52 to 1.67 V versus SCE, a constant dependence of the current
impedance spectroscopy (EIS study of the deactivation of DSA
density for oxygen evolution, jO2, on the current density for Ru
electrodes. An accelerated aging process was employed, wherein
the potential was changed from 1.35 to −0.32 V versus SCE at 60 dissolution, jRu, was found
Hz. These results support those reported several times before, jRu
that during electrolysis, corrosion results in the selective removal = 8 × 10−6
jO (42)
of Ru from the surface layers of the electrode. Tafel, CV and EIS 2

results indicated that deactivated electrodes (as determined by


In 2014 and 2015, a number of papers of Mayrhofer and co-
the standard passivation test of an increase in anode potential)
workers240−243 were published. In all papers, an in situ method to
behave similarly to fresh electrodes prepared with a 5% Ru
measure transition metal degradation products was applied, in
content. Furthermore, the authors note that fresh electrodes with
which an electrochemical flow cell is coupled to an inductively
5% Ru behave the same way in the low frequency region as aged coupled plasma-mass spectrometer (ICP-MS).244 First, Zer-
30% Ru electrodes, indicating that the deactivation process of adjanin et al.242 reported results on the connection between
DSA electrodes is due to removal of Ru from the surface, rather RuO2−SnO2 mixed oxide surface morphology and OER activity
than formation of an insulating TiO2 layer between the Ti and stability. Electrodes with two types of surface morphology,
substrate and the coating. either cracked or smooth, were prepared using thermal oxidation
In the same year, Eberil’ et al.227 presented results on aging of of sol−gel-synthesized precursors onto Ti substrates. The usage
DSA electrodes under standard chlorate production conditions. of more dilute solutions of the Ru and Sn precursors in
It was seen that the anode lifetime increased with the amount of isopropanol, together with repeated coating application cycles,
Ru in the electrode, indicating the relationship between anode enable the formation of smooth surfaces. Electrolysis was
lifetime and the dissolution of Ru from the coating. Furthermore, performed in a 0.1 M H2SO4 electrolyte at 25 °C. The sensitivity
as Tilak et al.226 did in the same year, Eberil’ et al. also performed of the ICP-MS instrument, allowing detection of nanograms of
tests to examine the proposed deactivation mechanism of Ru per square centimeter of electrode area, is such that
formation of a TiO2 layer between the coating and support. It was dissolution of Ru from the RuO2 component of the mixed
seen that a slight improvement in anode lifetime resulted from coating during cycling of the potential in the range from 1 to 1.6
the insertion of a Pt sublayer between the support and the V versus RHE could be measured. For an industrial catalyst, the
coating. However, introducing Pt into the coating itself resulted current density is normally kept constant, apart from at
in a significantly improved lifetime. This also indicates that the shutdown, so these results are not directly comparable to a
reason for DSA deactivation is the removal of Ru from the steady-state dissolution rate during industrial electrolysis.
coating, rather than the formation of a nonconductive interlayer. Furthermore, it is unclear what the ratio between RuIV and
Also in 2001, Bommaraju et al.237 presented results on SnIV components in the mixed coatings were. Still, it was found
deactivation of DSA electrodes. Much of the work was presented that the dissolution rate of Ru increased as the potential was
in Tilak et al.,226 but the authors also showed SIMS results increased over the reversible potential for OER of about 1.25 V
agreeing with those of Manli and Yanxi.235 versus RHE, and became especially high above 1.45 V versus
3013 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Table 4. Effects of Process Conditions on the Selectivity between Oxygen and Chlorine or Chlorate Formation Identified in
Literaturea
selectivity for oxygen
factor evolution refs comment
low pH (chlorine production) decreased 41,151,154, and 163 effect of pH is complex
pH for chlorate conditions optimum pH value exists 56, 67, 141, 162, and 173
higher cCl− decreased 66, 70, 75, 76, 121, 128, 154, 155, 162, 170, and 173 especially if cCl− < 100 g/dm3
increased j decreased 54, 56, 57, 62, 66, 69, 70, 75, 131, 154, and 162−164 increase at low cCl−170
higher cHOCl increased 54, 55, 57, 71, 116, 121, 130, 162, and 170
higher cNaClO3 decreased 6, 54, 162, and 173 related to viscosity?
increased cell temperature increased 54, 57, 74, 154, 162, and 164
anode potential (chlorate production) optimal anode potential exists 74 and 164
higher anode (changed geometry) decreased 57
larger holding volume decreased 55 and 62 current efficiency not directly measured
irradiance increase 6 and 94
a
The table focuses on trends described in concentrated NaCl solutions. This table gives a simplified view of the factors involved. The trends are
described in more detail in section 3.1.1.

Table 5. Effects of the Electrolyte Composition on the Selectivity between Oxygen and Chlorine or Chlorate Formation Identified
in Literaturea
selectivity for oxygen
factor evolution refs comment/contradictory results
F− possible decrease 137, 144, and 173 137
found no effect
PO3−
4 increase 137, 157, and 173 not with Pd71
137
SO2−
4 increase 138 and 157 found no effect
NO−3 increase 137
Na2Cr2O7 increase? 137 no effect,92,95,162 slight decrease173
Al no effect peroxide,96 oxide92
Ag increase oxide sources cited in ref 93 salt92 has no effect
As increase metal94
Ce no effect salt92
Co increase salt/ionic,86,92,94,97,100 metal,94 oxide93,95−97,105,106 and sources
in,93 unspecified168
Cr no effect salt,92 oxide95
Cu increase salt/ionic,90,97,103 metal,94 oxide,94,96,97,105 unspecified168 less active than Co, Ni
Fe increase? metal,94 oxide sources in93,96 salt,86,92 metal,93 oxide92,95 or unspecified form168
has no effect
Hg increase? oxide96
Ir increase salt/ionic,92,101,102 oxide,95 unspecified168
Mn increase? oxide sources in93,96 salt,86 metal93 has no effect
Mo no effect oxide92
Ni increase salt/ionic,103 metal,94 oxide95,96,105 and sources in93 unspecified168
Pt no effect unspecified168
Sn increase? metal94
RTO, Pt/Ir no effect particulates159 Ir increase according to ref 62
Ru no effect salt/ionic,92,95 oxide,92 unspecified168
Rh no effect salt/ionic95
U, W, Pd, Os, no effect oxide95
Tl, V
Sb, Pb, Zn, Cd no effect metal94
a
Several of the sources are studies of decomposition of bleach (sources cited in refs 93−95) or hypochlorous acid solutions (refs 90, 92, 93, 96, 97,
100, 102, 103, 105, and 106, which differ significantly from the electrolytes used in industrial chlor-alkali or sodium chlorate production. This table
gives a simplified view of the factors involved. Sources cited in ref 93 are published before year 1900. Each factor is discussed in more detail in section
3.1.2.

RHE, where the conversion of RuO2 to RuO4 can start. thereafter, Cherevko et al.241 reported on the dissolution of IrO2
Dissolution was also noted at potentials below 1.23 V versus films, prepared using spin-coating onto Ti substrates and
RHE, during cycling between 1.23 to 0 V versus RHE. It was conventional thermal oxidation, during OER in 1 M HClO4. It
found that the crack-free coating experienced a higher rate of was found that the stability of the films increased with increasing
decomposition in all potential ranges studied. Cherevko et al.240 calcination temperature, in the range from 250 to 550 °C, except
then reported results on dissolution of metallic electrodes of Ru, for electrodes prepared at 350 °C which was found to be the least
Ir, Rh, Pd, Pt, and Au, finding a clear correlation between a low stable. During constant-current experiments, performed at 20 A/
Tafel slope for OER and a high dissolution rate. Shortly m2, it was found that dissolution of both Ir and Ti occurred, with
3014 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Table 6. Effects of the Anode Composition on the Selectivity between Oxygen and Chlorine or Chlorate Formation Identified in
Literaturea
selectivity for oxygen
factor evolution refs comment
modified composition decrease 36, 158, 162, 167, 169, and several modified coatings exist
173
increased surface area increase 69 related to decrease in jloc?
larger coating crystallite decrease 78 related to increase in jloc?
particle size
higher oxide loading increase 75
less Ru in RTO coating decrease 67, 68, 75, 139, 154, 161, same effect as decreased j0 for reaction 11? Connected to increased chlorate
and 171 formation36,141,150,154
decreased j0 for reaction 11 decrease 69, 150, 159, and 167 but reaction 11 not only source
IrO2 substitution decrease? 74 doped into RTO; dorrosion products might increase selectivity in practice
Co substitution decrease? 78 doped into RuO2;82,83,189 saw increase
SnO2 substitution decrease 69, 71, 147, and 172 doped into RuO269,172 or together with Pd71
HfO2 substitution increase 69 doped into RuO2
Zn substitution increases 181, 182, and 189 doped into RuO2
Ni substitution decrease 77 doped into RuO2
Mg substitution decrease 193 doped into RuO2
Fe substitution decrease? 180 doped into RuO2, high Cl2 selectivity in Ar-saturated solutions
overlayer material decrease 60 overlayer of Pt mixed with IrO2
support material possible effect 8 and 179
a
The table is limited to oxides based on RuO2. “Modified composition” signifies studies where details about the composition or physical properties
of the electrodes were not given. Each effect is described in more detail in section 3.1.3. The table only includes studies where selectivity was
measured directly [i.e., not only by comparing differences in electrochemical behavior (e.g. Tafel slopes) in solutions with and without chloride ions].

the dissolution rates of both components decreasing with time. 3. DISCUSSION


In 2015, Reier et al.243 reported a combined OER activity and 3.1. Factors Affecting the Selectivity between Chlorine and
stability study of Ir−Ni oxides (rutile-like at Ni contents below Oxygen Evolution in the Production of Chlorine and
21% and of a layered brucite-like structure at higher Ni contents) Chlorate
in 0.1 M HClO4. It was found that coatings containing 67% Ni The research literature portrays some general trends in the
(as measured by inductively coupled plasma atomic emission selectivity between chlorine and oxygen evolution. Some factors
spectroscopy) were most active for OER, which was related to a have been examined in several articles and can be considered
well-understood. Others have been reported only once, or, in
higher active surface area of this composition. However, the Ir
some cases, with different authors finding different results.
dissolution rate of the coating with this composition was 34 times Sections 3.1.1, 3.1.2, and 3.1.3 will discuss the selectivity from the
higher than the dissolution rate of pure IrO2, even at a low point of view of process, electrolyte, and anode factors, while
current density of 10 A/m2. section 3.2 will discuss the relative importance of the different
Also in 2014, Frydendal et al.245 presented a study where reactions that can lead to formation of oxygen. Tables 4, 5, and 6
electrochemical quartz crystal microbalance (EQCM) and ICP- summarize the different factors, related to process conditions,
electrolyte contents, and electrode composition, respectively,
MS measurements were combined to monitor corrosion of rutile that have been found in literature, as well as the literature
RuO2 and amorphous MnOx. The electrodes had been prepared references for each factor. The purpose of the discussion is to
using sputter deposition on Au or EQCM Au quartz crystals. highlight the trends that have been reported, as well as to point to
Constant potential, at 1.8 V versus RHE for RuO2 and 1.8 and 1.9 factors that have been reported to have opposing effects in
V versus RHE for MnOx, and constant current, at 300 A/m2 for different studies. The discussion focuses specifically on the trends
that are relevant for chlor-alkali and sodium chlorate formation
RuO2 and 200 A/m2 for MnOx, measurements were carried out
and is thus limited to concentrated NaCl solutions. In some
in either 0.05 M H2SO4 in the case of RuO2 or 1 M KOH in the cases, the trends in dilute solutions are different, as shown by, for
case of MnOx. Constant mass loss rates were noted in both types example, the study of Czarnetzki and Janssen.170 Finally, the
of measurement, which were carried out for 2 h. These electrodes connection between anode stability and anode selectivity will be
thus did not exhibit any stabilization during electrolysis, in discussed in section 3.3.
3.1.1. Influence of Process Conditions. There are two
contrast with what has been found for both pure and mixed rutile
factors that, at first, seem relatively straightforward to consider:
oxides in other studies. A year later, Frydendal et al.246 found a the thermodynamics of the two main competing reactions 2 and
that the stability of MnO2 coatings for OER in 0.05 M H2SO4 11 (written again below) and the effect of the current density on
could be improved by codeposition of other transition metals. A the selectivity.
combination of theoretical and experimental results indicated The equilibrium between chloride ions and chlorine is
described by
that mixed MnO2−TiO2 coatings, prepared by cosputtering,
exhibited an improved stability at the cost of a slight decrease in 1
Cl− ⇌ Cl 2 + e−
activity. 2 (2)

3015 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

For the OER, the equilibrium can be written as ηp = A p + bp × log(j) (44)


