Thanks to visit codestin.com
Credit goes to www.scribd.com

100% found this document useful (3 votes)
897 views180 pages

Electrolysis Processes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (3 votes)
897 views180 pages

Electrolysis Processes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 180

Electrolysis

Processes
Edited by
Tanja Vidakovic-Koch
Printed Edition of the Special Issue Published in Processes

www.mdpi.com/journal/processes
Electrolysis Processes
Electrolysis Processes

Special Issue Editor


Tanja Vidakovic-Koch

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade • Manchester • Tokyo • Cluj • Tianjin
Special Issue Editor
Tanja Vidakovic-Koch
Max Planck Institute for
Dynamics of Complex Technical
Systems Sandtorstraße
Germany

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal
Processes (ISSN 2227-9717) (available at: https://www.mdpi.com/journal/processes/special issues/
electrolysis processes).

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.

ISBN 978-3-03936-386-5 (Pbk)


ISBN 978-3-03936-387-2 (PDF)


c 2020 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

About the Special Issue Editor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Tanja Vidaković-Koch
Editorial on Special Issue Electrolysis Processes
Reprinted from: Processes 2020, 8, 578, doi:10.3390/pr8050578 . . . . . . . . . . . . . . . . . . . . . 1

Jörn Brauns and Thomas Turek


Alkaline Water Electrolysis Powered by Renewable Energy: A Review
Reprinted from: Processes 2020, 8, 248, doi:10.3390/pr8020248 . . . . . . . . . . . . . . . . . . . . . 5

Nicole Vorhauer, Haashir Altaf, Evangelos Tsotsas and Tanja Vidakovic-Koch


Pore Network Simulation of Gas-Liquid Distribution in Porous Transport Layers
Reprinted from: Processes 2019, 7, 558, doi:10.3390/pr7090558 . . . . . . . . . . . . . . . . . . . . . 29

Haashir Altaf, Nicole Vorhauer, Evangelos Tsotsas and Tanja Vidaković-Koch


Steady-State Water Drainage by Oxygen in Anodic Porous Transport Layer of Electrolyzers:
A 2D Pore Network Study
Reprinted from: Processes 2020, 8, 362, doi:10.3390/pr8030362 . . . . . . . . . . . . . . . . . . . . . 53

Narjes Nabipour, Amir Mosavi, Alireza Baghban, Shahaboddin Shamshirband and


Imre Felde
Extreme Learning Machine-Based Model for Solubility Estimation of Hydrocarbon Gases
in Electrolyte Solutions
Reprinted from: Processes 2020, 8, 92, doi:10.3390/pr8010092 . . . . . . . . . . . . . . . . . . . . . 71

Jiabing Xia, Gerhart Eigenberger, Heinrich Strathmann and Ulrich Nieken


Acid-Base Flow Battery, Based on Reverse Electrodialysis with Bi-Polar Membranes:
Stack Experiments
Reprinted from: Processes 2020, 8, 99, doi:10.3390/pr8010099 . . . . . . . . . . . . . . . . . . . . . 83

Deepa Guragain, Camila Zequine, Ram K Gupta and Sanjay R Mishra


Facile Synthesis of Bio-Template Tubular MCo2 O4 (M = Cr, Mn, Ni) Microstructure and Its Electrochemical
Performance in Aqueous Electrolyte
Reprinted from: Processes 2020, 8, 343, doi:10.3390/pr8030343 . . . . . . . . . . . . . . . . . . . . . 97

Thorben Muddemann, Dennis Haupt, Bolong Jiang, Michael Sievers and Ulrich Kunz
Investigation and Improvement of Scalable Oxygen Reducing Cathodes for Microbial Fuel Cells
by Spray Coating
Reprinted from: Processes 2020, 8, 11, doi:10.3390/pr8010011 . . . . . . . . . . . . . . . . . . . . . 115

Kangdong Xu, Jianghua Peng, Pan Chen, Wankai Gu, Yunbai Luo and Ping Yu
Preparation and Characterization of Porous Ti/SnO2 –Sb2 O3 /PbO2 Electrodes for the Removal
of Chloride Ions in Water
Reprinted from: Processes 2019, 7, 762, doi:10.3390/pr7100762 . . . . . . . . . . . . . . . . . . . . . 131

Sheng Liu, Lin Gui, Ruichao Peng and Ping Yu


A Novel Porous Ni, Ce-Doped PbO2 Electrode for Efficient Treatment of Chloride Ion
in Wastewater
Reprinted from: Processes 2020, 8, 466, doi:10.3390/pr8040466 . . . . . . . . . . . . . . . . . . . . . 145

v
Yong Liu, Chao Zhang, Songsong Li, Chunsheng Guo and Zhiyuan Wei
Experimental Study of Micro Electrochemical Discharge Machining of Ultra-Clear Glass with
a Rotating Helical Tool
Reprinted from: Processes 2019, 7, 195, doi:10.3390/pr7040195 . . . . . . . . . . . . . . . . . . . . . 155

vi
About the Special Issue Editor
Tanja Vidakovic-Koch studied chemical engineering at University of Belgrade, Serbia,
where she also obtained her Magister of Science degree in electrochemistry. She pursued further
doctoral studies in process and systems engineering at Max Planck Institute for Dynamics of
Complex Technical Systems (MPI-DCTS), Magdeburg, Germany, where she obtained her Dr.-Ing.
degree as well as lecturer (Habilitation) qualification for Electrochemical Process Engineering (“venia
legendi”) from Otto von Guericke University (OvGU) in Magdeburg, Germany. Currently she is
Head of the Electrochemical Energy Conversion group at the MPI-DCTS, Magdeburg, Germany,
and she lectures at OvGU. Her research focuses on the model-based analysis of linear and nonlinear
dynamics of electrochemical energy conversion in polymer electrolyte membrane (PEM) fuel cells,
PEM electrolyzers, and enzymatic fuel cells.

vii
processes
Editorial
Editorial on Special Issue Electrolysis Processes
Tanja Vidaković-Koch
Electrochemical Energy Conversion, Max Planck Institute for Dynamics of Complex Technical Systems,
D-39106 Magdeburg, Germany; [email protected]

Received: 7 May 2020; Accepted: 7 May 2020; Published: 14 May 2020

Renewable energies such as solar, hydro or wind power are in principal abundant but subjected
to strong fluctuations. Therefore, development of new technologies for storage of these renewable
energies is of special interest. Electrochemical technologies are ideal candidates for the use of excess
current, and consequently an increased electrification of chemical processes is expected. In this respect,
there are different pathways to utilize excess current electrochemically. Intermediate energy storages in
(a) chemical energy carriers like hydrogen via water electrolysis or (b) electrochemical energy storage
devices like batteries are perhaps most accepted and discussed solutions. Additionally, excess current
can be used with the main goal not to be stored for later use, but to solve some environmental issue
or for construction purposes. Possible applications are waste water treatment and electromachining.
The article collection in this special issue consists of one review paper and nine original research papers
and discusses these topics in more detail. As a Guest Editor of this special issue, I thank all authors for
their contributions and wish all readers interesting insights and new inspirations for their works.

Electrolysis Processes for Intermediate Energy Storage in Chemical Energy Carriers Like
Hydrogen–Water Electrolysis
There are three main technologies for water electrolysis: alkaline water electrolysis (AEL),
proton exchange membrane (or polymer electrolyte membrane) electrolysis (PEMEL), and solid
oxide electrolysis (SOEL), with two of them (AEL and PEMEL) at high technical readiness level.
Despite these facts and intensive discussions on water electrolysis as a key technology for generation
of pure hydrogen using renewable electricity, currently less than 4% of hydrogen originates from
electrolysis, with the main part originating not from water but from from chlor-alkali electrolysis
where hydrogen is a by-product of chlorine production [1]. Broad introduction of cost competitive and
preferably zero carbon routes for hydrogen production is urgently required. In the review paper of
Brauns and Turek [1], some of the main challenges hindering broader penetration of water electrolysis
are discussed with a focus on AEL. As the authors write, AEL is a key technology for large-scale
hydrogen production powered by renewable energy. However, conventional electrolyzers are designed
for operation at fixed process conditions, therefore, the implementation of fluctuating and highly
intermittent renewable energy is challenging. Their system analysis enabled important insights and
a roadmap for more energy efficient systems. According to these authors, in order to be competitive
with the conventional path based on fossil energy sources, each component of a hydrogen energy
system needs to be optimized to increase the operation time and system efficiency. They stress that by
combining AEL with hydrogen storage tanks and fuel cells, power grid stabilization can be achieved.
As a consequence, the conventional spinning reserve can be reduced, which additionally lowers the
carbon dioxide emissions.
Water electrolysis was also in the focus of Vorhauer et al. [2] and Altaf et al. [3]. One of the
reasons for lowering energy efficiency of water electrolysis at high current densities is mass transport
limitation in porous transport layers (PTL) of water electrolyzers. This issue is intrinsic for both
low temperature water technologies (AEL and PEMEL) and is likely caused by the counter current
transport of oxygen gas produced at the anode to the educt (water or OH- ions). In two publications by
Vorhauer et al. [2] and Altaf et al. [3], this issue was studied with the help of porous network theory and

Processes 2020, 8, 578; doi:10.3390/pr8050578 1 www.mdpi.com/journal/processes


Processes 2020, 8, 578

with special emphasis on PEMEL. Pore network models (PNM) are powerful tools to simulate invasion
and transport processes in different porous media with applications across different disciplines like
geology, chemical engineering as well as electrochemical engineering (e.g., fuel cell applications).
In their pioneering contribution, Vorhauer et al. [2] described a first application of a PNM of drainage
for the prediction of the oxygen and water invasion process inside the anodic PTL at high current
densities. According to the authors, in the simulation with narrow pore size distribution, the volumetric
ratio of the liquid transporting clusters connected between the catalyst layer and the water supply
channel is only around 3% of the total liquid volume contained inside the pore network at the moment
when the water supply route through the pore network is interrupted; whereas around 40% of the
volume is occupied by the continuous gas phase. The majority of liquid clusters are disconnected
from the water supply routes through the pore network if liquid films along the walls of the porous
transport layer are disregarded. Moreover, these clusters hinder the countercurrent oxygen transport.
They also based on these sketches a new route for the extraction of transport parameters from Monte
Carlo simulations, incorporating pore scale flow computations and Darcy flow.
In the publication by Altaf et al. [3], results from Monte Carlo pore network simulations are
shown and compared qualitatively to microfluidic experiments from literature. The literature-based
experimental invasion patterns of different types of PTLs, such as felt, foam or sintered, have shown
good agreement with pore network simulations. Additionally, the impact of pore size distribution on
the phase patterns of oxygen and water inside the pore network was studied. These very promising
results further supported pore network modeling as a valuable tool for gaining new insights in the
transport processes in porous layers during water electrolysis.
Reliable models of gas-liquid equilibria in aqueous electrolytes are of significant importance
for proper description of many electrochemical processes including gases as products or reactants
(major examples are water, brine, CO2 electrolysis, but also reactions in polymer electrolyte fuel cells).
Nabipour et al. [4] proposed a novel solubility estimation tool for gases in aqueous electrolyte solutions
based on an extreme learning machine (ELM) algorithm. Although the presented study cases are not
directly relevant for water electrolysis, the here developed novel methodology has a potential to be
transferred to more relevant examples for electrochemical applications.

Electrolysis Processes for Intermediate Energy Storage in Electrochemical Energy Storage Devices
Different types of electrochemical energy storage devices are discussed in the framework of
intermediate energy storage from renewables. Common examples are rechargeable batteries like
Li-ion or redox flow batteries. In this special issue, two less discussed options are described,
a so-called acid-base flow battery [5] and supercapacitors [6]. An acid-base flow battery is proposed
by Xia et al. [5]. This very interesting solution is based on reverse electrodialalysis with bi-polar
membranes. During charging, the system operates in electrolysis mode; hydrogen and oxygen
evolution reactions take place at the cathode and anode, respectively; and acid and base solutions are
regenerated in corresponding compartments. Alike normal water electrolysis, the two half-reactions
take place at different pH values (acidic and alkaline conditions). During discharge, neutralization of
acid and base produces electricity in the process of reverse electrodialysis with bipolar membranes
in an apparently fairly overlooked flow battery concept. The authors demonstrated this concept
with stack experiments, consisting of 5 to 20 repeating cell units at lab scale. The first results are
very promising and the studied system shows already energy density in the similar range of redox
flow batteries. The challenges and measures to further increase energy efficiency of this new type of
acid-base flow battery are discussed.
Due to the high power fluctuations that are inherent to renewable energy sources, the dynamics
of the storage media is of great importance when designing storage concepts for renewable
energy. Electrochemical storage systems showing even better dynamics than batteries are so-called
supercapacitors. These devices can be quickly charged and discharged, but have lower energy density
than batteries. Still, due to their favorable dynamics, they can be a valuable addition to batteries

2
Processes 2020, 8, 578

in the framework of intermediate energy storage from renewables. The storage capacity of these
devices depends largely on employed materials. In this special issue, development of novel electrode
material for supercapacitor application based on pseudocapacitance is discussed. Guragain et al. [6]
developed a large-surface-area MCO2 O4 material in which a tubular microstructure leads to a noticeable
pseudocapacitive property with the excellent specific capacitance value exceeding 407.2 F/g at 2 mV/s
scan rate. In addition, a Coulombic efficiency ~100% and excellent cycling stability with 100%
capacitance retention was noted even after 5000 cycles. These tubular MCO2 O4 microstructures display
peak power density exceeding 7000 W/kg. Based on these authors, the superior performance of the
tubular MCO2 O4 microstructure electrode is attributed to their high surface area, adequate pore
volume distribution, and active carbon matrix, which allows effective redox reaction and diffusion of
hydrated ions.

Electrolysis Processes Aiming to Solve Environmental Issues or for Construction Purposes


In addition to clean energy, the requirement of clean water is of the upmost importance for
prosperity and healthiness of the world population. With respect to water, one issue is water scarcity,
causing 1.2 billion people to suffer from a lack of water [7]. Another issue is industry or domestic-based
water pollution [7–9]. Due to the widespread use of chlorinated compounds such as HCl, NaCl,
and MgCl2 in the industry, the content of chloride ion in wastewater is increasing [8,9]. If it is discharged
directly to the environment, diverse environmental issues will be created with consequences for water
resources, soil and human health. At present the most widely used method for Cl− waste water
treatment is chemical precipitation. Electrochemical oxidation processes have a high potential for
efficient removal of even trace amounts of Cl− due to high degradation efficiency, no secondary pollution,
modularity, flexibility, and use of renewable electricity, which can contribute to stabilization of the
energy grid. The degradation efficiency and degradation products of the electrochemical oxidation
process change with the anode material, therefore, two articles by Xu et al. [8] and Liu et al. [9] discuss
the development of new anode materials for electrochemical Cl− removal. Xu et al. [8] studied porous
Ti/SnO2 -Sb2 O3 /PbO2 electrodes for electrocatalytic oxidation of chloride ions by varying different
parameters. The results have shown that Ti/SnO2 -Sb2 O3 /PbO2 electrodes with 150 μm pore size had
the best removal effect on chloride ions with removal ratios amounting up to 98.5% when the initial
concentration of chloride ion was 10 g L−1 . The advanced electrode structure minimized oxygen
evolution as a side reaction, which further increased the removal effect of chloride ions. In the further
publication, Liu et al. [9] studied different material combinations. The porous Ti/Sb–SnO2 /Ni–Ce–PbO2
electrode was prepared by using a porous Ti plate as a substrate, an Sb-doped SnO2 as an intermediate,
and a PbO2 doped with Ni and Ce as an active layer. The authors studied also the mechanism of
electrochemical dechlorinating. They found out that the increase of OH− inhibits the degradation
of Cl− .
In addition to inorganic catalysts, oxidation of mainly organic pollutants in the waste water
treatment can be performed with the help of biological catalysts, for example microorganisms.
These systems can operate either as microbial fuel cells (MFC) or in electrolysis mode. A contribution
by Muddemann et al. [7] discuss current challenges in the scale up of these systems, with special
emphasis on oxygen reducing cathode. The authors demonstrated a strong increase in the MFC
performance and long term stability upon improving catalyst coating quality.
Finally Liu et al. [10] describe electrochemical discharge machining (ECDM) as an effective way to
fabricate micro structures in non-conductive materials, such as quartz glass and ceramics. This has
significant importance for the development of micro electromechanical systems (MEMS), such as micro
reactors and micro medical devices, which often consist of the micro structures of nonconductive
materials, such as glass, ceramics and silicon nitride and which are difficult to fabricate using the
traditional machining methods.

Funding: This research received no external funding.

3
Processes 2020, 8, 578

Conflicts of Interest: The author declares no conflict of interest.

References
1. Brauns, J.; Turek, T. Alkaline water electrolysis powered by renewable energy: A review. Processes 2020,
8, 248. [CrossRef]
2. Vorhauer, N.; Altaf, H.; Tsotsas, E.; Vidakovic-Koch, T. Pore network simulation of gas-liquid distribution in
porous transport layers. Processes 2019, 7, 558. [CrossRef]
3. Altaf, H.; Vorhauer, N.; Tsotsas, E.; Vidaković-Koch, T. Steady-state water drainage by oxygen in anodic
porous transport layer of electrolyzers: A 2D pore network study. Processes 2020, 8, 362. [CrossRef]
4. Nabipour, N.; Mosavi, A.; Baghban, A.; Shamshirband, S.; Felde, I. Extreme learning machine-based model
for solubility estimation of hydrocarbon gases in electrolyte solutions. Processes 2020, 8, 92. [CrossRef]
5. Xia, J.; Eigenberger, G.; Strathmann, H.; Nieken, U. Acid-base flow battery, based on reverse electrodialysis
with Bi-polar membranes: Stack experiments. Processes 2020, 8, 99. [CrossRef]
6. Guragain, D.; Zequine, C.; Gupta, R.; Mishra, S. Facile synthesis of bio-template tubular MCo2 O4 (M = Cr,
Mn, Ni) microstructure and its electrochemical performance in aqueous electrolyte. Processes 2020, 8, 343.
[CrossRef]
7. Muddemann, T.; Haupt, D.; Jiang, B.; Sievers, M.; Kunz, U. Investigation and improvement of scalable
oxygen reducing cathodes for microbial fuel cells by spray coating. Processes 2020, 8, 11. [CrossRef]
8. Xu, K.; Peng, J.; Chen, P.; Gu, W.; Luo, Y.; Yu, P. Preparation and characterization of porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes for the removal of chloride ions in water. Processes 2019, 7, 762. [CrossRef]
9. Liu, S.; Gui, L.; Peng, R.; Yu, P. A novel porous Ni, Ce-doped PbO2 electrode for efficient treatment of chloride
ion in wastewater. Processes 2020, 8, 466. [CrossRef]
10. Liu, Y.; Zhang, C.; Li, S.; Guo, C.; Wei, Z. Experimental study of micro electrochemical discharge machining
of ultra-clear glass with a rotating helical tool. Processes 2019, 7, 195. [CrossRef]

© 2020 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

4
processes
Review
Alkaline Water Electrolysis Powered by Renewable
Energy: A Review
Jörn Brauns * and Thomas Turek
Institute of Chemical and Electrochemical Process Engineering, Clausthal University of Technology,
Leibnizstr. 17, 38678 Clausthal-Zellerfeld, Germany; [email protected]
* Correspondence: [email protected]; Tel.: +49-5323-72-2473

Received: 23 January 2020; Accepted: 18 February 2020; Published: 21 February 2020

Abstract: Alkaline water electrolysis is a key technology for large-scale hydrogen production powered
by renewable energy. As conventional electrolyzers are designed for operation at fixed process
conditions, the implementation of fluctuating and highly intermittent renewable energy is challenging.
This contribution shows the recent state of system descriptions for alkaline water electrolysis and
renewable energies, such as solar and wind power. Each component of a hydrogen energy system
needs to be optimized to increase the operation time and system efficiency. Only in this way
can hydrogen produced by electrolysis processes be competitive with the conventional path based
on fossil energy sources. Conventional alkaline water electrolyzers show a limited part-load range
due to an increased gas impurity at low power availability. As explosive mixtures of hydrogen and
oxygen must be prevented, a safety shutdown is performed when reaching specific gas contamination.
Furthermore, the cell voltage should be optimized to maintain a high efficiency. While photovoltaic
panels can be directly coupled to alkaline water electrolyzers, wind turbines require suitable
converters with additional losses. By combining alkaline water electrolysis with hydrogen storage
tanks and fuel cells, power grid stabilization can be performed. As a consequence, the conventional
spinning reserve can be reduced, which additionally lowers the carbon dioxide emissions.