1
H 2O ⇌ O2 + 2H+ + 2e− for the parasitic reaction. If the Tafel slope b for the main reaction
2 (11)
is lower than the Tafel slope of the parasitic one, it is clearly seen
It is obvious that the latter equilibrium is dependent on the pH of that the overpotential for the parasitic reaction increases more
the solution. The reversible potential of the first reaction is also quickly with the current density. This is the case for the
influenced by the chloride concentration, as will be discussed. competition between the chlorine evolution and water oxidation
Oxygen evolution is increasingly unfavorable as the pH reactions. This simplified treatment assumes that the reactions
decreases, keeping other conditions (e.g., current density) take place on separate active sites, an assumption that might be
unchanged, due to the pH dependence of the electrode potential questionable if the two reactions compete for the same active
for oxygen evolution, unlike that for chlorine evolution.41,151,165 sites. It also neglects the effects of both mass transport of
Only in chlor-alkali processes with HCl addition can oxygen reactants to the anode surface, and of the formation of
evolving reactions be ruled out almost completely.41,151 This bubbles,247,248 on the local concentrations of reactants at the
would imply that oxygen evolution is a more severe problem in anode.247,248 Both of these effects will become important at the
chlorate processes, in which the pH of the process solution bulk high current densities employed industrially. As several different
is higher (pH 5.5−7). However, Trasatti and Lodi165 point out reactions yield oxygen, especially under chlorate production
that if water oxidation occurs at the anode, the water oxidation conditions, and since the effect of competitive adsorption is
reaction 11 acidifies the electrolyte close to the anode. Even if no important, this Tafel treatment of the influence of current density
water oxidation occurs, the formation of hypochlorous acid, gives a limited understanding of the effect of current density. A
reaction 4, acidifies the electrolyte. It is therefore considered more quantitative variant of this type of analysis is that by Tilak et
reasonable that, even in processes where the bulk pH is high, the al.,69 as given by eqs 26−28.
pH at the anode surface is in the acidic range. However, this Nevertheless, studies performed so far have found that
reasoning is not sufficient to actually rule out any of the oxygen- increasing current densities decrease the selectivity for oxygen
forming reactions, since the pH is probably not sufficiently acidic evolution (refs 54, 56, 57, 62, 66, 69, 70, 75, 162, 164, and 165).
at the anode surface to fully prevent water oxidation. Indeed, This could indirectly imply that the most important anodic side-
several studies54,56,67,162,165 have shown clear correlations reactions that yield oxygen have higher Tafel slopes than the
between pH and the oxygen evolution rate. The disagreement chlorine evolution reaction. However, this effect is likely tangled
between different studies of the influence of pH also highlights an with the effect of a decreasing pH at the anode surface with
important difficulty in studying pH effects in chlorate electro- increasing current density. The effect of current density is
lytes. Hammar and Wranglén54 did not find any influence of pH probably that exhibited also by electrodes with decreased surface
on the oxygen evolution rate between pH = 6.5 and pH = 11. area,69,78,79 which yields an increase in local current density.
However, Jaksić56 later found that the optimal pH for chlorate Moreover, it is important to also recall that Eberil et al.74 found a
production should be in the pH = 6−6.5 range. Jaksić believed maximum in chlorate formation efficiency close to the critical
this disagreement to be the result of improper handling of the anode potential (for RuO4 formation), meaning that increased
buffering effect of carbonates from absorption of CO2 from the current density increases the selectivity only up to a certain point
air. Moreover, Hardee and Mitchell162 only found an influence of at chlorate conditions. Although several studies have seen effects
pH at pH > 7, and no effect of pH on the rate of oxygen evolution of current density on the selectivity, further study of this factor
in the range of pH = 5−7. However, Hammar and Wranglén54 under chlorate conditions is warranted.
used graphite anodes. Nevertheless, chlorate electrolytes are The effect of increasing concentration of hypochlorite species
buffered both by dichromate species and hypochlorite species, is related to the decomposition of hypochlorite species to form
and several reactions in the process involve H+, making the study oxygen (refs 54, 55, 57, 71, 116, 121, 130, 162, and 170). This
of pH effects in chlorate electrolytes somewhat difficult. reaction will be discussed in more detail in section 3.2.
Eq 2 also shows the influence of chloride concentration. An effect of chlorate concentration, due to competitive
Several studies and reports66,70,75,76,121,128,154,155,162,170,173 note adsorption of chlorate on the electrode, on the oxygen evolution
that increasing concentrations of Cl− increase current efficiencies side reaction has been shown.54,162 However, Hammar and
for chlorine or chlorate production and decrease the selectivity Wranglén54 made estimations that showed that the increase in
for oxygen evolution. However, this might not only be related to electrolyte viscosity that chlorate brings could be the main reason
the decrease in the reversible electrode potential for chlorine for the decrease in oxygen evolution rate. This increase in
evolution but also due to increased rates of mass transfer of viscosity would result in a decrease in the mass transfer rate of
chloride to the anode surface as well as competitive adsorption of hypochlorite to the anode surface and a decrease in oxygen
chlorides on the surface of the anode, preventing anodic oxygen- produced through decomposition of hypochlorite.
evolving reactions.66,165 The effect of temperature is probably complicated to elucidate
The next fundamental factor that needs to be considered is in a detailed way, since it has effects on both thermodynamics and
that of the current density on the selectivity. It is simple to kinetics. This is true especially at chlorate cell conditions, since
consider the effect of current density on two separate reactions, several reactions are involved in the main reactions yielding
such as reactions 2 and 11. If Tafel expressions are written for chlorate and the side reactions yielding oxygen. Several studies
both reactions, with the respective subscripts m and p for the have found that increasing temperatures yield increased rates of
main (ClER) and the parasitic (OER) reaction, the following oxygen evolution at chlorate conditions.54,57,162,164 This has been
expressions are obtained suggested to be related to increasing rate of diffusion of
hypochlorite species to the anode surface at increasing
ηm = A m + bm × log(j) (43) temperatures54 and to increased rates of anodic and homoge-
neous decomposition of hypochlorite species at higher temper-
for the main reaction and atures.162 Eberil et al.74 found a complicated temperature
3016 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

dependence under chlorate conditions, at which the temperature evolution under chlorate process conditions, but they found that
had different effects at anode potentials on different sides of Ecr. addition of nitrates resulted in increased oxygen evolution rates.
Furthermore, the critical anode potential was associated with a The effects of all of these species seem to be related to
maximum in current efficiency for chlorate production and a competitive adsorption on the anode surface, and these effects
minimum in oxygen evolution current efficiency. This shows that might become less important at high chloride concentrations.
it is imperative that further study of the selectivity issue is Dichromate has been reported to both increase137 and
performed with equipment that allows precise correction of iR decrease173 the selectivity for oxygen evolution. Interestingly
drop, since the critical anode potential is reached only at high enough, other studies have seen only minor effects.162 All three
applied current densities. Using such equipment makes it studies were performed at chlorate process conditions. Studies of
possible to determine the effect of Ecr clearly. dilute hypochlorous acid solutions have not found any effect of
The effects of anode geometry that Wanngård57 found is likely dichromate on the selectivity for oxygen evolution.92,95
related both to a decrease in local current density, due to the A small number of metals, salts, and oxides of the elements
increased geometric surface area of the higher electrode blades, have been shown to exhibit catalytic effects on hypochlorite
and a change in the concentration gradient in the electrolyte flow decomposition. These are Co, Ni, Ir, and Cu. The effects of Fe,
across the blade for the higher type of blade (the “lower” blades Mn, Hg, and Sn are less clear, with some (especially very old
were 0.333 m high and 0.379 m wide, while the higher blades studies) reporting acceleration of hypochlorite decomposition.
were 0.8 m high and 0.3 m wide). Connick and Hurley99 suggested that ruthenate ions might
Increasing the external holding volume has been found to catalytically decompose hypochlorite ions (eqs 16 and 17).
increase the charge yield for chlorate formation.55,62 This is likely However, considering that the window of stability for the
related to the increased residence time for the electrolyte ruthenate ion is at high pH values of about 13−14 and potentials
afforded by a larger holding volume, enabling a higher conversion less than 1 V versus SHE,214 the prevalence of this reaction as an
in the homogeneous chlorate formation reaction. explanation for the oxygen evolution side reaction in either the
3.1.2. Influence of Additions and Contaminants in the chlor-alkali or the chlorate process is deemed quite low. No
Electrolyte. There have been comparatively few studies on the studies have been performed to compare the effects of different
effect of contaminants and additions to the electrolyte on the elements under industrial chlorate conditions. Some materials
oxygen evolution side reaction. In some cases, the species that have not been studied since the 1800s, and a more general study
have been studied have been found to have opposing effects. The of the heterogeneous activities of different metals, and possibly
great majority of studies are performed on systems that differ combinations of different metals, could serve to settle these
greatly from industrial chlorate or chlor-alkali conditions (e.g., on issues. An important aspect to consider in such studies is the
bleach or hypochlorous acid solutions), again indicating the speciation of the metallic compounds, as it is possible that salts
importance of performing further studies under conditions that are added will be oxidized to form metallic oxides active for
similar to those used industrially. decomposition of hypochlorite.92 It is clear that the influence of
Fluoride was found by both Fukuda et al.144 and Baolian and many different possible electrolyte contaminants and additives
Wenhua173 (based on the overall chlorate cell current efficiency) has not yet been examined fully under industrial conditions.
to decrease the rate of oxygen evolution. On the other hand, 3.1.3. Influence of the Anode Structure and Compo-
Jaksić et al.137 found essentially no influence of fluoride on the sition. The effect of the anode properties on the oxygen
oxygen evolution rate under chlorate conditions. The difference evolution side reaction has been studied by many. Unfortunately,
between the study of Fukuda et al. and Jaksić et al. might be many studies lack important details regarding the properties of
possible to rationalize based on the differences between the the anode. For example, proprietary coatings of unnamed
electrolytes studied (Jaksić et al. studied a chlorate electrolyte composition that lead to lower oxygen evolution rates have been
while Fukuda et al. studied an acidic electrolyte), if the results of used in several experimental studies.36,162,167,173 Most studies
Baolian and Wenhua were ignored. Still, the results might be have focused only on the selectivity between oxygen and chlorine
reconciled by noting that the effect seen by Baolian and Wenhua formation under chlor-alkali conditions. An exception is the
was significantly smaller (about 1%) at 200 g/dm3 NaCl than at study of a modified coating with decreased selectivity for oxygen
50 g/dm3 NaCl. Furthermore, the measurements of Jaksić et al. evolution at chlorate conditions performed by Spasojević et
were performed with a NaCl concentration of 310 g/dm3. al.71−73 Additionally, detailed characterization, such as of the
However, Wanngård168 notes that F− needs to be avoided since it actual composition of the final coating has usually not been
also has the effect of deactivating the anode. Kelsall and reported in literature, even if studies on proprietary anodes are
Robbins145 have predicted, based on thermodynamic calcu- excluded. This is a problem, especially since the preparation of
lations, that F− addition should result in dissolution of TiO2 to anodes used for research usually is a manual process where the

form TiF2− 6 . Still, it might be concluded that the effect of F coating is, for example, brushed onto a substrate. Furthermore, it
addition on selectivity essentially is that of competitive is known that the deposition efficiency of different compounds
adsorption on the electrode surface, reducing the rate of oxygen can vary widely depending on which precursors and which
formation in dilute electrolytes, but also possibly competing with preparation conditions are used.249 This lack of detailed materials
the adsorption of Cl−. characterization could explain some of the conflicting results in
Phosphates have been found to increase the rate of oxygen literature. This does not necessarily imply that the studies are
evolution.137,157,173 In contrast, the PdSn2-containing coating poorly done, since many of the studies were performed when
devised by Spasojević et al. made use of the adsorption of characterization methods that are common today, such as SEM,
phosphates on Pd to prevent oxygen evolution. Sulfates have XRD, XPS, and atomic force microscopy (AFM), had not
been reported to decrease the efficiency of chlorine evolu- entered general use. Furthermore, the usage of other X-ray
tion.157,165 Kazarinov and Andreev143 concluded that both spectroscopies (e.g., XAS), which allow a deeper understanding
phosphates and sulfates adsorb strongly on RTO, displacing of factors such as the local structure, were not as accessible as
chlorides. However, Jaksić et al.137 saw no effect on oxygen today. Nevertheless, even studies conducted today often do not
3017 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

make use of the possibilities afforded by modern characterization DFT studies by Rossmeisl et al.,251 Hansen et al.,183 and of Exner
methods. et al.207,210 In these studies, the importance of adjacent CUSs on
All the same, some overall factors can be identified. One effect, the anode surface were considered. First, it has been found that it
that of the active surface area of the electrode, is likely primarily is not only the metal CUSs themselves (found by Hansen et al.183
connected to its effect on the local current density, as discussed in to be the active sites on weaker-binding oxides) but also O-
section 3.1.1. While a large active surface area is obviously covered CUSs183,207 that are active sites for both ClER and OER.
important to achieve a high utilization of the coating, an Second, it has been found that the mechanism for OER does not
increased electrode surface area is also associated with a necessarily involve O−O recombination,183,207 thus not being
decreased local current density and resulting increased selectivity dependent on paired CUSs. The finding that the adsorption
for oxygen evolution. An increased electrode active surface area is energy of O on CUSs on the surface is the deciding factor for
often achieved by decreasing the size of the particles making up both chlorine and oxygen evolution activity might at first be hard
the coating. The selectivity effects from doping RuO2 or RTO to reconcile with experimental studies that show that materials
that are found in many studies are most likely due to this effect. with increased overpotential for oxygen evolution also show an
Arikawa et al.75 found that the selectivity is affected by the RTO increase in selectivity for chlorine evolution. This can be
loading on the anode, with increasing loading resulting in understood by studying Figure 23, which shows that there are
increased selectivity for OER. This effect is possibly also ΔE(Oc) values which allow decreased selectivity for oxygen
connected to an increase in active surface area, especially since evolution, compared with RuO2, with maintained low over-
Arikawa et al.75 studied electrodes with comparatively low potential for chlorine evolution.64 Indeed, Karlsson et al.64 have
loading. built upon the results of Hansen et al.183 to conclude that the
The atomic-level electronic effect of using different combina- increased selectivity exhibited by RTO electrodes likely is caused
tions of cations in mixed oxide electrodes (e.g., the combination by an electronic activation of Ti by Ru. The effect on selectivity
of Ru and Ti in Beer-type DSA) is more complex. Certain trends from changing the relative amounts of Ru and Ti in the binary
that are valid for Ru-based oxides might not apply for, for coating is thus not a matter of separation of active Ru sites67,68
example, Ir-based oxides, as has been shown by, for example, but related to the fundamental electronic properties of the mixed
Kuznetsova et al.189 (pure IrO2 being more Cl2-selective than any oxide coating and its effects on the activity of Ti sites.
of the doped oxides considered, while Ti or Sn-doped RuO2 Much work has been devoted to improving the selectivity by
oxides are more Cl2-selective than pure RuO2). Even if one using other materials than RTO, or by doping RTO. Studies
focuses only on “Ru-based” electrodes, the electronic structure which examine the effect of IrO2 have found it to have a lower
and therefore also the electrochemical properties of materials selectivity for oxygen evolution than pure RuO2,70 and Ir-doped
containing low concentrations of RuIV are significantly different RTO have also shown decreased O2 selectivities.74,165 At the
from materials containing high RuIV concentrations.64 However, same time, the dissolution of Ir compounds from the anode
it is important to consider the cautionary observation of might yield an increase in oxygen evolution in the industrial
Trasatti156 that the selectivity of different anode materials does setting due to catalysis of homogeneous hypochlorite decom-
not generally exhibit varying selectivities for chlorine evolution position. However, RuO2−Sb2O5−SnO2 electrodes have been
versus oxygen evolution. Materials highly active for chlorine found to exhibit similar selectivities in dilute solutions regardless
evolution are usually also highly active for oxygen evolution. A of Ir doping.185,186 PdO has been reported to be a stable and
first fundamental understanding for this behavior has been highly selective ClER anode.148 Shlyapnikov141 found that Pt and
provided by the DFT studies of Hansen et al.,183 showing that RuO2 exhibited higher current efficiencies for chlorate formation
ΔE(Oc) determines the activity for both chlorine and oxygen than graphite, lead oxide, or manganese oxide anodes. On the
evolution reactions on several metallic oxide materials. The other hand, platinized Ti has been reported to have a lower
fundamental factor that governs the selectivity is the descriptor chlorine evolution efficiency than RuO2−SnO2-coated Ti.187
value placing each material in a certain part of the activity volcano Co3O4 and related compounds have also been reported to have a
relevant for a certain set of linearly scaling reactions.183,199,250 lower selectivity for oxygen evolution.78,140,147,165 However, also
Nevertheless, while the same factor governs both reactions, there in this case, Co corroded from the surface of the electrode might
is some room for optimizing the selectivity for chlorine evolution yield increased oxygen evolution due to homogeneous
over oxygen evolution even on conventional rutile oxide hypochlorite decomposition. Combining Ru with Sn in oxide
electrodes, as has been explored theoretically63−65 and as electrodes has been found to have generally similar effects as the
experimental studies also indicate (see Table 6). combination of Ru and Ti in RTO electrodes.172 However,
The effect of altering the relative amounts of Ru and Ti in RTO studies have found that Ru−Sn oxide electrodes are less selective
coatings will be discussed first. A decrease in the Ru content has for OER than RTO electrodes.69,188 Tilak et al.69 noted similar
consistently been shown to result in decreased selectivity for effects when combining Ru and Hf in mixed oxide electrodes.
oxygen evolution (refs 67, 68, 75, 159, 161, 165, 167, and 171). Doping RuO2 with Zn, Ni, or Mg have also been found to yield
Similar trends regarding Ru content and OER selectivity have lower selectivities for oxygen evolution, in comparison with those
been noted for Ru−Sn oxide electrodes.172 This might be of pure RuO2.77,181,182,193
another aspect on the observation that an increased overpotential The insights from DFT63−65,183,184,207,210 are also interesting
for oxygen evolution of an anode material (i.e., a decreased j0 for when discussing the recent study of Spasojević et al.60 They
reaction 11) has been linked to a decreased selectivity for oxygen found that Pt-IrO2 overlayer coatings on RTO show decreased
evolution in chloride-containing solutions.69,150,159,167 These selectivity for OER as a side reaction in both chlor-alkali and
factors have been related to mechanistic differences between the chlorate electrolytes.60 The overlayer electrodes exhibited a
oxygen-evolving reaction and the chlorine-evolving reaction. It significantly lower electrochemically active surface area (in-
has been proposed that both reactions need Ru sites to progress, creased local current densities), which is likely part of the reason
but that the oxygen-evolving reaction needs two Ru adjacent for the improvement in selectivity (see section 3.1.1). However,
sites.67,68 Here an interesting new insight can be gained from the the work might also be reflective of the same fundamental effects
3018 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