Keywords: alkaline water electrolysis; hydrogen; renewable energy; sustainable; dynamic;


fluctuations; wind; solar; photovoltaic; limitations

1. Introduction
Hydrogen is considered a promising energy carrier for a sustainable future when it is produced
by utilizing renewable energy [1]. Today, less than 4% of hydrogen production is based on electrolysis
processes, of which the main part is hydrogen as a by-product of chlorine production. Hence, the major
share of the needed hydrogen depends on the fossil path through the steam reforming of natural
gas [2]. This situation is caused by the higher production costs of electrolysis processes compared to
the conventional fossil sources, due to high electricity costs and interfering laws [3]. To reduce CO2
emissions and to become independent of fossil energy carriers, the share of hydrogen produced using
renewable power sources needs to be increased significantly in the next few decades. Therefore, water
electrolysis is a key technology for splitting water into hydrogen and oxygen by using renewable
energy. After drying and removing oxygen impurities, the hydrogen purity is higher than 99.9%, and
the hydrogen can be directly used in the following processes or in the transport sector [4]. Solar and
wind energy are the preferred renewable power sources for hydrogen production, as their distribution
is the most widespread [5,6]. Hydropower, biomass, and geothermal energy are alternatives, and are
often utilized for the base load [7]. The main problem with using renewable energy is the unevenly
distributed and intermittent local availability [6]. With a higher share of renewable energy from wind
turbines or solar photovoltaic panels and fair CO2 emission costs, hydrogen production by water

Processes 2020, 8, 248; doi:10.3390/pr8020248 5 www.mdpi.com/journal/processes


Processes 2020, 8, 248

electrolysis will become more attractive. The combination of water electrolysis with renewable energy
is particularly advantageous, as excess electrical energy can be chemically stored in hydrogen to
balance the discrepancy between energy demand and production [6]. For large-scale applications,
the hydrogen can be stored in salt caverns, storage tanks, or the gas grid [8–12]. Smaller hydrogen
quantities can also be stored in metal hydrides [13,14].
For water electrolysis, there are three technologies available: Alkaline water electrolysis (AEL),
proton exchange membrane (or polymer electrolyte membrane) electrolysis (PEMEL), and solid
oxide electrolysis (SOEL) [15–18]. While the low-temperature technologies, AEL and PEMEL, both
provide high technology readiness levels, the high-temperature SOEL technology is still in the
development stage [19]. Alkaline water electrolysis uses concentrated lye as an electrolyte and requires
a gas-impermeable separator to prevent the product gases from mixing. The electrodes consist of
non-noble metals like nickel with an electrocatalytic coating. PEMEL uses a humidified polymer
membrane as the electrolyte and noble metals like platinum and iridium oxide as the electrocatalysts.
Both technologies are operated at temperatures from 50 to 80 ◦C and allow operation pressures of
up to 30 bar. The nominal stack efficiency of both technologies is around 70% [18,20]. SOEL is also
known as high-temperature (HTEL) or steam electrolysis, as gaseous water is converted into hydrogen
and oxygen at temperatures between 700 and 900 ◦C. Theoretically, stack efficiencies near 100%
are possible due to positive thermodynamic effects on power consumption at higher temperatures.
However, the increased thermal demand requires a suitable waste heat source from the chemical,
metallurgical, or thermal power generation industry for economical operation. Moreover, the corrosive
environment demands further material development [6,20,21]. As a consequence, SOEL provides only
small stack capacities below 10 kW, compared to 6 MW for AEL and 2 MW for PEMEL [20]. Hence,
the investment costs and the lifetime determine whether AEL or PEMEL is the most favorable system
design for a large-scale application. Today, the investment costs for AEL are from 800 to 1500 € kW−1
and for PEMEL from 1400 to 2100 € kW−1 . Furthermore, the lifetime of alkaline water electrolyzers is
higher and the annual maintenance costs are lower compared to a PEMEL system [15,20,22,23]. Often,
PEMEL systems are preferred for dynamic operation due to the short start-up time and a broad load
flexibility range. The shortcomings of AEL are gradually being overcome by further development [24].
Therefore, this review focuses on alkaline water electrolysis powered by renewable energy. To ensure
safety and high efficiency, alkaline water electrolyzers must be optimized for dynamic operation.
Hence, the process needs to be analyzed for how the dynamics will affect the system performance
and what aspects should be considered when fluctuating renewable energy is used instead of a
constant load [25]. Thus, this contribution shows model descriptions for alkaline water electrolysis,
photovoltaic panels, and wind turbines to identify the limitations when combining all components into
a hydrogen energy system. Furthermore, theoretical models can help to solve the existing problems
using intelligent system design and suitable operation strategies.
This study mainly contains literature that was obtained with the keywords alkaline electrolyzer
(or electrolyser or electrolysis) in combination with one of the following words: Renewable, sustainable,
green, dynamic, fluctuation, intermittent, solar, photovoltaic, wind, and power to gas. For a
broader overview, additional literature is also included. Figure 1 shows the number of annual
publications that are listed in the Web of Science Database for the given keywords from 1990
to 2019. Additionally, the keyword alkaline is replaced by other water electrolysis technologies
to show the share of technology-specific publications [26]. Around 2010, the number of annual
publications started to increase persistently due to the discussion about the energy turnaround,
especially in Germany and other European countries [9,27]. Furthermore, the topic is often discussed
technology-independently, as the number of technology-specific publications is small compared to
publications with unspecified water electrolysis technologies. While the low-temperature technologies,
AEL and PEMEL, show an equal share of technology-specific publications, the high-temperature
technology SOEL is mentioned less. This distribution reflects the recent considerations of which

6
Processes 2020, 8, 248

technology may be favored for sustainable hydrogen production. Particularly, alkaline water
electrolysis is considered as the most reliable method for large-scale hydrogen production [5,21].

50 AEL

number of publications / a−1


PEMEL
SOEL
40
unspecified

30

20

10

0
1990 2000 2010 2020
year / a
Figure 1. The number of publications per year from 1990 to 2019 containing the specified keywords.
Around 2010, the publication rate increases due to greater interest in the energy turnaround. While the
topic is often discussed technology-independently (unspecified), more publications for low-temperature
technologies, like alkaline water electrolysis (AEL) and proton exchange membrane electrolysis
(PEMEL), are available than for the high-temperature technology solid oxide electrolysis (SOEL) [26].

2. Alkaline Water Electrolysis


Alkaline water electrolysis is used to split water into the gases hydrogen and oxygen using electric
energy. The chemical reactions are given in the Equations (1)–(3). At the cathode, water molecules are
reduced by electrons to hydrogen and negatively charged hydroxide ions. At the anode, hydroxide
ions are oxidized to oxygen and water while releasing electrons. Overall, a water molecule reacts to
hydrogen and oxygen in the ratio of 2:1.

Cathode: 2 H2O(l) + 2 e– H2(g) + 2 OH–(aq) (1)


Anode: 2 OH–(aq) 0.5 O2(g) + H2O(l) + 2 e– (2)
Overall reaction: H2O(l) H2(g) + 0.5 O2(g) (3)

The required cell voltage for this electrochemical reaction can be determined by thermodynamics.
The free reaction enthalpy ΔR G in (4) can be calculated with the reaction enthalpy ΔR H,
the temperature T, and the reaction entropy ΔR S.

ΔR G = ΔR H − T · ΔR S (4)

The reversible cell voltage Urev in (5) is determined by the ratio of the free reaction enthalpy
ΔR G to the product of the number of exchanged electrons z = 2 and the Faraday constant F
(96,485 C mol−1 ) [28].

ΔR G
Urev = − (5)
z·F
At a temperature of 25 ◦C and an ambient pressure of 1 bar (standard conditions), the free reaction
enthalpy for the water splitting reaction is ΔR G = 237 kJ mol−1 , which leads to a reversible cell voltage
of Urev = −1.23 V. As the free reaction enthalpy is positive at standard conditions, the water splitting

7
Processes 2020, 8, 248

is a non-spontaneous reaction [28]. Due to irreversibilities, the actual cell voltage needs to be higher
than the reversible cell voltage for the water splitting reaction. The thermoneutral voltage Uth in
Equation (6) depends on the reaction enthalpy ΔR H, which is composed of the free reaction enthalpy
ΔR G and irreversible thermal losses T · ΔR S.

ΔR H
Uth = − (6)
z·F

At standard conditions, the reaction enthalpy for water electrolysis is ΔR H = 286 kJ mol−1 .
Hence, the thermoneutral voltage is Uth = −1.48 V [28].

3. System
A schematic flow diagram of alkaline water electrolysis is shown in Figure 2. The electrolyte is
pumped through the electrolysis stack, where the product gases are formed. While natural convection
can be a cost-efficient alternative, gas coverage of the electrode surface can raise the required cell
voltage and therefore increase the operational costs [29]. Additionally, most alkaline water electrolyzer
systems provide a temperature control for the electrolyte to maintain an optimal temperature range.

H2 O2

purification purification
demister/dryer demister/dryer
cathode anode
− +

gas separator gas separator

electrolysis stack
H2O
heat exchangers

pumps

equalization line

Figure 2. A schematic flow diagram of an alkaline water electrolyzer. The electrolyte is pumped
through the electrolysis cell where the gas evolution takes place. Adjacent gas separators split both
phases, and the liquid phase flows back to the electrolysis stack. Heat exchangers ensure that the
optimal temperature is maintained, and the product gases can be purified afterward.

The two-phase mixtures of liquid electrolyte and product gas leave the electrolysis cell and enter
subsequent gas separators. Mostly, the phase separation is realized with a high residence time in large
tanks. The product gas is demisted and dried before it is purified to the desired level [30]. The liquid
electrolyte leaves the gas separator and is pumped back to the electrolysis stack. As the product gases
are soluble in the electrolyte solution, the mixing of both electrolyte cycles causes losses and higher
gas impurities. An alternative can be to use partly separated electrolyte cycles with an equalization
line for liquid level balancing of both vessels [31,32]. With separated electrolyte cycles, the electrolyte
concentration will increase on the cathodic side due to water consumption and decrease on the anodic
side due to water production. Therefore, the electrolyte requires mixing, on occasion, to maintain an
optimal electrolyte conductivity.

8
Processes 2020, 8, 248

4. Cell Design and Cell Voltage


The design of the electrolysis stack depends on the manufacturer; however, some general
similarities can be observed. Two variants of cell designs are shown in Figure 3. Earlier alkaline
water electrolyzers used a conventional assembly with a defined distance between both electrodes.
Later, this concept was replaced by the zero-gap assembly, where the electrodes are directly pressed
onto the separator to minimize ohmic losses due to the electrolyte. Porous materials like Zirfon™ Perl
UTP 500 (AGFA) or dense anion exchange membranes can be used as the separator [33–37].

(a) conventional (b) zero-gap

cathode anode
− +

H2 O2 H2 O2
cathode anode
separator
− +
electrode distance

Figure 3. Different cell designs for alkaline water electrolysis. Whereas (a) shows a conventional
assembly with a defined distance between both electrodes, (b) depicts a zero-gap assembly where the
electrodes are directly pressed onto the separator [38].

During operation, the required cell voltage is always higher than the reversible cell voltage due to
different effects. A calculated cell voltage profile is displayed in Figure 4. In addition to the ohmic
losses, I · Rohm , there are activation overvoltages of the electrodes, ηact . The ohmic resistance of the cell
design is affected by the electronic conductivity of the electrode material, the specific conductivity of
the electrolyte, the ionic conductivity of the separator material, and gas bubble effects.

2:2

2
cell voltage / V

1:8

1:6 ”act

1:4
I · Rohm
1:2
Urev
1
0 0:1 0:2 0:3 0:4 0:5
current density / A cm−2
Figure 4. The calculated cell voltage of an atmospheric alkaline water electrolyzer at a temperature
of 60 ◦C according to Equation (8). The overall cell voltage consists of the reversible cell voltage Urev ,
ohmic losses I · Rohm , and activation overvoltages ηact [39,40].

9
Processes 2020, 8, 248

The zero-gap design tries to eliminate the electrolyte losses by minimizing the electrode distance.
There is still a minimal gap between both electrodes, which can increase the cell voltage. The activation
overvoltages are defined by the electrode materials. Whereas nickel is the most-used electrode
material, it provides very high overvoltages for the oxygen and hydrogen evolution reactions [41–44].
Hence, electrocatalytic materials are added to the electrodes. Iron is a cost-efficient catalyst for the
oxygen evolution reaction [41,42,45]. Molybdenum decreases the overvoltage for the evolution of
hydrogen at the cathode [44,46,47].
Several authors have proposed correlations for the modeling of cell voltage. Equation (7) considers
the operation temperature ϑ and the current density j by describing the dependencies with empirical
parameters. While the parameters ri reflect ohmic losses, s and ti stand for the activation overvoltages
of the hydrogen and oxygen evolution reactions [28].
  
t2 t
Ucell = Urev + (r1 + r2 · ϑ ) · j + s · log t1 + + 32 ·j+1 (7)
ϑ ϑ

This correlation can be extended with the effects of the operation pressure p in (8) by adding
the empirical parameters di , which specify the additional losses owing to pressurized operation [39].
In general, the reversible cell voltage increases with the pressure; however, the ohmic resistance caused
by the gas bubbles decreases as the bubble diameter becomes smaller. Hence, both effects equalize
each other and only small differences can be observed [48].
  
t2 t
Ucell = Urev + [(r1 + d1 ) + r2 · ϑ + d2 · p] · j + s · log t1 + + 32 ·j+1 (8)
ϑ ϑ

The correlations (7) and (8) are empirical and therefore only valid for the actual system to which
they are adjusted. The correlation parameters and a suitable equation for the reversible cell voltage
under atmospheric conditions can be found in the Appendix A in Table A1 and Equation (A1). Other
authors have proposed physically reasonable models based on actual dimensions and properties of
the system rather than on empirical correlations.
An example for such an approach is Equation (9), in which the terms are split into experimentally
determinable parts [49].

Ucell = Urev + ηact


c
+ ηact
a
+ I · ( Rc + Ra + Rele + Rmem ) (9)

The cell voltage Ucell is calculated with the reversible voltage Urev , the activation overvoltages ηact ,
and the ohmic resistances. Whereas Rc and Ra represent the reciprocal electronic conductivity of the
electrode materials, Rele stands for the ohmic loss caused by the electrolyte conductivity. Additionally,
the ohmic resistance Rmem of the separator material is taken into account. The activation overvoltages
ηact can be calculated with the Butler–Volmer equation. In most cases, the simplified Tafel equation is
sufficient to describe the resulting overpotentials [40]. The required Tafel slope and exchange current
density can be extracted from experimental data. Hence, those parameters are only valid for the
actual system design; however, they can be easily replaced by other data when needed. As the
ohmic resistances of the electrodes (Rc and Ra ) only depend on the electronic conductivity and
electrode dimensions, both values are known. In most cases, the ohmic resistance of the electrode is
comparably small and can be neglected. The electrolyte resistance Rele is determined by the specific
electrolyte conductivity and the cell design. Whereas the electrolyte gap is minimal in zero-gap
designs, conventional setups maintain a defined distance between both electrodes. As the specific
conductivity of the electrolyte gap is affected by gas bubbles, there is an optimal electrode distance
for conventional designs [50]. If the electrode distance is too small, the gas bubbles accumulate in the
gap and lower the conductivity. With increasing distance, the bubble detachment is enhanced and the
specific conductivity increases. It is a trade-off between a small electrolyte gap—as the ohmic resistance
increases linearly with this parameter—and a better conductivity of the space between both electrodes.

10
Processes 2020, 8, 248

In addition to the decreasing electrolyte conductivity with higher amounts of gas bubbles, the active
electrode surface can be blocked by gaseous compounds, which leads to additional losses [49]. As this
phenomenon depends on the cell design and operation concept, there are difficulties in describing it
properly. Therefore, it is often neglected, or empirical correlations referring to the gas hold-up are
utilized [49].
Furthermore, the installed separator material also has significant ohmic losses. While the porous
separator Zirfon™ Perl UTP 500 is often used, anion exchange membranes are promising alternatives.
For Zirfon™ -based materials, experimental data of the resistance at a fixed electrolyte concentration
for different temperatures are available [51].
The most-used electrolyte for alkaline water electrolysis is an aqueous solution of potassium
hydroxide (KOH) with 20 to 30 wt.% KOH, as the specific conductivity is optimal at the typical
temperature range from 50 to 80 ◦C [25]. A cheaper alternative would be a diluted sodium hydroxide
solution (NaOH), which has a lower conductivity [52]. Calculated specific electrolyte conductivities for
both electrolyte solutions at different temperatures are shown in Figure 5. While KOH provides
a specific conductivity around 95 S m−1 at 50 ◦C, NaOH reaches a value around 65 S m−1 . At a
temperature of 25 ◦C, a similar effect can be seen. The conductivity of KOH is around 40 to 50%
higher than the conductivity of a NaOH solution at the optimal weight percentage. Another aspect
is the solubility of the product gases inside the electrolyte, as this influences the resulting product
gas purity. In general, the gas solubility decreases with an increasing electrolyte concentration due to
the salting-out behavior [53]. NaOH also shows a slightly higher salting-out effect than that of KOH.
Hence, the product gas solubility is higher in a KOH solution [54–56].

120
specific conductivity / S m−1

100 KO
H,
50 ◦
C
80
KO
C

H,
0

60 25
,5


C
H
aO
N

40

5
,2
H
aO
N

20

0
0 10 20 30 40 50
electrolyte concentration / wt:%
Figure 5. The calculated specific electrolyte conductivity as a function of the electrolyte concentrations
of sodium hydroxide (NaOH) and potassium hydroxide (KOH) solutions at different temperatures
obtained by Equations (A2) and (A3). The correlation parameters can be found in Table A2 [52,57].

Another approach is to use ionic liquids (ILs) as the electrolyte or as an additive, owing to
their remarkable properties [5,6,21]. Ionic liquids are organic substances which are liquid at room
temperature and are electrically conductive [58]. A negligible vapor pressure, non-inflammability,
and thermal stability are promising arguments for their utilization in water electrolysis. Furthermore,
ILs can be used over a wide electrochemical window [59]. The absorption and separation of gases
is an additional area of application [60,61]. However, the toxicity of ILs is a current field of
research, and the viscosity is comparably high, which should be taken into account before any
large-scale implementation [6,58,59]. In addition to providing high-efficiency water electrolysis at low
temperatures, ILs are chemically inert and therefore do not require expensive electrode materials [62].

11
Processes 2020, 8, 248

5. Gas Purity
Gas purity is an important criterion of alkaline water electrolysis. While the produced hydrogen
typically has a purity higher than 99.9 vol.% (without additional purification), the gas purity of oxygen
is in the range of 99.0 to 99.5 vol.% [48]. As both product gases can form explosive mixtures in the
range of approximately 4 to 96 vol.% of foreign gas contamination, technical safety limits for an
emergency shutdown of the whole electrolyzer system are at a level of 2 vol.% [31,63]. Therefore,
the product gas impurity needs to be below this limit during operation to ensure continuous production.
Experimentally determined anodic gas impurities for alkaline water electrolysis are presented in
Figure 6 for different operation modes. The current densities are in the range from 0.05 to 0.7 A cm−2
and the system pressures range from 1 to 20 bar [64].

1 4
(a) separated (b) mixed

0:8
3
H2 in O2 / vol:%

0:6 H2 in O2 / vol:%
50 % lower explosion limit
2
0:4

1
0:2

0 0
0 0:2 0:4 0:6 0 0:2 0:4 0:6
−2
current density / A cm current density / A cm−2
1 bar 10 bar 20 bar

Figure 6. Anodic gas impurity (H2 in O2) in relation to the current density at different pressure
levels for (a) separated and (b) mixed electrolyte cycles, at a temperature of 60 ◦C, with an electrolyte
concentration of approximately 32 wt.% and an electrolyte volume flow of 0.35 L min−1 [64].

While the gas impurities with separated electrolyte cycles are below 0.7 vol.% for all tested current
densities and pressure levels, mixing of the electrolyte cycles increases the gas impurity significantly.
Furthermore, two similarities can be seen. The gas impurity lowers with increasing current density and
increases at higher pressure levels. Both effects are physically explainable. While the contamination
flux stays constant with varying current densities, the amount of produced gas becomes lower in
a linear relationship. Hence, at a higher current density, the contamination is more diluted than
at a lower current density [32,64]. As a consequence, the operation in the part-load range is more
critical due to the higher gas impurity. The amount of dissolved product gas increases with pressure;
thus, high concentration gradients for the diffusion through the separator material are available, and
more dissolved foreign gas reaches the other half-cell when mixing [64]. However, operation at slightly
elevated pressures is favorable, as the costly first mechanical compression level can be avoided by
the direct compression inside the electrolyzer system [65]. With mixed electrolyte cycles, the gas
impurity reaches critical values even at higher current densities during pressurized operation. While
at atmospheric pressure, the gas impurity is only at a current density of 0.05 A cm−2 , slightly above
the safety limit of 2 vol.% H2 in O2, this limit is already reached at 0.5 A cm−2 for a system pressure of
10 bar. At 20 bar, no sufficient gas purity could be measured, as even a current density of 0.7 A cm−2
results in a gas impurity of 2.5 vol.%.