discussed in several theoretical works.63−65,183,210 Doping TiO2 Finally, Tilak and Chen5 believe it has an insignificant influence
with Ru changes the descriptor value of the doped oxide toward under chlorate production conditions.
the part of the volcano optimal for chlorine evolution,64 and it is The influence of the water oxidation reaction, reaction 11, is
possible that Spasojević et al.60 is noting a similar effect when also uncertain. It is known that this reaction dominates at low
mixing IrO2 with Pt. PtO2 and IrO2 are located on opposite sides chloride concentrations and low electrode potentials,54 but at
of the volcano tip optimal for selective and active chlorine actual industrial current densities and concentrations, the
evolution.65,183,210 influence is unclear. Kotowski and Busse159 think it makes up a
Some have found effects on the selectivity for chlorine significant amount of the oxygen evolved in the membrane chlor-
evolution based of the support material.179 This influence is also alkali process, while Hardee and Mitchell162 and Spasojević et
hinted at in the original Beer patents.8,22 al.71,72 find it to be insignificant at chlorate conditions. Tilak and
Less attention has been devoted to the possibility that the Chen5 believe it is the only important source of oxygen at
electrode might heterogeneously catalyze the decomposition of chlorate conditions. The studies of Spasojević et al.71 and
hypochlorous acid species. While certain compounds (such as Spasojević et al.60 indicate that a majority of the oxygen is formed
Co) are generally not considered as electrode components, due through water oxidation (see Figures 8 and 18) in dilute
to their well-known catalytic effect as electrolyte contaminants solutions with similar hypochlorite concentrations as in chlorate
(see the previous section), there are fewer studies of production. It seems that the rate of water oxidation is
heterogeneous hypochlorite decomposition. When it comes to significantly lower in concentrated chlorate solutions with
metallic and oxidic materials (see Table 5), Co, Ni, Ir, and Cu are chromate buffering,57,71,72,162 making it a less important source
materials which have been found to catalyze the bulk of oxygen than the decomposition of hypochlorite in sodium
decomposition. Still, Kotowski and Busse159 did not find that chlorate production. The water oxidation reaction is likely
Pt/Ir particles were active for hypochlorite decomposition, kinetically controlled, while the oxidation of hypochlorite species
despite the Ir content. Kuhn and Mortimer136 did see indications is diffusion controlled,60,71,72 at least at the current densities
applied in industrial chlor-alkali and chlorate production.
that RuO2 heterogeneously decomposes hypochlorite in a Cl2-
The homogeneous decomposition of hypochlorous acid to
saturated solution, while Kokoulina and Bunakova155 and
form oxygen gas, reaction 14, is well-known. The catalytic effect
Kotowski and Busse159 found no effect when RTO electrodes
of different electrolyte contaminants on this reaction has been
were used. To get complete certainty in this matter, studies
discussed in section 3.1.2. Hypochlorite does decompose to form
where both RuO2, RTO and doped RTO or RuO2 electrodes are oxygen even in the absence of any catalyst, although the rate is
tested under the same conditions, with the same equipment and significantly lower than that of chlorate production.89,91,92 There
methods, are likely required. are indications that both reactions share an intermediate.89,92
Finally, some studies36,74,141,150,154 indicate that for the chlor- While measures to control the electrolyte pH can be applied to
alkali process, it might not be enough to simply optimize anodes prevent the formation of hypochlorous acid species in the chlor-
for minimal oxygen selectivity, as this is then associated with alkali process,36 this is obviously not a possible recourse in the
increased rates of chlorate production. This might partly be a pH chlorate process. Further research is needed to elucidate the
effect, as the OER decreases pH at the anode and therefore also connection between oxygen and chlorate formation, so that
decreases the rate of chlorate formation.154 A reduced rate of the measures might be found to influence the selectivity in the
OER with a low-oxygen anode might thus require additional direction of chlorate formation. More studies of this reaction are
anolyte acidification,36 although this might yield an increase in needed to fully understand its importance in chlorate production.
cell voltage due to decreased conductivity of the cation- Still, it can be said that purification of the electrolyte to remove
permeable membrane.2 Whether the effect could be associated metallic contaminants should minimize the importance of
with changes in activity or selectivity for electrochemical chlorate homogeneous hypochlorite decomposition in both chlor-alkali
formation on the surface of the anode is yet to be determined. and sodium chlorate production.
3.2. Extent of Oxygen Evolution from Different Reactions Finally, the influence of anodic decomposition of hypochlorite
species, resulting in formation of oxygen but not chlorate, is
We now change focus to discuss the relative importance of the controversial. Kotowski and Busse159 have suggested one
different reactions thought to result in oxygen production. Only a mechanism for this reaction, eq 23. Older studies consider the
few studies have attempted to quantify the five most probable anodic decomposition of hypochlorite a possible source of
oxygen-evolving reactions, reactions 10, 11, 12, 14, and 23, that oxygen,116,121,130 but most studies performed before 1980 seem
are present in varying degrees at chlorate and chlor-alkali to conclude that it is a minor source of oxygen. Several studies
conditions. Due to the low pH at the anode surface, anodic performed since the end of the 1980s have shown it to be an
hydroxide oxidation, reaction 12, is generally assumed to be an important factor for oxygen evolution in chlorate sys-
unimportant source of oxygen. The importance of the other tems.54,57,159,162,177 The fact that the anodic decomposition of
reactions is less clear. hypochlorite was not thought to be an important side reaction at
The influence of the first reaction, the anodic chlorate chlorate process conditions until the second half of the 1980s
formation reaction, is controversial. Studying chlorate process shows that indirect methods used to measure current efficiency
conditions, Hammar and Wranglén,54 Ibl and Vogt,62 and Jaksić for chlorate production are insufficient to give a full under-
et al.55 thought it was the most important oxygen-evolving standing of the process. It took studies where the proper
reaction. Recent modeling studies make the same assumption.191 detection of the oxygen evolved was performed to find that other
These models describe the overall current efficiency of the reactions might also be important sources of oxygen. Never-
process well. However, Hardee and Mitchell162 seem to think it theless, some still do not agree that anodic decomposition of
has a small influence. Studying a membrane cell chlor-alkali hypochlorite species is important,5,167 warranting further studies
system, Kotowski and Busse159 thought it could make up of how hypochlorite is actually decomposed at the anode. The
anywhere from 20% to 60% of the oxygen evolved in the system. relative importance of heterogeneous chemical decomposition
3019 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

and electrocatalytic decomposition of hypochlorite species on, at both acidic, neutral, and alkaline pH values, while the solubility
for example, the RTO electrode surface, is also an open question. of pure TiO2 is predicted to be low under such conditions.145
Furthermore, in studies where the influence of anodic Despite these predictions for the pure oxides, the lifetime of
hypochlorite decomposition has been considered under industrial RuO2−TiO2 mixed-oxide DSAs is about 10 years.253
chlorate-process-like conditions, some studies find that the rate Additionally, the anode potential of electrodes in operation does
of oxygen evolution shows characteristics of mass transport not change until the RuIV content has become very low.226 The
limitations (even on graphite),54 while others find it shows constancy in activity during corrosion could perhaps be seen as
characteristics of kinetic limitations (on RTO).57 Studies have support for our recent suggestion that also Ti sites can be active
found that the reaction could make up as much as 40−80% of the in DSAs and that only low concentrations of Ru are necessary to
overall oxygen production, even in membrane chlor-alkali achieve these effects.64 Perhaps it might even be conceivable that
cells.159 Wanngård57 found that if the homogeneous decom- the last sites that are active on DSA are Ti sites activated by Ru, as
position reaction can be avoided (reaction 14), the anodic these sites might be more stable.65
decomposition of hypochlorite (the reaction where hypochlorite While not directly relevant to the selectivity issue, some
is decomposed electrochemically on the anode surface, proposed conclusions can be mentioned regarding the supposed
by Kotowski and Busse159 to proceed via reaction 23) is the main deactivation of DSA due to formation of a nonconductive
contributor to the oxygen evolved under chlorate conditions.57 interlayer of TiO2, between the support and the coating, during
However, the actual mechanism of the decomposition, whether electrolysis.223,236,237,239,254 This deactivation mechanism is
hypochlorous acid or hypochlorite is the main reactant, is still not frequently mentioned, but there seems to be little direct proof
known. Physics-based macroscopic modeling has suggested that of it being an important deactivation mechanism for electrodes
hypochlorite (OCl−) is the most probable reactant,176 while used in chlor-alkali or sodium chlorate production. It has been
estimations based on a probable low pH at the anode surface suggested that the deactivation is more likely the result of
suggest that HOCl should be the reactant.170 The recent study of complete removal of Ru from the surface layers of the
Macounová et al.,84 suggesting the possibility of radical reactions coating.226,227 Indeed, the fact that reactivation by Pt deposition
involving ClO· during oxidation of hypochlorite, also indicates onto the surface of used industrial electrodes seems to be
the need for further research on the topic. possible supports this claim.2
3.3. Connection between Anode Stability and Selectivity
We would like to point out that much of the detailed
knowledge about the connection between selectivity and stability
The literature clearly shows the connection between the rate of has resulted from the radiochemical studies performed almost
DSA degradation and the rate of the OER. However, it must be exclusively by scientists in the Soviet Union between 1960 and
pointed out that the relationship is more complex than it may 1990. These methods have allowed the corrosion of the anode to
seem. In chloride-containing electrolyte, at very low pH values be studied in situ during electrolysis at industrially relevant
below 2, degradation due to the formation of soluble chlorides is current densities and electrolyte composition and not only based,
likely a dominating mechanism of anode degradation.213 At these for example, on an increase in potential in accelerated lifetime
pH values, the corrosion rate of Ti has been found to exceed that tests. The latter type of stability test, which gives very little
of Ru.229 The rate of corrosion of DSA has been found to be understanding of the character of the degradation during
minimal in the pH range of 2−3,151 which agrees with the area of electrolysis, has been the dominating one employed in the
maximal stability according to the Pourbaix diagram.214 As the West even up until year 2015. The reason is perhaps the practical
pH is increased, the rate of corrosion increases in a similar rate as challenge in employing the radiochemical method, as proper
the OER.163,217,218 In this area, the corrosion rate of Ti has been safeguards must be employed, and possibly also the long time
found to be constant, but the Ru corrosion rate increases periods needed to study anode degradation at industrially
significantly.229 It may be noted that this increase coincides with relevant conditions. The study of Gajić-Krstajić et al.,239 which
the chlorine-hypochlorite equilibrium being pushed toward employed a spectrophotometric method to allow similarly
hypochlorite, see reaction 4 and Figure 1. The oxidizing power of sensitive detection of Ru as afforded by the radiochemical
hypochlorite, and the fact that hypochlorite, as opposed to method, could perhaps be used to avoid handling radioactive
chlorine, does not form gas bubbles in the electrolyte, may materials and give time-resolved understanding of the degrada-
accelerate the oxidation of ruthenium. There is a clear link tion process. Furthermore, the recent studies of Mayrhofer and
between OER and RuIV corrosion in the pH range relevant to co-workers240−243 and Chorkendorff and co-workers245,246 have
especially industrial chlor-alkali production (pH 2 to 4). While shown how ICP-MS and EQCM can be used to monitor
less data is available under pH values suitable for chlorate degradation of DSAs during use as OER catalysts. Nevertheless,
production, it is likely that the trend is maintained even there.231 long-term tests in the correct conditions, so that the effect of the
However, there is no clear link between the rate of ClER and that relevant reactions are studied, are still necessary. Additionally, the
of RuIV corrosion in the pH range relevant for chlor-alkali or clear link between OER and degradation that the literature
chlorate production. At higher pH values (above 10), the rate of indicates means that, for example, an accelerated lifetime test in
corrosion of both TiIV and RuIV increase,229 possibly the result of sulfuric acid, where OER is the only reaction occurring,
formation of metal oxyhydroxides.142,213,230 constitutes a test of a completely different property than what
Literature indicates that the surface layers of electrodes used in is relevant for an anode which is to evolve almost exclusively
chlor-alkali or sodium chlorate production will be stripped of chlorine gas.
RuIV. This has been shown directly both radiochemically229 and As seen in several studies (refs 74, 164, 174, 219, 222, 223, 231,
by, for example, selective ion mass spectrometry or and 238), there is a change in the Tafel slope for OER and ClER
AES,235−237,252 and indirectly by other means.226,227 This is in on Ru-based anodes at the critical anode potential, Ecr. This
agreement with thermodynamic predictions,214 which indicate potential is still not well understood but has been related to the
that soluble Ru compounds should form during anodic formation of RuO4.219 Indeed, the Pourbaix diagram indicates
polarization of pure RuO2 in concentrated chloride solutions, the connection between Ecr and the start of formation of
3020 DOI: 10.1021/acs.chemrev.5b00389
Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

ruthenate species.214 Furthermore, an increase in the rate of Ru 4. CONCLUSIONS


dissolution from DSA has been noted at electrode potentials It is clear that the issue of the selectivity between the main
above Ecr.231 Thus, operation at anode potentials higher than Ecr reactions in chlor-alkali and chlorate processes and the oxygen-
should be avoided to limit DSA corrosion. From an industrial evolving side reactions requires further study, as has also been
perspective, the risk of exceeding Ecr is higher in the chlorate recommended by Wanngård168 and Tilak and Chen.5 While the
process than in the chlor-alkali process.238 This is partly effect of several factors, most importantly of the current density,
connected to the decrease in Ecr which occurs as pH is are relatively well-known, the details of the effects of process
increased223 and which in turn likely is related to the pH effect on parameters, electrolyte composition, and electrode character-
the potential required for ruthenate formation, which follows the istics on the selectivity issue, as well as the mechanisms and
same trend.214 Interestingly, Eberil et al.74 found an optimum in relative importance of the suggested oxygen-forming reactions,
selectivity for chlorate production close to Ecr. Another are still not well-elucidated. Several studies have been conducted
important industrial aspect is the finding that Ru corrosion under conditions similar to those used in industrial chlor-alkali
increases by repeated interruptions in the electrolysis.224 Thus, production, while fewer studies have examined the selectivity
repeated starts and stops should be avoided. under conditions similar to those used during industrial chlorate
In closing, we would like to provide a discussion of how the production. For chlor-alkali production, a practical way of
lifetime of industrial DSA can be maximized in the intermediate increasing the selectivity through acidification of the electrolyte is
pH range relevant for chlor-alkali and sodium chlorate formation. available for state-of-the-art membrane cells. However, a higher
The difference in corrosion behavior between the Ru and Ti pH is necessary to facilitate chlorate formation, thus excluding
component of DSA in the intermediate pH range is most likely the method of acidification in industrial chlorate production, and
due to the possibility of forming ruthenates and perruthenates.213 oxygen evolution still accounts for about 5% current efficiency
Ti does not form such compounds.230 The formation of gaseous loss. Additionally, the possible influence of the critical anode
RuVIII products during electrolysis using DSA is well-known both potential, Ecr, on the selectivity for chlorate production is still not
well understood. Its appearance at high, industrially relevant,
in literature (see, for example, refs 151, 154, 221, 224, and 229)
current densities makes it imperative that future studies employ
and in industrial chlor-alkali and sodium chlorate production.
accurate iR compensation methods. Such methods could then be
This aspect has devoted significant experimental work, with, for combined with measurements of both the gas and the liquid
example, the study of Kötz et al.212 clearly showing the phase composition which might also give further understanding
connection between the onset of OER and the onset of RuO4 for the relative importance of the different oxygen-evolving
formation on pure RuO2 electrodes. It is reasonable to assume reactions.
that the formation of RuO4 occurs via an intermediate that is For both processes, modern experimental and theoretical
shared with the OER, as surface oxides need to form in both methods should be used to gain a deeper understanding of the
cases. Both the rate of OER and the percentage of the Ru released interplay between anode activity and selectivity, composition,
into the gas phase increases with pH.151,154 Furthermore, it was and structure. The combination of theoretical modeling and
clearly shown by Pecherskii et al.67 that Ru electrodes that exhibit modern characterization methods (e.g., the different X-ray
improved selectivity for ClER also exhibit improved stability. spectroscopic methods available today) is well-suited to
Even for other Ru mixed oxides, such as Ru−Sn oxides, improved exploring the details of heterogeneous and homogeneous
chlorine evolution selectivity69,71,172 is associated with improved (electro)catalytic reactions. These tools now allow for detailed
stability.234 RuIV corrosion rates have been found to decrease understanding of the electronic structure of practical catalysts,
significantly with increasing Cl− concentration.154,218,231 As has whereas before mostly macroscopic parameters regarding
been discussed in the preceding sections, increasing the Cl− structure and composition of electrodes have been possible to
concentration is a very simple way of increasing the current study in detail. The understanding resulting from studies
efficiency for Cl2. Increasing the current density is another way of combining theory and experiments could then be used to
improving the current efficiency for Cl2, at least as long as the improve the selectivity and activity of electrodes used in chlor-
density is not so high as to, for example, result in mass transport alkali and sodium chlorate production. They could also be used
limitations, and this also results in decreased current efficiency to start exploring the details of the relatively poorly understood
for corrosion.154,221,224,231 Increasing the temperature decreases bulk-phase reactions, where catalytic processes are involved both
the selectivity for chlorine evolution and increases the rate of in unwanted oxygen evolution reactions as well as in the
corrosion of DSA.154,163 However, a higher temperature might formation of sodium chlorate. The latter aspect is especially
still be viable, as it improves the kinetics of the chlorine- and interesting, as an understanding of the catalysis of sodium
chlorate formation could result in new ways of accelerating the
chlorate-producing reactions and decreases the cell potential
reaction, with important consequences for sodium chlorate
difference. At pH values relevant for industrial chlor-alkali and
process design.
sodium chlorate formation, high oxygen production rates are
indicative of a high rate of RuIV corrosion due to the formation of AUTHOR INFORMATION
ruthenates or perruthenates. Thus, chlor-alkali production
Corresponding Author
should operate in the pH range of 2−3 to achieve a maximal
chlorine current efficiency41,151 and anode lifetime.151 For *E-mail: amco@kth.se.
chlorate production, the optimal pH for current efficiency of Notes
chlorate production is significantly higher, at about pH 6−7. In The authors declare no competing financial interest.
this pH range, it is not possible to maximize both chlorate
production efficiency and DSA stability in the same way, and Biographies
both factors are connected to the critical potential. Anodes in Rasmus K. B. Karlsson received his Master’s degree in Chemistry and
both processes should be operated below the critical potential. Chemical Engineering in 2011 and his Ph.D. in Chemical Engineering in

3021 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

2015, both from KTH Royal Institute of Technology. His Ph.D. thesis CV cyclic voltammetry
research was primarily focused on combining experimental and differential electrochemical mass spectrosco-
theoretical methods to elucidate the fundamental mechanisms deciding DEMS py
the selectivity between chlorine and oxygen evolution in the chlor-alkali DSA dimensionally stable anode
and chlorate processes. EDX energy-dispersive X-ray spectroscopy
Ann Cornell received her MSc in Chemical Engineering in 1987 and EIS electrochemical impedance spectroscopy
then worked for 15 years in industry with research and development EQCM electrochemical quartz crystal microbalance
within electrochemical engineering, mainly related to industrial chlorate EXAFS extended X-ray absorption fine-structure
electrolysis. She completed her Ph.D. in 2002 at KTH with a thesis on spectroscopy
electrode reactions in the chlorate process and has, since 2003, worked GC gas chromatography
in the group of Applied Electrochemistry at KTH. Ann is now an ICP-MS inductively coupled plasma mass spectrome-
associate professor and still conducts research related to industrial try
electrolysis but has also broadened her interests to areas such as MS mass spectrometry
cellulose-based Li-ion batteries. OER oxygen evolution reaction
ORTA ruthenium−titanium oxide anode, abbrevia-
tion frequently used in Soviet and Russian
ACKNOWLEDGMENTS literature
The financial support of the Swedish Energy Agency and of RDE rotating disc electrode
Permascand AB are hereby acknowledged. Furthermore, the RHE relative hydrogen electrode, the SHE value
constructive comments given by Dr. John Gustavsson adjusted for the pH of the solution (i.e., that
(Permascand AB) on a draft of this review are gratefully measured with a SHE immersed in the same
acknowledged. Staffan Sandin is acknowledged for bringing some solution as the working electrode)
of the papers on decomposition of hypochlorous acid species to RTO ruthenium−titanium oxide
our attention. Professor Kiyotaka Asakura (the Catalysis SCE saturated calomel electrode
Research Center of Hokkaido University, Japan) is acknowl- SECM scanning electrochemical microscopy
edged for assistance with providing a high-resolution version of SEM scanning electron microscopy
Figure 8. Professor Lars G. M. Pettersson (Stockholm SHE standard hydrogen electrode
University) is gratefully acknowledged for constructive com- SIMS secondary ion mass spectrometry
ments on section 2.2.5.2. TEM transmission electron microscopy
TNT TiO2 nanotube
NOMENCLATURE XANES X-ray absorption near-edge spectroscopy
η overpotential, V XAS X-ray absorption spectroscopy
Φep current efficiency (or charge yield) of product XPS X-ray photoelectron spectroscopy
p XRD X-ray diffraction spectroscopy
ac adsorbate a on a CUS of the rutile (110)
surface
abr adsorbate a on a bridge site of the rutile (110)
surface REFERENCES
ai activity of ion i (1) Schmittinger, P.; Florkiewicz, T.; Curlin, L. C.; Lüke, B.; Scannell,
ci concentration of species i, mol/dm3 R.; Navin, T.; Zelfel, E.; Bartsch, R. Ullmann’s Encyclopedia of Industrial
E electrode potential, V versus SHE, or energy Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim,
(e.g., of adsorption), kJ Germany, 2012.
E0,i equilibrium (or reversible) electrode potential (2) O’Brien, T. F.; Bommaraju, T. V.; Hine, F. Vol. I: Fundamentals;
for reaction i, V versus SHE Springer, 2005; Vol. 1.
Ecr critical anode potential (3) Schlag, S. Sodium Chlorate. Chemical Economics Handbook Product
G Gibbs free energy, kJ/mol Review; 2012.
iR compensation Correction for the electrolyte resistance in (4) Fauvarque, J. The Chlorine Industry. Pure Appl. Chem. 1996, 68,
1713−1720.
electrochemical measurements (5) Tilak, B. V.; Chen, C.-P. In Chlor-alkali and chlorate technology: R.B.
j current density, A/m2 MacMullin Memorial Symposium; Burney, H. S., Furuya, N., Hine, F.,
ji partial current density for a certain reaction, Ota, K.-I., Eds.; The Electrochemical Society, 1999.
A/m2 (6) Viswanathan, K.; Tilak, B. V. Chemical, Electrochemical, and
j0 exchange current density, A/m2 Technological Aspects of Sodium Chlorate Manufacture. J. Electrochem.
jcr critical current density, the current density at Soc. 1984, 131, 1551−1559.
Ecr (7) Vogt, H.; Balej, J.; Bennett, J. E.; Wintzer, P.; Sheikh, S. A.; Gallone,
q* electrochemically active surface area, C/m2 P.; Vasudevan, S.; Pelin, K. Ullmann’s Encyclopedia of Industrial
t temperature in degrees centigrade Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim,
U electrode potential Germany, 2010.
(8) Beer, H. B. Method of Making an Electrode Having a Coating
AES atomic emission spectroscopy Containing a Platinum Metal Oxide Thereon. U.S. Patent 4052271,
AFM atomic force microscopy 1977.
BEP Brønsted−Evans−Polanyi (9) Roginskaya, Y. E.; Galyamov, B. S.; Lebedev, V. M.; Belova, I. D.;
CHE computational hydrogen electrode Venevtsev, Y. N. Phase Composition and Electrical Transport
ClER chlorine evolution reaction Properties of the RuO2-TiO2 System. Zh. Neorg. Khim. 1977, 22,
CUS coordinatively unsaturated site 499−504.