12
Processes 2020, 8, 248

6. Periphery
The operation of alkaline water electrolyzers is also affected by the installed periphery, including
power supplies. The output signal may contain a specific number of ripples, which directly influences
the process performance [66]. Power supplies with a high ripple propensity lower the overall efficiency
and, therefore, signal smoothing is necessary. The ripple formation is avoidable, but the component
costs will be higher [67]. In general, thyristor-based power supplies tend to deliver a higher degree
of fluctuation, and transistor-based systems output a smoother signal. Additionally, a higher ripple
frequency does not affect the system performance as much as the occurrence of low-frequency
ripples [68]. Furthermore, a coherence between the ripple behavior of a power supply and the
product gas quality of alkaline water electrolysis can be observed [69].

7. Renewable Energy
The combination of alkaline water electrolysis with renewable energy is essential for sustainable
hydrogen production without significant carbon dioxide emissions. While solar and wind energy are
often favored due to their wide availability, other renewable energies, such as hydropower, biomass,
and geothermal energy, are frequently utilized for the base load [7]. The direct usage of renewable
energy in the power grid is difficult due to the mismatch between energy demand and production
and the limited storage possibilities for electricity. Hence, excess electric energy should be chemically
stored in hydrogen for later usage [6]. Due to the fluctuating and intermittent behavior of solar and
wind power, alkaline water electrolyzers must be adapted to a dynamic operation. To evaluate the
requirements, local weather data can be used to extract the amplitudes and frequencies of fluctuations.
Typical time-related profiles for solar radiation and wind velocity are shown in Figure 7. The data
were measured by the weather station of the Clausthal University of Technology on the rooftop of a
university building. Whereas the wind velocity shows a mean value of around 3.8 m s−1 , the significant
solar radiation is only available during the daytime. Hence, the averaged value over the whole day is
233 W m−2 for a sunny day and only 29 W m−2 for a cloudy day.

600 10
(a) (b)

8
solar radiation / W m−2

wind velocity / m s−1

400 sunny day


6

4
200
cloudy day 2

0 0
0 6 12 18 24 0 6 12 18 24
time / h time / h
Figure 7. Typical time-related profiles for (a) solar radiation and (b) wind velocity, measured by the
weather station of the Clausthal University of Technology. Though solar radiation peaks around noon,
wind velocity shows sinusoidal oscillations.

The volume flow of the produced hydrogen directly follows the renewable energy profile used
for operation [70]. Only a short delay is noticeable for the gas purity, which is defined by the system
volume [71]. Due to the possibility of direct coupling of water electrolysis and photovoltaic panels,
this technology is highly appropriate for renewable hydrogen production [29,72,73]. As photovoltaic

13
Processes 2020, 8, 248

panels require high investment costs, wind power is often favored for large-scale hydrogen production.
In comparison with photovoltaic power, wind power shows a higher degree of fluctuation and is very
intermittent. Therefore, the dynamic operation of alkaline water electrolyzers is more challenging [4].
Hence, the dynamic behavior of an alkaline water electrolyzer can be used to develop
suitable system designs and to operate existing systems safely and efficiently. As measurements
of solar radiation and wind velocity are often available for a given location, the theoretically
available renewable energy can be calculated and used as an input during the system design.
Different approaches exist for the calculation of solar photovoltaic power and wind turbine power.
While the current–voltage characteristics of photovoltaic panels can be expressed as a function of
manufacturer data and solar radiation, the power of wind turbines is a fraction of the maximum
available wind power, which is defined by the wind speed and a performance coefficient [72,74].

7.1. Solar Photovoltaic Power


The behavior of photovoltaic panels can be described by single-diode and two-diode models with
varying degrees of complexity. Often, the solution must be obtained iteratively or with numerical
methods when very detailed models are utilized [75,76]. Simple models with analytical solutions
are a recent research topic, as a short processing time can be needed for online characterization and
the optimization of existing systems [75]. In Figure 8, the coupling possibilities of an alkaline water
electrolyzer and solar photovoltaic panels are shown. Additional losses of a DC/DC transformer can
be avoided when a direct coupling of the systems can be realized. Otherwise, the transformation
ensures a fit of both systems by an indirect coupling [73,77,78].

H2 O2
photovoltaic panel (optional)

energy energy alkaline water


DC/DC H2O
converter electrolysis

Figure 8. Schematic of alkaline water electrolysis powered by solar energy. Photovoltaic panels convert
the solar radiation into electricity, which can be used for the operation. The implementation of a
DC/DC power converter is optional, as direct and indirect coupling is possible [70,78,79].

When a direct coupling of both systems is to be realized, the possible operation points can be
determined by the intersection of the current–voltage curves. A typical current–voltage characteristic
of an alkaline water electrolyzer is given by (8). The resulting current of a photovoltaic cell IPV at
different solar radiation levels can be described by (10) with a suitable single-diode model as a function
of the voltage UPV [29,72,73]. Therefore, specific data from the photovoltaic (PV) panel and the ambient
conditions are required in order to calculate the photocurrent Iph , the reverse saturation current Is ,
and the thermal voltage UT . Furthermore, the serial Rs and parallel Rp resistance of the photovoltaic
panel must be available.
   
UPV + IPV · Rs U + IPV · Rs
IPV = Iph − Is · exp − 1 − PV (10)
UT Rp

The photocurrent Iph is defined in (11), which shows a linear relationship with the solar radiation
Esun absorbed by the photovoltaic cell. A higher cell temperature Tc increases the photocurrent.
 
Iph = 0.003 m2 V−1 + 10−7 m2 V−1 K−1 · Tc · Esun (11)

The reverse saturation current Is can be calculated by (12) with the short-circuit current Isc ,
the open-cell voltage Uoc , and the thermal voltage UT . Whereas the short-circuit current and the

14
Processes 2020, 8, 248

open-cell voltage are provided by the manufacturer, the thermal voltage depends on the physical
properties.

Isc
Is =   (12)
Uoc
exp UT − 1

An equation for the thermal voltage is given in (13), which is based on the Boltzmann constant
kB (1.3806·10−23 J K−1 ) and the electron charge e (1.602 19·10−19 C) [72]. Additionally, the number of
serially connected cells, ns , and the cell temperature are required. Furthermore, the non-ideality factor
m contains any deviations from the theoretical behavior.

ns · kB · Tc
UT = m · (13)
e
In addition to these equations, the calculation of the resulting current of a photovoltaic cell requires
knowledge of the serial (Rs ) and parallel (Rp ) resistance of the system. By adding parallel photovoltaic
cells, the current multiplies by the amount of parallel paths np . Suitable parameters of an existing
photovoltaic cell setup are given in Table 1. For this exemplary calculation, a constant temperature of
the photovoltaic cell is assumed. Otherwise, the cell temperature increases with the absorbed solar
radiation. While simple linear approaches already result in a good agreement with experimental data,
a complete energy balance is the best way to determine the temperature exactly [29,72].

Table 1. Parameters for the example calculation of the photovoltaic current using Equation (10).
The number of serial ns and parallel np connected photovoltaic cells, the short-circuit current Isc ,
the open-cell voltage Uoc , the serial Rs and parallel resistance Rp , and the non-ideality factor m are
setup-specific data. A constant cell temperature Tc is assumed [29,72,73].

ns np Isc Uoc Rs Rp m Tc
– – A V Ω Ω – ◦C

9 2 5.98 4.615 0.099 20 1.6 48

The results of the example calculation are shown in Figure 9. The current–voltage characteristics
are given for different solar radiation levels from 200 to 1000 W m−2 , in combination with a typical
polarization curve of an alkaline water electrolyzer (10 cm2 electrode area) from (8) in Figure 9a.
The power–voltage curves for the photovoltaic cell are shown in Figure 9b. The maximal power point
(MPP) for each radiation level is marked with a dot in both diagrams.
In Figure 9a, the characteristics of the alkaline water electrolyzer deviate from the MPP curve.
Therefore, the photovoltaic cell cannot deliver the maximal power, and the overall efficiency decreases.
Hence, both systems should be optimized until the alkaline water electrolyzer performs close to the
maximal power output [73,80]. The alternative would be an indirect coupling of both systems with the
integration of a DC/DC converter, which also implies losses, with an efficiency of around 90% [81,82].

15
Processes 2020, 8, 248

20
(a) (b)
1000 W m−2 AEL
6 MPP
MPP
15

2
m−
800 W m−2

W
current / A

power / W

00
4

10

2
m−
600 W m−2
10

0W
80
2

m
400 W m−2 0 W
60
2 −2
5 400
W
m

200 W m−2
−2
m
200 W

0 0
0 1 2 3 4 5 0 1 2 3 4 5
voltage / V voltage / V
Figure 9. Example calculation results of the (a) current–voltage characteristics of a photovoltaic
panel at different solar radiation levels and the corresponding (b) power–voltage curve.
Additionally, a current–voltage characteristic of an alkaline water electrolyzer (AEL) is implemented.
The intersections determine the possible operation points. For an efficient operation, the distance to the
maximal power points (MPP) should be minimal [29,72,73].

7.2. Wind Power


As the power from photovoltaic cells is only available during the daytime, wind power is another
important energy source for the renewable production of hydrogen. The schematic concept is shown
in Figure 10. For the implementation of conventional wind turbines, an AC/DC converter is essential.
The efficiency of an AC/DC conversion is also approximately 90% [82,83].

H2 O2

wind energy energy alkaline water


AC/DC H2O
turbine converter electrolysis

Figure 10. Schematic of alkaline water electrolysis powered by wind energy. Wind turbines convert the
available wind power into electricity, which can be used for the operation. The implementation of a
suitable AC/DC converter is mandatory [74,79].

For the calculation of the wind turbine power, the exact wind velocity at the height of the turbine
rotor should be known. Often, the wind velocity is measured at rooftops or special measurement
facilities with a defined height of approximately 10 m, which is significantly lower than the height of
a wind turbine, around 100 m [84]. Therefore, the measured data should be corrected to the desired
height by (14).
 
ln zwind
z0
vwind = vwind,ref ·  z  (14)
ln wind,ref
z0

16
Processes 2020, 8, 248

The wind velocity vwind at the height zwind can be determined from the measured wind velocity
vwind,ref at the height zwind,ref in combination with the roughness of the terrain z0 [48]. To obtain the
output power of a wind turbine Pturbine , first, the theoretical wind power Pwind needs to be calculated
using (15). Therefore, the air density ρ (from 1.22 to 1.3 kg m−3 ), the area spanned by the rotor blades
A, and the wind velocity are needed [74,85].

1
Pwind = · ρ · A · v3wind (15)
2
The maximal wind power cannot be completely converted into wind turbine power.
This circumstance is considered by the implementation of the performance coefficient CP , which
lowers the maximal reachable power output. The actual wind turbine power results from the product
of the wind power and the performance coefficient in (16).

Pturbine = Pwind · Cp (16)

The determination of the correct performance coefficient is a complete research topic in itself,
which consists of empirical correlations and computational fluid dynamics (CFD) simulation studies.
Often, experimental data are used to fit the correlations to the measurements [74]. An example equation
for the performance coefficient is shown in (17) [74,79].
   
116 12.5
Cp = 0.22 · − 0.4 · β − 5 · exp − (17)
λi λi

Therefore, the pitch angle of the turbine blades β has to be defined and the tip speed ratio λ
needs to be calculated in (18) from the turbine blade radius R, the rotational speed ω, and the wind
speed [74].

R·ω
λ= (18)
vwind

The calculation of the performance coefficient also requires the parameter λi , which is described
by (19) based on the tip speed ratio and the blade pitch angle [74].

1 1 0.035
= − 3 (19)
λi λ + 0.08 · β β +1

For the blade radius, a value of 46.5 m is assumed, which is a typical blade length for a wind
turbine with a rated power of 2 MW [74]. In Figure 11, the calculation results for the performance
coefficient, depending on the tip speed ratio and the turbine power at different wind velocities, are
shown. The performance coefficient of conventional wind turbines is limited at Cp = 0.593 [74]. In this
example, a maximal performance coefficient of approximately Cp = 0.450 is reached for a blade pitch
angle of β = 0°. With an increasing pitch angle, the maximum of the performance coefficient decreases
and shifts towards smaller tip speed ratios.

17
Processes 2020, 8, 248

0:5 3
(a) (b)
0° MPP
performance coefficient / 1

12
0:4 3°

m
turbine power / MW

s

1

2

0:3

10
:8
12°

m
s

15°

1
9:
0:2 6
m
s−
1 1
8:
4m
s −1
0:1 7:2
m −
s 1
6m
s −1

0 0
0 5 10 15 0 10 20 30 40
−1
tip speed ratio / 1 rotational speed / min
Figure 11. Example calculation results of (a) the performance coefficient for various rotor blade
pitch angles using Equation (17) and (b) the wind turbine power for different wind velocities using
Equation (16). The maximum power point (MPP) trajectory is marked [74,79].

For the calculation of the turbine power in Figure 11b, a pitch angle of β = 6° is assumed.
With increasing wind velocity, the value of the maximal power point (MPP) becomes higher and shifts
towards faster rotational speeds. The rated wind speed of this exemplary wind turbine is at 11 m s−1
with rotational speeds from 6 to 17 min−1 . The cut-in wind speed is 3 m s−1 and the cut-out wind speed
is 22 m s−1 [74]. In comparison with the power characteristics of photovoltaic panels, the polarization
curve of alkaline water electrolyzers can not be directly optimized towards the MPP trajectory, as the
optimal operation point highly depends on the wind turbine design and weather conditions. Therefore,
an efficient AC/DC converter is the best option for maintaining an efficient operation of an alkaline
water electrolyzer [82].

8. Hydrogen Energy System and Power Grid Stabilization


An exemplary process scheme for a hydrogen energy system is provided in Figure 12. Photovoltaic
panels and wind turbines are connected with suitable converters to a DC bus, from which alkaline
water electrolyzers are powered. The produced hydrogen can be stored for later application in fuel cells.
To raise the fuel cell efficiency, the produced oxygen can be used instead of air. Therefore, an additional
storage tank must be available, which incurs further costs [86].
The fuel cells are also connected to the DC bus, and the power can be used by the electricity
grid with DC/AC converters. At lower energy demands, hydrogen can be produced and converted
back into energy when it is needed. As conventional alkaline water electrolyzers are designed for
operation at constant conditions, occurring fluctuations may be damped by additional energy storage
devices like batteries, supercapacitors, or flywheels [25,28,82]. When excess energy is available, this
energy storage can be charged to be fully available when needed. The damping quantity is limited to
a certain degree of fluctuation, as the energy storage amount is also restricted to the capacity of all
installed devices. Additionally, the produced hydrogen can also be used for the decarbonization of
industrial processes or as a fuel in the transport sector [87–89]. To raise the overall efficiency, some
DC/DC converters could be neglected by optimized system designs by lowering the system flexibility.
Furthermore, when the alkaline water electrolyzers are able to operate under dynamic conditions,
additional energy storage devices are not required or, at least, the number of such devices could
be lowered. There are still some challenges for electrolyzer manufacturers to overcome before this
possibility becomes available.

18
Processes 2020, 8, 248

grid connection
wind
energy
photovoltaic storage

DC/DC AC/DC DC/AC DC/DC

DC bus
DC/DC DC/DC
alkaline water fuel
electrolysis cell
H2
H2 storage H2

H2O optional H2O


O2 O2 O2 or air
storage

Figure 12. The schematic process scheme of a hydrogen energy system. Photovoltaic panels and wind
turbines generate renewable energy to power alkaline water electrolyzers, and stored hydrogen can be
converted back into electricity by fuel cells. Therefore, either oxygen or air can be utilized. Additional
energy storage devices can damp fluctuations, and the complete hydrogen energy system can be used
for power grid stabilization [25,28,82,87].

With an increasing share of renewable energies in the power grid, it is difficult to maintain a
constant power frequency. Such hydrogen energy systems or alkaline water electrolyzers can be
used to stabilize the power frequency by damping the fluctuations. An additional benefit would be
the reduction of the conventional spinning reserve, which reduces costs and CO2 emissions [87,90].
A predictive control can be used for stable and efficient operation. Pressurized alkaline electrolyzers
are more suitable for damping fast fluctuations, whereas atmospheric units can handle the slow
fluctuations [87].

9. Limitations and Solution Approaches


The implementation of a hydrogen energy system into the existing power grid is a challenging
task with some limitations which must be overcome in order to guarantee high system availability.
The main problem of an alkaline water electrolyzer powered by renewable energy is the high gas
impurity in the part-load range, which can cause a safety shutdown when reaching a foreign gas
contamination of 2 vol.% [31,91]. Hence, the annual operation time is limited to the time spans with
sufficient renewable energy [91].

9.1. Limited Operation Time


The limited operation time leads to a high number of startup and shutdown cycles, which can
exceed the maximal start/stop count defined by the manufacturer and, therefore, can lower the
expected system lifetime or warranty agreements. Mainly, the electrodes are affected by the repetitive
start/stop behavior and the electrode degradation is accelerated [48,82]. Nickel electrodes are known
to degrade significantly after 5000 to 10,000 start/stop cycles. When operating with photovoltaic
power only, 7000 to 11,000 cycles are already reached in the period of 20 to 30 years. The fluctuating
nature of renewable energy amplifies the electrode degradation, as this phenomenon acts partly as a
start/stop process [92]. This issue can be solved by the development of stable electrode compositions
or self-repairing electrode surfaces [92].

19
Processes 2020, 8, 248

To circumvent the drawbacks of having only one renewable power source, such as in the
daytime-limited operation with solar power, the combination of several energy sources enhances
the overall efficiency. While the operation with only PV shows a faradaic efficiency of approximately
40%, wind power leads to a faradaic efficiency of around 80%. The combination of both technologies
enhances the faradaic efficiency above 85% [79].
To hinder the gas impurity from reaching the lower explosion limit, the part-load range of most
alkaline electrolyzers is limited to 10 to 25% of their nominal load [82,91]. Fluctuations below the
minimal load can be balanced out with the implementation of energy storage devices, as shown
in Figure 12; however, in some scenarios, the available energy storage will not be sufficient. When
the gas impurity is still in a tolerable region, short periods without an electrode polarization can
be allowed. The cathode starts to degrade noticeably below a voltage of around 0.25 V [82]. Thus,
the complete shutdown can be held until reaching this voltage limit. The available time depends on the
electrode composition, as the electrochemical double layer acts as a capacitor and delays the voltage
breakdown after a power loss. Experimentally, a time span of around 10 min has been reported [82].

9.2. Optimal System Design and Operation Strategies


To mitigate the rise of gas impurities during low power availability, an optimal system design
can allow enough time until sufficient energy is available again. While the gas volume inside the
system acts as a buffer tank and dilutes the gas contamination, the liquid and the solid volume of an
electrolyzer buffers the system temperature during part-load operation [25,71].
Furthermore, to maintain an efficient operation, the system temperature has to be in an
optimal range of 50 to 80 ◦C for an electrolyte solution with 20 to 30 wt.% KOH [25]. As most
renewable-energy-powered alkaline water electrolyzers will not provide a separate heating unit,
the temperature needs to be reached and maintained only by the heat of the reaction [4]. Temperatures
above 80 ◦C should be avoided with a suitable cooling system to prevent high degradation rates.
An alternative would be the operation at low temperatures to damp electrode degradation, but then,
very active electrocatalysts are needed to reach a sufficient efficiency [86].
More experimental and theoretical work is needed to fully understand the dynamic behavior
of alkaline water electrolyzers powered by renewable energy [25]. In addition to an optimal system
design, suitable dynamic operation strategies can be beneficial for lowering the gas impurity. While
low gas impurities occur with separated electrolyte cycles, high gas impurities result in combined
mode. The measured stationary gas impurities in Figure 6 are reached after a specific duration. When
the electrolyzer is able to switch between both operation modes automatically, this can be used to
switch to the separated mode when the gas impurity is too high, and then combine again when a
sufficient gas production rate is available. Experimental work shows that the gas impurity can be
almost halved by this approach [31].
The primary reason for high gas contamination is the continuous operation at low current densities.
This circumstance can be prevented by reducing the overall cell area (overloading) or by subdividing
the system into several smaller blocks [91]. While the implementation of electrolyzers with smaller
electrode areas also limits the maximal load compared to larger systems, partial system operation is
a more elegant method. During low power availability, single stacks or compartments of a system
with multiple stacks can be powered off, which lowers the available electrode area and therefore
results in higher current densities [93]. Obviously, this strategy causes problems in maintaining the
optimal system temperature due to the disabled components. An alternative method to prevent
adverse process states is the use of predictive control systems. For example, when low renewable
power availability is forecasted, the system can change the temperature, pressure, or operation mode
to a more suitable state before negative effects occur [87].