3022 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

(10) Hine, F.; Yasuda, M.; Yoshida, T. Studies on the Oxide-Coated (38) Brinkmann, T.; Santonja, G. G.; Schorcht, F.; Roudier, S.; Sanco,
Metal Anodes for Chlor-Alkali Cells. J. Electrochem. Soc. 1977, 124, 500− L. D. Best Available Techniques (BAT) Reference Document for the
505. Production of Chlor-alkali; JRC Science and policy reports EUR 26844
(11) Gerrard, W. A.; Steele, B. C. H. Microstructural Investigations on EN; Industrial Emissions Directive 2010/75/EU (Integrated pollution
Mixed RuO2−TiO2 Coatings. J. Appl. Electrochem. 1978, 8, 417−425. prevention and control), 2014.
(12) Over, H. Surface Chemistry of Ruthenium Dioxide in (39) Eurochlor, Chlorine Industry Review 2013−2014; 2014.
Heterogeneous Catalysis and Electrocatalysis: from Fundamental to (40) Wendt, H.; Vogt, H.; Kreysa, G.; Kolb, D. M.; Engelmann, G. E.;
Applied Research. Chem. Rev. 2012, 112, 3356−3426. Ziegler, J. C.; Goldacker, H.; Jüttner, K.; Galla, U.; Schmieder, H. e.
(13) Kavan, L.; Grätzel, M.; Gilbert, S. E.; Klemenz, C.; Scheel, H. J. Ullmanna’s Encyclopedia of Industrial Chemistry [Online]; Wiley-VCH:
Electrochemical and Photoelectrochemical Investigation of Single- Weinheim, Germany, 2009; Chapter Electrochemistry (accessed May
Crystal Anatase. J. Am. Chem. Soc. 1996, 118, 6716−6723. 20, 2011).
(14) Glassford, K. M.; Chelikowsky, J. R. Electronic Structure of (41) Bergner, D. Reduction of By-product Formation in Alkali
TiO2:Ru. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 12550− Chloride Membrane Electrolysis. J. Appl. Electrochem. 1990, 20, 716−
12553. 722.
(15) García-Mota, M.; Vojvodic, A.; Abild-Pedersen, F.; Nørskov, J. K. (42) Valensi, G.; Deltombe, E.; De Zoubov, N.; Vanleugenhaghe, C.;
Electronic Origin of the Surface Reactivity of Transition Metal Doped Pourbaix, M. In Atlas of Electrochemical Equilibria In Aqueous Solutions;
TiO2(110). J. Phys. Chem. C 2013, 117, 460−465. Pourbaix, M., Ed.; Pergamon Press Ltd., 1966; Chapter IV, Section 20.0,
(16) Beer, B. H. Noble Metal Coated Titanium Electrode and Method pp 343−349.
of Making and Using It. U.S. Patent 3096272, 1963. (43) Leah, R. T.; Brandon, N. P.; Vesovic, V.; Kelsall, G. H. Numerical
(17) Beer, H. B. Electrode and Method of Making Same. U.S. Patent Modeling of the Mass Transport and Chemistry of a Simplified
3234110, 1966. Membrane-Divided Chlor-Alkali Reactor. J. Electrochem. Soc. 2000, 147,
(18) Beer, H. B. Electrolysis with Precious Metal-coated Titanium 4173.
Anode. U.S. Patent 3236756, 1966. (44) Cheng, C. Y.; Kelsall, G. H. Models of Hypochlorite Production in
(19) Beer, H. B. Electrode and Coating Therefore. U.S. Patent Electrochemical Reactors with Plate and Porous Anodes. J. Appl.
3632498, 1972. Electrochem. 2007, 37, 1203−1217.
(20) Beer, H. B. Electrode Having Platinum Metal Oxide Coating (45) Ping Ong, S.; Wang, L.; Kang, B.; Ceder, G. Li-Fe-P-O2 Phase
Thereon, and Method of Use Thereof. U.S. Patent 3711385, 1973. Diagram from First Principles Calculations. Chem. Mater. 2008, 20,
(21) Beer, H. B. Method of Making An Electrode Having A Coating 1798−1807.
containing a platinum metal oxide thereon. U.S. Patent 3864163, 1975. (46) Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.;
(22) Beer, H. B. Coating of Protected Electrocatalytic Material on an Persson, K. A.; Ceder, G. Formation Enthalpies by Mixing GGA and
Electrode. U.S. Patent 3933616, 1976. GGA + U Calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011,
(23) Hayfield, P. C. S. Development of the Noble Metal/oxide Coated 84, 045115−1−045115−10.
Titanium Electrode. Part I: the Beginning of the Story. Platinum Met. (47) Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.;
Rev. 1998, 42, 27−33. Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G. e.; Persson, K.
(24) Trasatti, S. Electrocatalysis: Understanding the Success of DSA. A. Commentary: the Materials Project: a Materials Genome Approach
Electrochim. Acta 2000, 45, 2377−2385. to Accelerating Materials Innovation. APL Mater. 2013, 1, 011002−1−
(25) Duby, P. The History of Progress in Dimensionally Stable 011002−11.
Anodes. JOM 1993, 45, 41−43. (48) Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.;
(26) Trasatti, S. In The Electrochemistry of Novel Materials; Lipkowski, Halow, I.; Bailey, S. M.; Churney, K. L.; Nuttall, R. L. The NBS Tables of
J., Ross, P. N., Eds.; VCH Publishers, 1994; Chapter 5, pp 207−295. Chemical Thermodynamic Properties: Selected Values for Inorganic
(27) Trasatti, S. Physical Electrochemistry of Ceramic Oxides. and C1 and C2 Organic Substances in SI Units. J. Phys. Chem. Ref. Data,
Electrochim. Acta 1991, 36, 225−241. Suppl. 1982, 11 (2), 1−407.
(28) Daghetti, A.; Lodi, G.; Trasatti, S. Interfacial Properties of Oxides (49) Wendt, H., Kreysa, G. Electrochemical Engineering; Springer
Used as Anodes in the Electrochemical Technology. Mater. Chem. Phys. Science + Business Media, 1999.
1983, 8, 1−90. (50) van der Stegen, J. H. G. The State of the Art of Modern Chlor
(29) Electrodes of Conductive Metallic Oxides. Part B; Trasatti, S., Ed.; Alkali Electrolysis with Membrane Cells. Ph.D. Thesis, University of
Elsevier, 1981. Twente, Holland, 2000.
(30) Electrodes of Conductive Metallic Oxides. Part A; Trasatti, S., Ed.; (51) Colman, J. E. Electrolytic Production of Sodium Chlorate. AIChE
Elsevier, 1980. Symp. Ser. 1981, 77, 244−263.
(31) Trasatti, S. Progress in the Understanding of the Mechanism of (52) De Valera, V. On the Theory of Electrochemical Chlorate
Chlorine Evolution at Oxide Electrodes. Electrochim. Acta 1987, 32, Formation. Trans. Faraday Soc. 1953, 49, 1338−1351.
369−382. (53) Landolt, D.; Ibl, N. On the Mechanism of Anodic Chlorate
(32) Hayfield, P. C. S. Development of the Noble Metal/oxide Coated Formation in Concentrated NaCl Solutions. Electrochim. Acta 1970, 15,
Titanium Electrode. Part II: the Move from Platinum/iridium to 1165−1183.
Ruthenium Oxide Electrocatalysts. Platinum Met. Rev. 1998, 42, 46−55. (54) Hammar, L.; Wranglén, G. Cathodic and Anodic Efficiency
(33) Hayfield, P. C. S. Development of the Noble Metal/oxide Coated Losses in Chlorate Electrolysis. Electrochim. Acta 1964, 9, 1−16.
Titanium Electrode. Part III: Coated Titanium Anodes inWidely (55) Jaksić, M. M.; Despić, A.; Csonka, I.; Nikolić, B. Studies on
Ranging Oxygen Evolving Situations. Platinum Met. Rev. 1998, 42, 116− Chlorate Cell Process V. Theory and Practice of a Modified Technology
122. for Electrolytic Chlorate Production. J. Electrochem. Soc. 1969, 116,
(34) Over, H. Atomic Scale Insights into Electrochemical Versus Gas 1316−1322.
Phase Oxidation of HCl Over RuO2-based Catalysts: a Comparative (56) Jaksić, M. M. Mutual Effect of Current Density, pH, Temperature,
Review. Electrochim. Acta 2013, 93, 314−333. and Hydrodynamic Factors on Current Efficiency in the Chlorate Cell
(35) Diebold, U. The Surface Science of Titanium Dioxide. Surf. Sci. Process. J. Electrochem. Soc. 1974, 121, 70−79.
Rep. 2003, 48, 53−229. (57) Wanngård, J. Oxygen Formation in Chlorate Electrolysis. Poster
(36) Bergner, D.; Hartmann, M. Chlorate and Oxygen Formation in presented at the Joint Meeting of Dechema and the Society of Chemical
Alkali Chloride Membrane Electrolysis. J. Appl. Electrochem. 1993, 23, Industry on Electrochemical Cell Design and Optimization Procedures,
103−107. 1990.
(37) Standard Potentials in Aqueous Solutions; Bard, A. J., Parsons, R., (58) Foerster, F.; Müller, E. Beiträge zur Theorie der Elektrolyse von
Jordan, J., Eds.; CRC Press, 1985. Alkalichloridlösungen. Z. Electrochem. 1903, 9, 171−185.

3023 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

(59) Foerster, F. The Electrolysis of Hypochlorite Solutions. Trans. (80) Macounová, K.; Makarova, M.; Krtil, P. Oxygen Evolution on
Am. Electrochem. Soc. 1924, 46, 23−50. Nanocrystalline RuO2 and Ru0.9Ni0.1O2−δ Electrodes - DEMS Approach
(60) Spasojević, M.; Ribić-Zelenović, L.; Spasojević, P. Microstructure to Reaction Mechanism Determination. Electrochem. Commun. 2009, 11,
of New Composite Electrocatalyst and Its Anodic Behavior for Chlorine 1865−1868.
and Oxygen Evolution. Ceram. Int. 2012, 38, 5827−5833. (81) Petrykin, V.; Bastl, Z.; Franc, J.; Macounova, K.; Makarova, M.;
(61) Ibl, N.; Landolt, D. On The Mechanism of Anodic Chlorate Mukerjee, S.; Ramaswamy, N.; Spirovova, I.; Krtil, P. Local Structure of
Formation in Dilute NaCl solutions. J. Electrochem. Soc. 1968, 115, 713− Nanocrystalline Ru1−xNixO2−δ Dioxide and Its Implications for Electro-
720. catalytic Behaviors - an XPS and XAS Study. J. Phys. Chem. C 2009, 113,
(62) Ibl, N.; Vogt, H. In Comprehensive Treatise of Electrochemistry; 21657−21666.
Bockris, J. O., Ed.; Plenum Press, 1981; Vol. 2, Chapter 3, pp 167−250. (82) Petrykin, V.; Macounová, K.; Okube, M.; Mukerjee, S.; Krtil, P.
(63) Exner, K. S.; Anton, J.; Jacob, T.; Over, H. Controlling Selectivity Local Structure of Co Doped RuO2 Nanocrystalline Electrocatalytic
in the Chlorine Evolution Reaction over RuO2-Based Catalysts. Angew. Materials for Chlorine and Oxygen Evolution. Catal. Today 2013, 202,
Chem., Int. Ed. 2014, 53, 11032−11035. 63−69.
(64) Karlsson, R. K.; Hansen, H. A.; Bligaard, T.; Cornell, A.; (83) Halck, N. B.; Petrykin, V.; Krtil, P.; Rossmeisl, J. Beyond the
Pettersson, L. G. Ti Atoms in Ru0.3Ti0.7O2 Mixed Oxides Form Active Volcano Limitations in Electrocatalysis−Oxygen Evolution Reaction.
and Selective Sites for Electrochemical Chlorine Evolution. Electrochim. Phys. Chem. Chem. Phys. 2014, 16, 13682−13688.
Acta 2014, 146, 733−740. (84) Macounová, K. M.; Simic, N.; Ahlberg, E.; Krtil, P. Electro-
(65) Karlsson, R. K. B.; Cornell, A.; Pettersson, L. G. M. The chemical Water-Splitting Based on Hypochlorite Oxidation. J. Am.
Electrocatalytic Properties of Doped TiO2. Electrochim. Acta 2015, 180, Chem. Soc. 2015, 137, 7262−7265.
514−527. (85) Chapin, R. M. The Effect of Hydrogen-Ion Concentration on the
(66) Buné, N.; Shilyaeva, G.; Losev, V. Electrode Kinetics of Side Decomposition of Hypohalites. J. Am. Chem. Soc. 1934, 56, 2211−2215.
Reactions Oxygen Evolution from Chloride Solutions at Titanium- (86) Lister, M. Decomposition of Sodium Hypochlorite: the
ruthenium Oxide Anodes. Elektrokhimiya 1977, 13, 1540−1546. Uncatalyzed Reaction. Can. J. Chem. 1956, 34, 465−478.
(67) Pecherskii, M.; Gorodetskii, V.; Buné, N.; Losev, V. Kinetics of the (87) Bhaduri, I. Ü ber die Umsetzung der Hypochlorite in Chlorate. Z.
Side Electrode Processes on Ruthenium-titanium Anodes in Chloride Anorg. Chem. 1897, 13, 385−406.
Solutions. Elektrokhimiya 1982, 18, 415−422. (88) Lister, M. W. The Decomposition of Hypochlorous Acid. Can. J.
(68) Buné, N.; Losev, V. V.; Reznik, M. F.; Zaripova, E. Selectivity and Chem. 1952, 30, 879−889.
Electrochemical Behavior of Titanium-ruthenium Oxide Anodes (89) Lister, M. W.; Petterson, R. C. Oxygen Evolution from Sodium
Containing Different Levels of Ruthenium Dioxide. Elektrokhimiya Hypochlorite Solutions. Can. J. Chem. 1962, 40, 729−733.
1986, 22, 396−398. (90) Church, J. A. Kinetics of the Uncatalyzed and Copper(II)-
(69) Tilak, B. V.; Tari, K.; Hoover, C. Metal Anodes and Hydrogen catalyzed Decomposition of Sodium Hypochlorite. Ind. Eng. Chem. Res.
Cathodes: Their Activity Towards O2 Evolution and ClO3−Reduction 1994, 33, 239−245.
Reactions. J. Electrochem. Soc. 1988, 135, 1386−1392. (91) Adam, L. C.; Gordon, G. Hypochlorite Ion Decomposition:
(70) Kuhn, A. T.; Mortimer, C. J. The Efficiency of Chlorine Evolution Effects of Temperature, Ionic Strength, and Chloride Ion. Inorg. Chem.
in Dilute Brines on Ruthenium Dioxide Electrodes. J. Appl. Electrochem. 1999, 38, 1299−1304.
1972, 2, 283−287. (92) Sandin, S.; Karlsson, R. K. B.; Cornell, A. The Catalyzed and
(71) Spasojević, M.; Krstajić, N.; Jaksić, M. A Selective Catalyst for Uncatalyzed Decomposition of Hypochlorite. Ind. Eng. Chem. Res. 2015,
Titanium Anodes: Development and Optimization: II. Selectivity 54, 3767−3774.
Features. J. Res. Inst. Catal., Hokkaido Univ. 1984, 32, 29−36. (93) Howell, O. R. The Catalytic Decomposition of Sodium
(72) Spasojević, M.; Krstajić, N.; Jaksić, M. Electrocatalytic Hypochlorite by Cobalt Peroxide. Proc. R. Soc. London, Ser. A 1923,
Optimization of Faradaic Yields in the Chlorate Cell Process. Surf. 104, 134−152.
Technol. 1984, 21, 19−26. (94) White, A. D. Action of Solutions of Bleaching Powder and of
(73) Krstajic, N.; Spasojević, M.; Jaksic, M. A Selective Catalyst for Hypochlorous acid on Metals. J. Soc. Chem. Ind. 1903, 22, 132−134.
Titanium Anodes: Development and Optimization I. Catalyst Structure, (95) Hofmann, K. A.; Ritter, K. Beständigkeit und Oxydationspotential
Activity and Durability. J. Res. Inst. Catal., Hokkaido Univ. 1984, 32, 19− der Hypochlorite, Beiträge zur Katalyse und über ein Hypochlorit-
28. Kohle-Element. Ber. Dtsch. Chem. Ges. 1914, 47, 2233−2244.
(74) Eberil, V. I.; Fedotova, N. S.; Novikov, E. A.; Mazanko, A. F. (96) Chirnoaga, E. CCXXII. - the Catalytic Decomposition of
Studying the Link Between the Potential of a Metal-oxide Anode, the Solutions of Sodium Hypochlorite by Finely Divided Metallic Oxides.
Current Efficiency for Chlorate, and the Current Losses for the Oxygen J. Chem. Soc. 1926, 129, 1693−1703.
and Chlorine Evolution in a Wide Range of the Chlorate Electrolysis (97) Lewis, J. R. The Catalytic Decomposition of Sodium
Conditions. Russ. J. Electrochem. 2000, 36, 1296−1302. Hypochlorite Solutions. I. Mechanism of the Reaction. J. Phys. Chem.
(75) Arikawa, T.; Murakami, Y.; Takasu, Y. Simultaneous Determi- 1927, 32, 243−254.
nation of Chlorine and Oxygen Evolving at RuO2/Ti and RuO2−TiO2/ (98) Lewis, J. R. The Catalytic Decomposition of Sodium
Ti Anodes by Differential Electrochemical Mass Spectroscopy. J. Appl. Hypochlorite Solutions. III. Promoter Action of Hydrated Magnaesium
Electrochem. 1998, 28, 511−516. Oxide in the Hydrated Copper Oxide Catalysis of Sodium Hypochlorite.
(76) Jirkovský, J.; Hoffmannová, H.; Klementová, M.; Krtil, P. Particle J. Phys. Chem. 1930, 35, 915−919.
Size Dependence of the Electrocatalytic Activity of Nanocrystalline (99) Connick, R. E.; Hurley, C. R. Chemistry of Ru(VI), -(VII) and
RuO2 Electrodes. J. Electrochem. Soc. 2006, 153, E111−E118. -(VIII). Reactions, Oxidation Potentials and Spectra. J. Am. Chem. Soc.
(77) Macounová, K.; Makarova, M.; Jirkovský, J.; Franc, J.; Krtil, P. 1952, 74, 5012−5015.
Parallel Oxygen and Chlorine Evolution on Ru1−xNixO2−y Nano- (100) Patel, M. C.; Mankad, B. N. A Note on the Effect of Salts on the
structured Electrodes. Electrochim. Acta 2008, 53, 6126−6134. Decomposition of Sodium Hypochlorite. J. Indian Chem. Soc., Ind. News
(78) Makarova, M. V.; Jirkovský, J.; Klementová, M.; Jirka, I.; Ed. 1954, 17, 236−237.
Macounová, K.; Krtil, P. The Electrocatalytic Behavior of Ru0.8Co0.2O2−x (101) Ayres, G. H.; Booth, M. H. Catalytic Decomposition of
- the Effect of Particle Shape and Surface Composition. Electrochim. Acta Hypochlorite Solution by Iridium Compounds. I. The pH-Time
2008, 53, 2656−2664. Relationship. J. Am. Chem. Soc. 1955, 77, 825−827.
(79) Macounová, K.; Jirkovský, J.; Makarova, M. V.; Franc, J.; Krtil, P. (102) Ayres, G. H.; Booth, M. H. Catalytic Decomposition of
Oxygen Evolution on Ru1−xNixO2−y Nanocrystalline Electrodes. J. Solid Hypochlorite Solution by Iridium Compounds. II. Kinetic Studies. J.
State Electrochem. 2009, 13, 959−965. Am. Chem. Soc. 1955, 77, 828−833.

3024 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

(103) Lister, M. Decomposition of Sodium Hypochlorite: the (128) Müller, E. Experimentaluntersuchung über die Bildung von
Catalyzed Reaction. Can. J. Chem. 1956, 34, 479−488. Hypochlorit und Chlorat bei der Elektrolyse von Alkalichloriden. Z.
(104) Gray, E. T.; Taylor, R. W.; Margerum, D. W. Kinetics and Anorg. Chem. 1899, 22, 33−90.
Mechanisms of the Coppercatalyzed Decomposition of Hypochlorite (129) Müller, E.; Koppe, P. Einfluss der Stromkonzentration auf die
and Hypobromite. Properties of a Dimeric Copper(III) Hydroxide Elektrolytische Chloratbildung. Z. Elektrochem. 1911, 17, 421−430.
Intermediate. Inorg. Chem. 1977, 16, 3047−3055. (130) Knibbs, N.; Palfreeman, H. The Theory of Electrochemical
(105) Hamano, A.; Mori, H. Activity of Metal Oxide Catalysts. 4. Chlorate and Perchlorate Formation. Trans. Faraday Soc. 1920, 16,
Activity and Surface Area of Transition Metal Oxide Catalysts. Sasebo 402−433.
Kogyo Koto Senmon Gakko Kenkyu Hokoku 1977, 14, 65−72. (131) Joffe, W. S. Ü ber den Anodenprozess bei der Elektrolyse
(106) Hamano, A.; Takakura, K.; Nakamori, I. The Effect of pH on the wässriger Alkalichloridlösungen mit imprägnierten und nichtimpräg-
Catalytic Decomposition of Aqueous Sodium Hypochlorite by Cobalt nierten Graphitelektroden. II: Gemeinsame Abscheidung von Chlor
Oxide. Kayaku Gakkaishi 1998, 59, 174−180. und Sauerstoff an der Anode. Zeitschr. Electrochem. 1936, 42, 76−79.
(107) Baldwin, R. T. History of the Chlorine Industry. J. Chem. Educ. (132) Rius, A.; Llopis, J. Sobre La Oxidacion Anodica De Los
1927, 4, 313−319. Hipocloritos Alcalinos. Anales De Fisica y Quimica 1945, 41, 1030−1053.
(108) Kershaw, J. B. C. The Present Position and Future Prospects of (133) de Valera, V. The Mechanism of Chloride Electrolysis. Trans.
the Electrolytic Alkali and Bleach Industry. Trans. Faraday Soc. 1907, 3, Faraday Soc. 1956, 52, 250−260.
38−47. (134) Despić, A.; Jaksić, M.; Nikolić, B. The Effect of Kinetic and
(109) Müller, E. Alkalichloridelektrolyse. Z. Elektrochem. Angew. Phys. Hydrodynamic Factors on Current Efficiency in the Chlorate Cell
Chem. 1907, 13, 800−804. Process. J. Appl. Electrochem. 1972, 2, 337−343.
(110) Oettel, F. Studien Ü ber Die Elektrolytische Bildung Von (135) Jaksić, M. M.; Nikolić, B.; Csonka, I.; Djordjević, A. Studies on
Unterchlorig-sauren Und Chlorsauren Salzen. Z. Elektrochem. Angew. Chlorate Cell Process IV. Effect of Neutral Salts on Conversion of
Phys. Chem. 1894, 1, 354−361. Available Chlorine to Chlorate. J. Electrochem. Soc. 1969, 116, 684−687.
(111) Oettel, F. Studien Ü ber Die Elektrolytische Bildung Von (136) Kuhn, A. T.; Mortimer, C. J. The Kinetics of Chlorine Evolution
Unterchlorig-sauren Und Chlorsauren Salzen. II. Z. Elektrochem. Angew. and Reduction on Titanium-SupportedMetal Oxides Especially RuO2
Phys. Chem. 1895, 1, 474−480. and IrO2. J. Electrochem. Soc. 1973, 120, 231−236.
(112) Oettel, F. Zur Elektrolyse von Chlorcalcium-Lösungen. Z. (137) Jaksić, M. M.; Despić, A.; Nikolić, B.; Maksić, S. Effect of Some
Elektrochem. Angew. Phys. Chem. 1898, 5, 1−5. Anions on the Chlorate Cell Process. Croat. Chem. Acta 1972, 44, 61−
(113) Bischoff, H.; Foerster, F. Beitrag Zur Kenntnis Der Elektrolyse 66.
Von Chlorcalciumlösungen. Z. Elektrochem. 1898, 4, 464−470. (138) Bondar, R.; Borisova, A. A.; Kalinovskii, E. A. Titanium-
(114) Foerster, F.; Müller, E.; Jorre, F. Zur Kenntnis der Vorgänge bei ruthenium Dioxide Anodes in the Electrolysis of Chloride-sulphate
der Elektrolyse der Alkalichloridlösungen. Z. Elektrochem. Angew. Phys. Solutions. Elektrokhimiya 1974, 10, 44−48.
(139) Veselovskaya, I. E.; Spasskaya, E. K.; Sokolov, V. A.; Takachenko,
Chem. 1899, 6, 11−26.
V. I.; Yakimenko, L. M. Titanium-ruthenium Dioxide Anodes in the
(115) Foerster, F.; Jorre, F. Zur Kenntniss der Beziehungen der
Electrolysis of Chloridesulphate Solutions. Elektrokhimiya 1974, 10,
unterchlorigsauren Salze zu den chlorsauren Salzen. J. Prakt. Chem.
70−73.
1899, 59, 53−101.
(140) Agapova, R. A.; Kokhanov, G. N. Electrochemical Properties of
(116) Foerster, F.; Sonneborn, H. Zur Kenntnis der anodischen
Cobalt Oxide Anodes. Elektrokhimiya 1976, 12, 1649−1653.
Sauerstoffentwicklung bei der Elektrolyse von Alkalichloridlösungen. Z.
(141) Shlyapnikov, V. A. Role of Anode Material in Electrosynthesis of
Elektrochem. Angew. Phys. Chem. 1900, 6, 597−604. Chlorates. J. Appl. Chem. USSR 1976, 49, 90−94.
(117) Foerster, F.; Jorre, F. Zur Kenntnis der Erscheinungen bei der (142) Uzbekov, A. A.; Lambrev, V. G.; Yazikov, I. F.; Buné, N. Y.;
Elektrolyse von Alkalichloridlösungen mit Diaphragma. Z. Anorg. Chem. Gorodetskii, V. V.; Losev, V. V. Possibility of Improving Chloride
1900, 23, 158−219. Electrolysis by Use of Titanium-supported Ruthenium Dioxide Anodes.
(118) Foerster, F. Zur Theorie der elektrolytischen Bildung von Elektrokhimiya 1978, 14, 1303.
Hypochlorit und Chlorat. Z. Anorg. Chem. 1899, 22, 1−32. (143) Kazarinov, V. E.; Andreev, V. N. Investigation of the Adsorption
(119) Foerster, F.; Müller, E. Zur Kenntnis der Elektrolyse, zumal der of Ions and the Structure of the Electric Double Layer on Ruthenium-
Alkalichloride, an platinierten Elektroden. Z. Elektrochem. Angew. Phys. titanium Oxide Anodes. Elektrokhimiya 1978, 14, 577−579.
Chem. 1902, 8, 515−540. (144) Fukuda, K.-i.; Iwakura, C.; Tamura, H. Effect of the Addition of
(120) Foerster, F.; Müller, E. Zur Theorie der Einwirkung der NH4F on Anodic Behaviors of DSA-type Electrodes in H2SO4-
Halogene auf Alkalien. Z. Elektrochem. Angew. Phys. Chem. 1902, 8, 921− (NH4)2SO4 Solutions. Electrochim. Acta 1979, 24, 367−371.
926. (145) Kelsall, G.; Robbins, D. Thermodynamics of Ti-H2O-F(-Fe)
(121) Foerster, F.; Müller, E. Ü ber die bei der elektrolytischen systems at 298 K. J. Electroanal. Chem. Interfacial Electrochem. 1990, 283,
Darstellung von Alkali- Hypochloriten und -Chloraten erreichbaren 135−157.
Strom- und Energie-Ausbeuten. Z. Elektrochem. Angew. Phys. Chem. (146) Iwakura, C.; Inai, M.; Tamura, H. Foreign Metal-doped SnO2
1902, 8, 8−17. Film Anodes for Oxygen and Chlorine Evolution. Chem. Lett. 1979, 8,
(122) Foerster, F.; Müller, E. Ü ber das Verhalten der unterchlorigen 225−228.
Säure und ihrer Salze bei der Elektrolyse. Z. Elektrochem. Angew. Phys. (147) Caldwell, D. L.; Hazelrigg, M. J. In Modern Chlor-Alkali
Chem. 1902, 8, 633−638. Technology; Coulter, M. O., Ed.; SCI, 1980; Vol. 1; Chapter 10, pp 137−
(123) Foerster, F. Ü ber kÜ nstlichen Graphit und über Platiniridium als 144.
Anodenmaterialien. Z. Elektrochem. Angew. Phys. Chem. 1902, 8, 143− (148) Saito, S. In Modern Chlor-Alkali Technology; Coulter, M. O., Ed.;
147. SCI, 1980; Vol. 1; Chapter 11, pp 137−144.
(124) Foerster, F.; Dolch, P. Die Umwandlung von Hypochlorit in (149) Harjanto, S.; Cao, Y.; Shibayama, A.; Naitoh, I.; Nanami, T.;
Chlorat in alkalischer Lösung. Z. Elektrochem. 1917, 23, 137−147. Kasahara, K.; Okumura, Y.; Liu, K.; Fujita, T. Leaching of Pt, Pd and Rh
(125) Foerster, F.; Müller, E. Ü ber das Verhalten der unterchlorigen from Automotive Catalyst Residue in Various Chloride Based Solutions.
Säure und ihrer Salze bei der Elektrolyse. Z. Elektrochem. Angew. Phys. Mater. Trans. 2006, 47, 129−135.
Chem. 1902, 8, 665−672. (150) Bergner, D. Stromausbeuten, Chlorat- und Sauerstoffbildung bei
(126) Sproesser, L. Ü ber Alkalichlorid-Elektrolyse an Kohlenanoden. der Alkali-chlorid- Elektrolyse nach dem Membranverfahren. Chem.-Ztg.
Z. Elektrochem. Angew. Phys. Chem. 1901, 7, 971−976. 1980, 104, 215−224.
(127) Oechsli, W. Ü ber die Elektrolytische Perchloratbildung. Z. (151) Gorodetskii, V. V.; Pecherskii, M. M.; Yanke, V. B.; Buné, N. Y.;
Elektrochem. Angew. Phys. Chem. 1903, 9, 807−828. Busse-Machukas, V. B.; Kubasov, V. L.; Losev, V. V. Effect of Acidity on