20
Processes 2020, 8, 248

10. Conclusions
The combination of alkaline water electrolysis and renewable energy for sustainable hydrogen
production is an essential step towards the decarbonization of industrial processes and the transport
sector [87–89]. To determine the most relevant limitations and to propose suitable solution approaches,
the technologies have to be fully understood [25]. Whereas the process of alkaline water electrolysis can
be defined by current–voltage characteristics and the resulting gas impurity, photovoltaic panels and
wind turbines should be operated at the maximal power point [73,74,79]. Therefore, the influencing
parameters must be known. Different model approaches exist, out of which the most suitable one
should be chosen. While empirical correlations are often only valid for the specific experimental setup,
physically reasonable models can be used in a more general way to develop new solutions. For alkaline
water electrolysis, many experimental and theoretical data are available to calculate and analyze the
cell voltage under operation conditions. As the actual system design and cell arrangement differ
for every electrolyzer, certain parameters have to be determined experimentally to use the proposed
models for another system. Mainly, this issue exists for electrode compositions and separator materials.
To describe the gas purity of hydrogen and oxygen mathematically, only models and correlations on an
empirical basis are currently available due to the high number of influencing variables [31,32]. As the
gas impurity mainly determines the system availability of an alkaline water electrolyzer, more research
for the development of physically-based models is needed. The dynamic system behavior should
be analyzed, as optimized dynamic operation strategies can be beneficial for the overall system
efficiency. Many models with different complexity levels are available for the description of the
current–voltage characteristics of photovoltaic panels. Most models rely on physical principles and
manufacturer data [75]. Thus, proper modeling for different systems is possible. The power conversion
by wind turbines can be described by system properties and suitable correlations for the performance
coefficient [74]. As this variable is influenced by many parameters, including the design of the turbine
blades, the correlation should only be used for very similar wind turbines, or the parameters must be
determined experimentally or by simulation.
To conclude, there are appropriate models available for all components of a hydrogen energy
system. However, some descriptions need further improvement to be applicable to a variety of different
system designs. With this knowledge and with experimental studies, many researchers have already
examined the limitations of renewable-powered alkaline water electrolyzers [48,79,82]. The central
prospect is to increase the operation time through intelligent system designs and advantageous
operational concepts. While the implementation of conventional energy storage devices to damp the
dynamics is a first logical step, alkaline water electrolyzers should be enabled to handle all dynamics
directly to reduce costs and to enhance the efficiency [25]. As the hydrogen production from fossil
energy carriers is less expensive than hydrogen from electrolysis processes, only optimized systems
with the use of excess renewable energy can be competitive.

Author Contributions: Conceptualization, methodology, software, validation, formal analysis, investigation,


data curation, writing—original draft preparation, visualization, J.B.; resources, writing—review and editing, J.B.
and T.T.; supervision, project administration, funding acquisition, T.T. All authors have read and agreed to the
published version of the manuscript.
Funding: This work is funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)
Project numbers: 290019031; 391348959.
Acknowledgments: The authors thank the Institute of Electrical Information Technology (IEI) of the Clausthal
University of Technology for providing the weather data.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.

21
Processes 2020, 8, 248

Abbreviations

AC Alternating current
AEL Alkaline water electrolysis
CFD Computational fluid dynamics
DC Direct current
HTEL High-temperature electrolysis
ILs Ionic liquids
MPP Maximum power point
PEMEL Proton exchange membrane electrolysis
PV Photovoltaic
SOEL Solid oxide electrolysis

Appendix A. Correlations and Parameters


A correlation for the reversible cell voltage Urev of alkaline water electrolysis is given in (A1).
The obtained value can be used for the calculation of the cell voltage in (7) or (8) at atmospheric
conditions. For a pressurized system, extended correlations are required, as the reversible cell voltage
increases at higher pressures [40]. The empirical correlation parameters for the calculation of cell
voltage by (7) and (8) are given in Table A1.
   2
T T
Urev = 1.503 42 V − 9.956 · 10−4 V · + 2.5 · 10−7 V · (A1)
K K

Table A1. Parameters for the calculation of cell voltage by Equations (7) and (8) [28,39,94].

Parameter Equation (7) [28,94] Equation (8) [39] Unit


r1 8.05·10−5 4.451 53·10−5 Ω m2
r2 −2.5·10−7 6.888 74·10−9 Ω m2 ◦ C − 1
s 0.185 0.338 24 V
t1 1.002 −0.015 39 m2 A − 1
t2 8.424 2.001 81 m2 ◦C A−1
t3 247.3 15.241 78 m2 ◦C2 A−1
d1 – −3.129 96·10−6 Ω m2
d2 – 4.471 37·10−7 Ω m2 bar−1

The correlations for the calculation of specific electrolyte conductivity for KOH and NaOH can
be found in (A2) and (A3). The required correlation parameters are listed in Table A2. The validity
range for (A2) is a temperature T from 258.15 to 373.15 K and KOH mass fractions wKOH between 0.15
and 0.45. Equation (A3) is valid for temperatures ϑ between 25 and 50 ◦C and NaOH mass fractions
wNaOH from 0.08 to 0.25 [52,57].

σKOH =K1 · (100 · wKOH ) + K2 · T + K3 · T 2 + K4 · T · (100 · wKOH )


T (100 · wKOH ) (A2)
+K5 · T 2 · (100 · wKOH )K6 + K7 · + K8 ·
(100 · wKOH ) T

σNaOH = K1 + K2 · ϑ + K3 · wNaOH
3
+ K4 · wNaOH
2
+ K5 · wNaOH (A3)

22
Processes 2020, 8, 248

Table A2. Parameters for the calculation of the specific electrolyte conductivities of KOH and NaOH
solutions by Equations (A2) and (A3) [52,57].

Parameter Equation (A2) [57] Unit Equation (A3) [52] Unit


K1 27.984 480 3 S m−1 −45.7 S m−1
K2 −0.924 129 482 S m−1 K−1 1.02 S m−1 ◦C−1
K3 −0.014 966 037 1 S m−1 K−2 3200 S m−1
K4 −0.090 520 955 1 S m−1 K−1 −2990 S m−1
K5 0.011 493 325 2 S m−1 K−2 784 S m−1
K6 0.1765 – – –
K7 6.966 485 18 S m−1 K−1 – –
K8 −2898.156 58 S K m−1 – –

References
1. Sherif, S.; Barbir, F.; Veziroglu, T. Towards a Hydrogen Economy. Electr. J. 2005, 18, 62–76. [CrossRef]
2. Suleman, F.; Dincer, I.; Agelin-Chaab, M. Environmental Impact Assessment and Comparison of Some
Hydrogen Production Options. Int. J. Hydrog. Energy 2015, 40, 6976–6987. [CrossRef]
3. Dincer, I.; Acar, C. Review and Evaluation of Hydrogen Production Methods for Better Sustainability. Int. J.
Hydrog. Energy 2015, 40, 11094–11111. [CrossRef]
4. Gandía, L.M.; Oroz, R.; Ursúa, A.; Sanchis, P.; Diéguez, P.M. Renewable Hydrogen Production: Performance
of an Alkaline Water Electrolyzer Working under Emulated Wind Conditions. Energy Fuels 2007, 21,
1699–1706. [CrossRef]
5. Santos, D.M.F.; Sequeira, C.A.C.; Figueiredo, J.L. Hydrogen Production by Alkaline Water Electrolysis.
Quim. Nova 2013, 36, 1176–1193. [CrossRef]
6. Wang, M.; Wang, Z.; Gong, X.; Guo, Z. The Intensification Technologies to Water Electrolysis for Hydrogen
Production—A Review. Renew. Sustain. Energy Rev. 2014, 29, 573–588. [CrossRef]
7. Gahleitner, G. Hydrogen from Renewable Electricity: An International Review of Power-to-Gas Pilot Plants
for Stationary Applications. Int. J. Hydrog. Energy 2013, 38, 2039–2061. [CrossRef]
8. Carpetis, C. Estimation of Storage Costs for Large Hydrogen Storage Facilities. Int. J. Hydrog. Energy 1982, 7,
191–203. [CrossRef]
9. Michalski, J.; Bünger, U.; Crotogino, F.; Donadei, S.; Schneider, G.S.; Pregger, T.; Cao, K.K.; Heide, D.
Hydrogen Generation by Electrolysis and Storage in Salt Caverns: Potentials, Economics and Systems
Aspects with Regard to the German Energy Transition. Int. J. Hydrog. Energy 2017, 42, 13427–13443.
[CrossRef]
10. Pellow, M.A.; Emmott, C.J.M.; Barnhart, C.J.; Benson, S.M. Hydrogen or Batteries for Grid Storage? A Net
Energy Analysis. Energy Environ. Sci. 2015, 8, 1938–1952. [CrossRef]
11. Qadrdan, M.; Abeysekera, M.; Chaudry, M.; Wu, J.; Jenkins, N. Role of Power-to-Gas in an Integrated Gas
and Electricity System in Great Britain. Int. J. Hydrog. Energy 2015, 40, 5763–5775. [CrossRef]
12. Schouten, J.; Janssenvanrosmalen, R.; Michels, J. Modeling Hydrogen Production for Injection into the
Natural Gas Grid: Balance between Production, Demand and Storage. Int. J. Hydrog. Energy 2006, 31,
1698–1706. [CrossRef]
13. Goncharov, A.; Guglya, A.; Melnikova, E. On the Feasibility of Developing Hydrogen Storages Capable of
Adsorption Hydrogen Both in Its Molecular and Atomic States. Int. J. Hydrog. Energy 2012, 37, 18061–18073.
[CrossRef]
14. Goncharov, A.; Guglya, A.; Melnikova, E. Corrigendum to “On the Feasibility of Developing Hydrogen
Storages Capable of Adsorption Hydrogen Both in Its Molecular and Atomic States” [Int J Hydrogen Energy,
37 (2012) 18061–18073]. Int. J. Hydrog. Energy 2013, 38, 3521. [CrossRef]
15. Schmidt, O.; Gambhir, A.; Staffell, I.; Hawkes, A.; Nelson, J.; Few, S. Future Cost and Performance of Water
Electrolysis: An Expert Elicitation Study. Int. J. Hydrog. Energy 2017, 42, 30470–30492. [CrossRef]
16. Schalenbach, M.; Zeradjanin, A.R.; Kasian, O.; Cherevko, S.; Mayrhofer, K.J. A Perspective on
Low-Temperature Water Electrolysis-Challenges in Alkaline and Acidic Technology. Int. J. Electrochem. Sci.
2018, 13, 1173–1226. [CrossRef]