3025 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

the Electrochemical and Corrosion Behavior of Titanium-ruthenium Mathemeatical and Computational Methods in Electrochemical Engineer-
Oxide Anodes in Chloride Solutions. Elektrokhimiya 1981, 17, 513−517. ing; Electrochemical Society: Pennington, NJ, 1993; 147−154.
(152) Érenburg, R. G.; Krishtalik, L. I.; Yaroshevskaya, I. P. Kinetics of (173) Baolian, Y.; Wenhua, Z. Current Efficiency in the Chlorate Cell
chlorine evolution and ionization at ruthenium oxide and ruthenium- Process with an Oxygen Cathode (III): a Gas Analysis Method for
titanium oxide anodes. Elektrokhimiya 1975, 11, 1072−1074. Chlorate Current. J. Appl. Electrochem. 1994, 24, 503−508.
(153) O’Grady, W.; Iwakura, C.; Huang, J.; Yeager, E. Report Number (174) Cornell, A.; Håkansson, B.; Lindbergh, G. Ruthenium Based
37: Ruthenium Oxide Catalysts for the Oxygen Electrode; Office of Naval DSA® in Chlorate Electrolysis - Critical Anode Potential and Reaction
Research: Department of Chemistry, Case Western Reserve University: Kinetics. Electrochim. Acta 2003, 48, 473−481.
Cleveland, Ohio, 1974. (175) Gritzner, G.; Kreysa, G. Nomenclature, Symbols and Definitions
(154) Denton, D. A.; De Souza, J. T.; Entwisle, J. H.; Lee, D.; Wilson, in Electrochemical Engineering. Pure Appl. Chem. 1993, 65, 1009−1020.
H. G. In Developments in Coatings for Metal Anodes; Jackson, C., Ed.; Ellis (176) Byrne, P.; Fontes, E.; Parhammar, O.; Lindbergh, G. A
Horwood: Chichester, 1983; Vol. 2, Chapter 13. Simulation of the Tertiary Current Density Distribution from a
(155) Kokoulina, D.; Bunakova, L. Oxygen Evolution at the Oxide Chlorate Cell: I. Mathematical Model. J. Electrochem. Soc. 2001, 148,
Electrodes RuO2 and RuO2+TiO2 in Chloride Solutions. J. Electroanal. D125−D132.
Chem. Interfacial Electrochem. 1984, 164, 377−383. (177) Byrne, P. Mathematical Modelling and Experimental Simulation
(156) Trasatti, S. Electrocatalysis in the Anodic Evolution of Oxygen of Chlorate and Chloralkali Cells. Ph.D. Thesis, KTH The Royal
and Chlorine. Electrochim. Acta 1984, 29, 1503−1512. institute of technology, Stockholm, Sweden, 2001.
(157) Buné, N. Y.; Portnova, M. Y.; Filatov, V. P.; Losev, V. V. Effect of (178) Santana, M. H.; De Faria, L. A. Oxygen and Chlorine Evolution
Foreign Anions on the Kinetics of Chlorine and Oxygen Evolution on on RuO2+TiO2+CeO2+ Nb2O5 Mixed Oxide Electrodes. Electrochim.
the Ruthenium-titanium Oxide Anodes Under the Conditions of Acta 2006, 51, 3578−3585.
Chlorine Evolution. Elektrokhimiya 1984, 20, 1291−1295. (179) Panić, V.; Jovanović, V.; Terzić, S.; Barsoum, M.; Jović, V.;
(158) Cairns, J. F.; Couper, A. M.; Denton, D. A. In Modern Chlor- Dekanski, A. The Properties of Electroactive Ruthenium Oxide
alkali Technology; Wall, K., Ed.; Ellis Horwood Publishers, 1986; Vol. 3, Coatings Supported by Titanium-based Ternary Carbides. Surf. Coat.
Chapter 21, pp 293−308. Technol. 2007, 202, 319−324.
(159) Kotowski, S.; Busse, B. In Modern Chlor-alkali Technology; Wall, (180) Macounová, K.; Makarova, M.; Franc, J.; Jirkovský, J.; Krtil, P.
K., Ed.; Ellis Horwood Publishers, 1986; Vol. 3, Chapter 22, pp 310− Influence of Oxygen on Reactivity of Ru1−xFexO2−y-Doped Materials.
323. Electrochem. Solid-State Lett. 2008, 11, F27−F29.
(160) Wranglén, G. Kloratelektrolysens Teori. Tek. Tidskr. 1962, 10, (181) Petrykin, V.; Macounova, K.; Shlyakhtin, O.; Krtil, P. Tailoring
1−4. the Selectivity for Electrocatalytic Oxygen Evolution on Ruthenium
(161) Spasojević, M.; Krstajić, N.; Jakšić, M. Structure, Properties and Oxides by Zinc Substitution. Angew. Chem., Int. Ed. 2010, 49, 4813−
Optimization of an Anodic Electrocatalyst: RuO2/TiO2 on Titanium. J. 4815.
Mol. Catal. 1987, 40, 311−326. (182) Petrykin, V.; Macounova, K.; Franc, J.; Shlyakhtin, O.;
(162) Hardee, K. L.; Mitchell, L. K. The Influence of Electrolyte Klementova, M. K.; Mukerjee, S.; Krtil, P. Zn-doped RuO2 Electro-
Parameters on the Percent Oxygen Evolved from a Chlorate Cell. J. catalysts for Selective Oxygen Evolution: Relationship Between Local
Electrochem. Soc. 1989, 136, 3314−3318. Structure and Electrocatalytic Behavior in Chloride Containing Media.
(163) Zhinkin, N. V.; Novikov, E. A.; Fedotova, N. S.; Éberil, V. I.; Chem. Mater. 2011, 23, 200−207.
Busse-Macukas, V. B. Oxygen Evolution and Ruthenium Losses from (183) Hansen, H. A.; Man, I. C.; Studt, F.; Abild-Pedersen, F.; Bligaard,
Active Coatings of Titanium-ruthenium Oxide Anodes in Sodium T.; Rossmeisl, J. Electrochemical Chlorine Evolution at Rutile Oxide
Chloride Solutions at Different Current Densities, Ph and Temper- (110) Surfaces. Phys. Chem. Chem. Phys. 2010, 12, 283−290.
atures. Elektrokhimiya 1989, 25, 1094−1099. (184) Rossmeisl, J.; Qu, Z.-W.; Zhu, H.; Kroes, G.-J.; Nørskov, J. K.
(164) Eberil, V. I.; Dobrov, Y. G.; Novikov, E. A.; Fedotova, N. S. The Electrolysis of Water on Oxide Surfaces. J. Electroanal. Chem. 2007, 607,
Current Efficiency for Chlorate and Current Losses Due to Oxygen 83−89.
Formation as a Function of Anodic Potential in the Electrolysis of (185) Chen, S.; Zheng, Y.; Wang, S.; Chen, X. Ti/RuO2-Sb2O5-SnO2
Chloride-chlorate Solutions at RTO Anodes. Russ. J. Electrochem. 1997, Electrodes for Chlorine Evolution from Seawater. Chem. Eng. J. 2011,
33, 570−572. 172, 47−51.
(165) Trasatti, S.; Lodi, G. In Electrodes of Conductive Metallic Oxides. (186) Wang, S.; Xu, H.; Yao, P.; Chen, X. Ti/RuO2-IrO2SnO2-Sb2O5
Part B; Trasatti, S., Ed.; Elsevier Scientific Publishing Company, 1981; anodes for Cl2 Evolution from Seawater. Electrochemistry 2012, 80, 507−
Chapter 10, pp 521−614. 511.
(166) Boodts, J. F. C.; Trasatti, S. Effect of Composition on the (187) Neodo, S.; Rosestolato, D.; Ferro, S.; De Battisti, A. On the
Electrocatalytic Activity of the Ternary Oxide Ru0:3Ti(0:7−x)SnxO2. J. Electrolysis of Dilute Chloride Solutions: Influence of the Electrode
Electrochem. Soc. 1990, 137, 3784−3789. Material on Faradaic Efficiency for Active Chlorine, Chlorate and
(167) Couper, A.; Brooks, W.; Denton, D. In Modern Chlor-alkali Perchlorate. Electrochim. Acta 2012, 80, 282−291.
Technology; Prout, N., Moorhouse, J., Eds.; Elsevier Applied Science, (188) Xiong, K.; Deng, Z.; Li, L.; Chen, S.; Xia, M.; Zhang, L.; Qi, X.;
1990; Vol. 4, Chapter 8, pp 71−83. Ding, W.; Tan, S.; Wei, Z. Sn and Sb Co-doped RuTi Oxides Supported
(168) Wanngård, J. In Modern Chlor-alkali Technology; Wellington, T. on TiO2 Nanotubes Anode for Selectivity Toward Electrocatalytic
C., Ed.; Elsevier Applied Science, 1992; Vol. 5; Chapter 25, pp 295−306. Chlorine Evolution. J. Appl. Electrochem. 2013, 43, 847−854.
(169) Traini, C.; Meneghini, G. In Modern Chlor-alkali technology; (189) Kuznetsova, E.; Petrykin, V.; Sunde, S.; Krtil, P. Selectivity of
Wellington, T. C., Ed.; Elsevier Applied Science, 1992; Vol. 5; Chapter Nanocrystalline IrO2- based Catalysts in Parallel Chlorine and Oxygen
23, pp 269−280. Evolution. Electrocatalysis 2015, 6, 198−210.
(170) Czarnetzki, L.; Janssen, L. Formation of Hypochlorite, Chlorate (190) Le Luu, T.; Kim, J.; Yoon, J. Physicochemical Properties of RuO2
and Oxygen During NaCl Electrolysis from Alkaline Solutions at an and IrO2 Electrodes Affecting Chlorine Evolutions. J. Ind. Eng. Chem.
RuO2/TiO2 Anode. J. Appl. Electrochem. 1992, 22, 315−324. 2015, 21, 400−404.
(171) Takasu, Y.; Arikawa, T.; Sunohara, S.; Yahikozawa, K. Direct (191) Spasojevic, M.; Ribic-Zelenovic, L.; Spasojevic, P.; Nikolic, B.
Detection of Competitively Evolving Chlorine and Oxygen at Anodes. J. Current Efficiency In The Chlorate Cell Process. J. Serb. Chem. Soc.
Electroanal. Chem. 1993, 361, 279−281. 2014, 79, 677−688.
(172) Nakajima, Y.; Nakamatsu, S.; Shimamune, T.; Matsuda, Y. (192) Zeradjanin, A. R.; Menzel, N.; Schuhmann, W.; Strasser, P. On
Oxide-Coated Titanium Anode for Chlor-alkali Industry. Proceedings of the Faradaic Selectivity and the Role of Surface Inhomogeneity During
the Symposia on Chlor-Alkali and Chlorate Production and New the Chlorine Evolution Reaction at Ternary Ti-Ru-Ir Mixed Metal

3026 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

Oxide Electrocatalysts. Phys. Chem. Chem. Phys. 2014, 16, 13741− Evolution Reaction (OER) on RuO 2 (110). ChemElectroChem 2015, 2,
13747. 707−713.
(193) Abbott, D. F.; Petrykin, V.; Okube, M.; Bastl, Z.; Mukerjee, S.; (212) Kötz, R.; Stucki, S.; Scherson, D.; Kolb, D. In-situ Identification
Krtil, P. Selective Chlorine Evolution Catalysts Based on Mg-doped of RuO4 as the Corrosion Product During Oxygen Evolution on
Nanoparticulate Ruthenium Dioxide. J. Electrochem. Soc. 2015, 162, Ruthenium in Acid Media. J. Electroanal. Chem. Interfacial Electrochem.
H23−H31. 1984, 172, 211−219.
(194) Calle-Vallejo, F.; Koper, M. T. First-principles Computational (213) De Zoubov, N.; Pourbaix, M. In Atlas of Electrochemical
Electrochemistry: Achievements and Challenges. Electrochim. Acta Equilibria In Aqueous Solutions; Pourbaix, M., Ed.; Pergamon Press Ltd.,
2012, 84, 3−11. 1966; Chapter IV, Section 13.1, pp 343−349.
(195) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; (214) Loučka, T. The Potential-pH Diagram for the Ru-H2O-Cl−
Kitchin, J. R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for System at 25°c. J. Appl. Electrochem. 1990, 20, 522−523.
Oxygen Reduction at a Fuel-cell Cathode. J. Phys. Chem. B 2004, 108, (215) Llopis, J.; Vázquez, M. Passivation of Ruthenium in Hydro-
17886−17892. chloric Acid Solution. Electrochim. Acta 1966, 11, 633−640.
(196) Filhol, J.-S.; Neurock, M. Elucidation of the Electrochemical (216) Llopis, J.; Gamboa, J.; Alfayate, J. Radiochemical Study of the
Activation of Water Over Pd by First Principles. Angew. Chem., Int. Ed. Anodic Corrosion of Ruthenium. Electrochim. Acta 1967, 12, 57−65.
2006, 45, 402−406. (217) Gorodetskii, V. V.; Pecherskii, M. M.; Skuratnik, Y. B.;
(197) Otani, M.; Sugino, O. First-Principles Calculations of Charged Dembrovskii, M. A.; Losev, V. V. Anodic Behavior of Ruthenium.
Surfaces and Interfaces: a Plane-wave Nonrepeated Slab Approach. Phys.
Elektrokhimiya 1973, 9, 894−897.
Rev. B: Condens. Matter Mater. Phys. 2006, 73, 115407−1−115407−11.
(218) Pecherskii, M. M.; Gorodetskii, V. V.; Pulina, V. M.; Losev, V. V.
(198) Rossmeisl, J.; Skúlason, E.; Björketun, M. E.; Tripkovic, V.;
Effect of Acidity on the Anodic Behavior of Ruthenium. Elektrokhimiya
Nørskov, J. K. Modeling the Electrified Solid-liquid Interface. Chem.
Phys. Lett. 2008, 466, 68−71. 1976, 12, 1445−1448.
(199) Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Bahn, S.; Hansen, L.; (219) Veselovskaya, I. E.; Khodkevich, S. D.; Malkina, R. I.;
Bollinger, M.; Bengaard, H.; Hammer, B.; Sljivancanin, Z.; Mavrikakis, Yakimenko, L. M. Electrochemical Behavior of Ruthenium Anode in
M.; et al. Universality in Heterogeneous Catalysis. J. Catal. 2002, 209, Chlorine and Oxygen Evolution Processes. Elektrokhimiya 1974, 10,
275−278. 74−77.
(200) Skúlason, E.; Karlberg, G. S.; Rossmeisl, J.; Bligaard, T.; Greeley, (220) Iwakura, C.; Hirao, K.; Tamura, H. Anodic Evolution of Oxygen
J.; Jónsson, H.; Nørskov, J. K. Density Functional Theory Calculations on Ruthenium in Acidic Solutions. Electrochim. Acta 1977, 22, 329−334.
for the Hydrogen Evolution Reaction in an Electrochemical Double (221) Uzbekov, A. A.; Lambrev, V. G.; Yazikov, I. F.; Rodin, N. N.;
Layer on the Pt(111) Electrode. Phys. Chem. Chem. Phys. 2007, 9, 3241− Zabrodskaya, L. M.; Klement’eva, V. S.; Vlodov, Y. M. Corrosion of
3250. Titanium-supported Ruthenium Dioxide Anodes - Nature of Anode
(201) Skúlason, E.; Tripkovic, V.; Björketun, M. E.; Gudmundsdóttir, Consumption and Ruthenium Dissolution Kinetics in Chloride
S.; Karlberg, G.; Rossmeisl, J.; Bligaard, T.; Jónsson, H.; Nørskov, J. K. Solutions. Elektrokhimiya 1978, 14, 1150−1159.
Modeling the Electrochemical Hydrogen Oxidation and Evolution (222) Kokoulina, D. V.; Krasovitskaya, Y. I.; Ivanova, T. V. Oxygen
Reactions on the Basis of Density Functional Theory Calculations. J. Evolution at Ruthenium Oxide Anodes and Its Connection with Anode
Phys. Chem. C 2010, 114, 18182−18197. Destruction. Elektrokhimiya 1978, 14, 470−474.
(202) Greeley, J.; Stephens, I. E. L.; Bondarenko, A. S.; Johansson, T. (223) Bondar, R. U.; Kalinovskii, E. A. Electrochemical Stability of
P.; Hansen, H. A.; Jaramillo, T. F.; Rossmeisl, J.; Chorkendorff, I.; Titanium-ruthenium Oxide Anodes. Elektrokhimiya 1978, 14, 730−733.
Nørskov, J. K. Alloys of Platinum and Early Transition Metals as Oxygen (224) Gorodetskii, V. V.; Pecherskii, M. M.; Yanke, V. B.; Shub, D. M.;
Reduction Electrocatalysts. Nat. Chem. 2009, 1, 552−556. Losev, V. V. Kinetics of Dissolution of Ruthenium-titanium Oxide
(203) Siahrostami, S.; Verdaguer-Casadevall, A.; Karamad, M.; Deiana, Anodes in the Electrolysis of Chloride Solutions. Elektrokhimiya 1979,
D.; Malacrida, P.; Wickman, B.; Escudero-Escribano, M.; Paoli, E. A.; 15, 559−562.
Frydendal, R.; Hansen, T. W.; et al. Enabling Direct H2O2 Production (225) Barral, G.; Guitton, J.; Montella, C.; Vergara, F. Étude des
Through Rational Electrocatalyst Design. Nat. Mater. 2013, 12, 1137− anodes à couche d’oxyde de ruthénium sur substrat de titane III:
1143. Influence de quelques paramétres de préparation des couches d’oxyde et
(204) Over, H.; Kim, Y. D.; Seitsonen, A. P.; Wendt, S.; Lundgren, E.; des conditions d’électrolyse sur l’usure lors d’un dégagement d’oxygéne.
Schmid, M.; Varga, P.; Morgante, A.; Ertl, G. Atomic-Scale structure and Surf. Technol. 1980, 10, 25−45.
Catalytic reactivity of the RuO2 (110) surface. Science 2000, 287, 1474− (226) Tilak, B. V.; Birss, V. I.; Wang, J.; Chen, C.-P.; Rangarajan, S. K.
1476. Deactivation of Thermally Formed Ru/Ti Oxide Electrodes: An AC
(205) Mars, P.; van Krevelen, D. Oxidations Carried Out by Means of Impedance Characterization Study. J. Electrochem. Soc. 2001, 148, D112.
Vanadium Oxide Catalysts. Chem. Eng. Sci. 1954, 3, 41−59. (227) Eberil’, V. I.; Novikov, E. A.; Mazanko, A. F. Reasons for the DSA
(206) Fang, Y.-H.; Liu, Z.-P. Mechanism and Tafel Lines of Electro-
Passivation during Chlorate Electrolysis and the Means for Extending
oxidation of Water to Oxygen on RuO2(110). J. Am. Chem. Soc. 2010,
the Anode Service Life. Russ. J. Electrochem. 2001, 37, 1054−1058.
132, 18214−18222.
(228) Burke, L. D.; Lyons, M. E.; McCarthy, M. Oxygen Evolution and
(207) Exner, K. S.; Anton, J.; Jacob, T.; Over, H. Chlorine Evolution
Corrosion at Ruthenium Dioxide-Based Anodes. Proceedings 4th World
Reaction on RuO2(110): Ab initio Atomistic Thermodynamics Study -
Pourbaix Diagrams. Electrochim. Acta 2014, 120, 460−466. Energy Conference, Pasadena, CA, 1982; IEEE: Piscataway, NJ, pp 267−
(208) Wang, H.; Schneider, W. F.; Schmidt, D. Intermediates and 278.
Spectators in O2 Dissociation at the RuO2 (110) Surface. J. Phys. Chem. (229) Uzbekov, A. A.; Klement’eva, V. S. Radiochemical Investigation
C 2009, 113, 15266−15273. of the Selectivity of Dissolution of Components of Ruthenium-titanium
(209) Siahrostami, S.; Vojvodic, A. Influence of Adsorbed Water on the Oxide Anodes (ORTA) in Chloride Solutions. Elektrokhimiya 1985, 21,
Oxygen Evolution Reaction on Oxides. J. Phys. Chem. C 2015, 119, 758−763.
1032−1037. (230) Schmets, J.; Van Muylder, J.; Pourbaix, M. In Atlas of
(210) Exner, K. S.; Anton, J.; Jacob, T.; Over, H. Microscopic Insights Electrochemical Equilibria In Aqueous Solutions; Pourbaix, M., Ed.;
into the Chlorine Evolution Reaction on RuO2(110): a Mechanistic Ab Pergamon Press Ltd, 1966; Chapter IV, pp 213−222.
Initio Atomistic Thermodynamics Study. Electrocatalysis 2015, 6, 163− (231) Uzbekov, A. A.; Klement’eva, V. S.; Kubasov, V. L.; Lambrev, V.
172. G. Corrosion Behavior of Ruthenium Oxide-titanium Anodes During
(211) Exner, K. S.; Anton, J.; Jacob, T.; Over, H. Ligand Effects and Production of Sodium Chlorate by Electrolysis of Chloride-chlorate
Their Impact on Electrocatalytic Processes Exemplified with the Oxygen Solutions. J. Appl. Chem. USSR 1985, 58, 686−688.

3027 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028
Chemical Reviews Review

(232) Kötz, R.; Stucki, S. Stabilization of RuO2 by IrO2 for Anodic Spectroscopic Surface Composition Studies of Titanium-ruthenium
Oxygen Evolution in Acid Media. Electrochim. Acta 1986, 31, 1311− Oxide Anodes (ORTA) Under the Conditions of Chlorine Electrolysis.
1316. Elektrokhimiya 1981, 17, 79−83.
(233) Kolotyrkin, Y. M. Relationship Between Kinetic and Corrosion (253) Takasu, Y.; Sugimoto, W.; Nishiki, Y.; Nakamatsu, S. Structural
Characteristics of Anodic Materials. Indian J. Technol. 1986, 24, 381− analyses of RuO2−TiO2/Ti and IrO2−RuO2−TiO2/Ti anodes used in
387. industrial chlor-alkali membrane processes. J. Appl. Electrochem. 2010,
(234) Lyons, M. E. G.; Burke, L. D. Mechanism of Oxygen Reactions at 40, 1789−1795.
Porous Oxide Electrodes. Part 1. - Oxygen Evolution at RuO2 and (254) Loučka, T. The Reason for the Loss of Activity of Titanium
RuxSn1−xO2 Electrodes in Alkaline Solution Under Vigorous Electrolysis Anodes Coated with a Layer of RuO2 and TiO2. J. Appl. Electrochem.
Conditions. J. Chem. Soc., Faraday Trans. 1 1987, 83, 299−321. 1977, 7, 211−214.
(235) Manli, S.; Yanxi, C. In Modern Chlor-alkali Technology; Prout, N.,
Moorhouse, J., Eds.; Springer: The Netherlands, 1990; Vol. 4; Chapter
14, pp 149−157.
(236) Hardee, K. L.; Kus, R. A. In Modern Chlor-alkali Technology;
Sealey, S., Ed.; SCI, 1998; Vol. 7, Chapter 7, pp 43−54.
(237) Bommaraju, T. V.; Chen, C.-P.; Birss, V. I. In Modern Chlor-alkali
Technology; Moorhouse, J., Ed.; Wiley-Blackwell, 2001; Vol. 8, Chapter
5, pp 57−81.
(238) Nylén, L.; Cornell, A. Critical Anode Potential in the Chlorate
Process. J. Electrochem. Soc. 2006, 153, D14−D20.
(239) Gajić-Krstajić, L.; Trišović, T.; Krstajić, N. Spectrophotometric
Study of the Anodic Corrosion of Ti/RuO2 Electrode in Acid Sulfate
Solution. Corros. Sci. 2004, 46, 65−74.
(240) Cherevko, S.; Zeradjanin, A. R.; Topalov, A. A.; Kulyk, N.;
Katsounaros, I.; Mayrhofer, K. J. J. Dissolution of Noble Metals during
Oxygen Evolution in Acidic Media. ChemCatChem 2014, 6, 2219−2223.
(241) Cherevko, S.; Reier, T.; Zeradjanin, A. R.; Pawolek, Z.; Strasser,
P.; Mayrhofer, K. J. Stability of Nanostructured Iridium Oxide
Electrocatalysts During Oxygen Evolution Reaction in Acidic Environ-
ment. Electrochem. Commun. 2014, 48, 81−85.
(242) Zeradjanin, A. R.; Topalov, A. A.; Van Overmeere, Q.; Cherevko,
S.; Chen, X.; Ventosa, E.; Schuhmann, W.; Mayrhofer, K. J. J. Rational
Design of the Electrode Morphology for Oxygen Evolution - Enhancing
the Performance for Catalytic Water Oxidation. RSC Adv. 2014, 4,
9579−9587.
(243) Reier, T.; Pawolek, Z.; Cherevko, S.; Bruns, M.; Jones, T.;
Teschner, D.; Selve, S.; Bergmann, A.; Nong, H. N.; Schlögl, R. e.; et al.
Molecular Insight in Structure and Activity of Highly Efficient, Low-Ir
Ir-Ni Oxide Catalysts for Electrochemical Water Splitting (OER). J. Am.
Chem. Soc. 2015, 137, 13031−13040.
(244) Topalov, A. A.; Katsounaros, I.; Auinger, M.; Cherevko, S.;
Meier, J. C.; Klemm, S. O.; Mayrhofer, K. J. J. Dissolution of Platinum:
Limits for the Deployment of Electrochemical Energy Conversion?
Angew. Chem., Int. Ed. 2012, 51, 12613−12615.
(245) Frydendal, R.; Paoli, E. A.; Knudsen, B. P.; Wickman, B.;
Malacrida, P.; Stephens, I. E. L.; Chorkendorff, I. Benchmarking the
Stability of Oxygen Evolution Reaction Catalysts: The Importance of
Monitoring Mass Losses. ChemElectroChem 2014, 1, 2075−2081.
(246) Frydendal, R.; Paoli, E. A.; Chorkendorff, I.; Rossmeisl, J.;
Stephens, I. E. L. Toward an Active and Stable Catalyst for Oxygen
Evolution in Acidic Media: Ti-Stabilized MnO2. Adv. Energy Mater.
2015, 5, 1−9.
(247) Gijsbers, H. F. M.; Janssen, L. J. J. Distribution of mass transfer
over a 0.5-m-tall hydrogen-evolving electrode. J. Appl. Electrochem. 1989,
19, 637−648.
(248) Vogt, H.; Stephan, K. Local microprocesses at gas-evolving
electrodes and their influence on mass transfer. Electrochim. Acta 2015,
155, 348−356.
(249) Comninellis, C.; Vercesi, G. P. Problems in DSA® Coating
Deposition by Thermal Decomposition. J. Appl. Electrochem. 1991, 21,
136−142.
(250) Jones, G.; Bligaard, T.; Abild-Pedersen, F.; Nørskov, J. K. Using
Scaling Relations to Understand Trends in the Catalytic Activity of
Transition Metals. J. Phys.: Condens. Matter 2008, 30, 064239−1−
064239−6.
(251) Rossmeisl, J.; Logadottir, A.; Nørskov, J. K. Electrolysis of Water
on (Oxidized) Metal Surfaces. Chem. Phys. 2005, 319, 178−184.
(252) Gorodetskii, V. V.; Zorin, P. N.; Pecherskii, M. M.; Busse-
Machukas, V. B.; Kubasov, V. L.; Tomashpol’skii, Y. Y. Auger

3028 DOI: 10.1021/acs.chemrev.5b00389


Chem. Rev. 2016, 116, 2982−3028

Potrebbero piacerti anche