23
Processes 2020, 8, 248

17. Zeng, K.; Zhang, D. Recent Progress in Alkaline Water Electrolysis for Hydrogen Production and
Applications. Prog. Energy Combust. Sci. 2010, 36, 307–326. [CrossRef]
18. Carmo, M.; Fritz, D.L.; Mergel, J.; Stolten, D. A Comprehensive Review on PEM Water Electrolysis. Int. J.
Hydrog. Energy 2013, 38, 4901–4934. [CrossRef]
19. David, M.; Ocampo-Martínez, C.; Sánchez-Peña, R. Advances in Alkaline Water Electrolyzers: A Review.
J. Energy Storage 2019, 23, 392–403. [CrossRef]
20. Buttler, A.; Spliethoff, H. Current Status of Water Electrolysis for Energy Storage, Grid Balancing and Sector
Coupling via Power-to-Gas and Power-to-Liquids: A Review. Renew. Sustain. Energy Rev. 2018, 82, 2440–2454.
[CrossRef]
21. Rashid, M.; Mesfer, M.K.A.; Naseem, H.; Danish, M. Hydrogen Production by Water Electrolysis: A Review
of Alkaline Water Electrolysis, PEM Water Electrolysis and High Temperature Water Electrolysis. Int. J. Eng.
Adv. Technol. 2015, 4, 80–93.
22. Götz, M.; Lefebvre, J.; Mörs, F.; McDaniel Koch, A.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, T. Renewable
Power-to-Gas: A Technological and Economic Review. Renew. Energy 2016, 85, 1371–1390. [CrossRef]
23. Marini, S.; Salvi, P.; Nelli, P.; Pesenti, R.; Villa, M.; Berrettoni, M.; Zangari, G.; Kiros, Y. Advanced Alkaline
Water Electrolysis. Electrochim. Acta 2012, 82, 384–391. [CrossRef]
24. Seibel, C.; Kuhlmann, J.W. Dynamic Water Electrolysis in Cross-Sectoral Processes. Chem. Ing. Tech. 2018, 90,
1430–1436. [CrossRef]
25. Shen, X.; Zhang, X.; Li, G.; Lie, T.T.; Hong, L. Experimental Study on the External Electrical Thermal and
Dynamic Power Characteristics of Alkaline Water Electrolyzer. Int. J. Energy Res. 2018, 42, 3244–3257.
[CrossRef]
26. Clarivate Analytics. Web of Science Database. 2020. Available online: http://apps.webofknowledge.com
(accessed on 14 January 2020).
27. Ehret, O.; Bonhoff, K. Hydrogen as a Fuel and Energy Storage: Success Factors for the German Energiewende.
Int. J. Hydrog. Energy 2015, 40, 5526–5533. [CrossRef]
28. Ulleberg, O. Modeling of Advanced Alkaline Electrolyzers: A System Simulation Approach. Int. J.
Hydrog. Energy 2003, 28, 21–33. [CrossRef]
29. Ðukić, A.; Firak, M. Hydrogen Production Using Alkaline Electrolyzer and Photovoltaic (PV) Module. Int. J.
Hydrog. Energy 2011, 36, 7799–7806. [CrossRef]
30. LeRoy, R.L. The Thermodynamics of Aqueous Water Electrolysis. J. Electrochem. Soc. 1980, 127, 1954.
[CrossRef]
31. Haug, P.; Koj, M.; Turek, T. Influence of Process Conditions on Gas Purity in Alkaline Water Electrolysis.
Int. J. Hydrog. Energy 2017, 42, 9406–9418. [CrossRef]
32. Haug, P.; Kreitz, B.; Koj, M.; Turek, T. Process Modelling of an Alkaline Water Electrolyzer. Int. J.
Hydrog. Energy 2017, 42, 15689–15707. [CrossRef]
33. Renaud, R.; Leroy, R. Separator Materials for Use in Alkaline Water Electrolysers. Int. J. Hydrog. Energy 1982,
7, 155–166. [CrossRef]
34. Kraglund, M.R.; Aili, D.; Jankova, K.; Christensen, E.; Li, Q.; Jensen, J.O. Zero-Gap Alkaline Water Electrolysis
Using Ion-Solvating Polymer Electrolyte Membranes at Reduced KOH Concentrations. J. Electrochem. Soc.
2016, 163, F3125–F3131. [CrossRef]
35. Kraglund, M.R.; Carmo, M.; Schiller, G.; Ansar, S.A.; Aili, D.; Christensen, E.; Jensen, J.O. Ion-Solvating
Membranes as a New Approach towards High Rate Alkaline Electrolyzers. Energy Environ. Sci. 2019,
12, 3313–3318. [CrossRef]
36. Hnát, J.; Paidar, M.; Schauer, J.; Žitka, J.; Bouzek, K. Polymer Anion-Selective Membranes for Electrolytic
Splitting of Water. Part II: Enhancement of Ionic Conductivity and Performance under Conditions of Alkaline
Water Electrolysis. J. Appl. Electrochem. 2012, 42, 545–554. [CrossRef]
37. Hnát, J.; Plevová, M.; Žitka, J.; Paidar, M.; Bouzek, K. Anion-Selective Materials with
1,4-Diazabicyclo[2.2.2]Octane Functional Groups for Advanced Alkaline Water Electrolysis. Electrochim. Acta
2017, 248, 547–555. [CrossRef]
38. Phillips, R.; Dunnill, C.W. Zero Gap Alkaline Electrolysis Cell Design for Renewable Energy Storage as
Hydrogen Gas. RSC Adv. 2016, 6, 100643–100651. [CrossRef]

24
Processes 2020, 8, 248

39. Sánchez, M.; Amores, E.; Rodríguez, L.; Clemente-Jul, C. Semi-Empirical Model and Experimental Validation
for the Performance Evaluation of a 15 kW Alkaline Water Electrolyzer. Int. J. Hydrog. Energy 2018,
43, 20332–20345. [CrossRef]
40. Hammoudi, M.; Henao, C.; Agbossou, K.; Dubé, Y.; Doumbia, M. New Multi-Physics Approach for Modelling
and Design of Alkaline Electrolyzers. Int. J. Hydrog. Energy 2012, 37, 13895–13913. [CrossRef]
41. Koj, M.; Gimpel, T.; Schade, W.; Turek, T. Laser Structured Nickel-Iron Electrodes for Oxygen Evolution in
Alkaline Water Electrolysis. Int. J. Hydrog. Energy 2019, 44, 12671–12684. [CrossRef]
42. Koj, M.; Qian, J.; Turek, T. Novel Alkaline Water Electrolysis with Nickel-Iron Gas Diffusion Electrode for
Oxygen Evolution. Int. J. Hydrog. Energy 2019, 44, 29862–29875. [CrossRef]
43. Hall, D.E. Electrodes for Alkaline Water Electrolysis. J. Electrochem. Soc. 1981, 128, 740. [CrossRef]
44. Huot, J.Y. Low Hydrogen Overpotential Nanocrystalline Ni-Mo Cathodes for Alkaline Water Electrolysis.
J. Electrochem. Soc. 1991, 138, 1316. [CrossRef]
45. Rauscher, T.; Bernäcker, C.I.; Mühle, U.; Kieback, B.; Röntzsch, L. The Effect of Fe as Constituent in Ni-Base
Alloys on the Oxygen Evolution Reaction in Alkaline Solutions at High Current Densities. Int. J. Hydrog.
Energy 2019, 44, 6392–6402. [CrossRef]
46. Fan, C. Study of Electrodeposited Nickel-Molybdenum, Nickel-Tungsten, Cobalt-Molybdenum,
and Cobalt-Tungsten as Hydrogen Electrodes in Alkaline Water Electrolysis. J. Electrochem. Soc. 1994,
141, 382. [CrossRef]
47. Rauscher, T.; Müller, C.I.; Schmidt, A.; Kieback, B.; Röntzsch, L. Ni–Mo–B Alloys as Cathode Material for
Alkaline Water Electrolysis. Int. J. Hydrog. Energy 2016, 41, 2165–2176. [CrossRef]
48. Ursúa, A.; San Martín, I.; Barrios, E.L.; Sanchis, P. Stand-Alone Operation of an Alkaline Water Electrolyser
Fed by Wind and Photovoltaic Systems. Int. J. Hydrog. Energy 2013, 38, 14952–14967. [CrossRef]
49. Henao, C.; Agbossou, K.; Hammoudi, M.; Dubé, Y.; Cardenas, A. Simulation Tool Based on a Physics Model
and an Electrical Analogy for an Alkaline Electrolyser. J. Power Sources 2014, 250, 58–67. [CrossRef]
50. Balabel, A.; Zaky, M.S.; Sakr, I. Optimum Operating Conditions for Alkaline Water Electrolysis Coupled
with Solar PV Energy System. Arab. J. Sci. Eng. 2014, 39, 4211–4220. [CrossRef]
51. Vermeiren, P. Zirfon® : A New Separator for Ni-H2 Batteries and Alkaline Fuel Cells. Int. J. Hydrog. Energy
1996, 21, 679–684. [CrossRef]
52. Le Bideau, D.; Mandin, P.; Benbouzid, M.; Kim, M.; Sellier, M. Review of Necessary Thermophysical
Properties and Their Sensivities with Temperature and Electrolyte Mass Fractions for Alkaline Water
Electrolysis Multiphysics Modelling. Int. J. Hydrog. Energy 2019, 44, 4553–4569. [CrossRef]
53. Shoor, S.K.; Walker, R.D.; Gubbins, K.E. Salting out of Nonpolar Gases in Aqueous Potassium Hydroxide
Solutions. J. Phys. Chem. 1969, 73, 312–317. [CrossRef]
54. Grover, P.K.; Ryall, R.L. Critical Appraisal of Salting-Out and Its Implications for Chemical and Biological
Sciences. Chem. Rev. 2005, 105, 1–10. [CrossRef] [PubMed]
55. Randall, M.; Failey, C.F. The Activity Coefficient of Non-Electrolytes in Aqueous Salt Solutions from Solubility
Measurements. The Salting-out Order of the Ions. Chem. Rev. 1927, 4, 285–290. [CrossRef]
56. Lang, W.; Zander, R. Salting-out of Oxygen from Aqueous Electrolyte Solutions: Prediction and Measurement.
Ind. Eng. Chem. Fundam. 1986, 25, 775–782. [CrossRef]
57. See, D.M.; White, R.E. Temperature and Concentration Dependence of the Specific Conductivity of
Concentrated Solutions of Potassium Hydroxide. J. Chem. Eng. Data 1997, 42, 1266–1268. [CrossRef]
58. de Souza, R.F.; Padilha, J.C.; Gonçalves, R.S.; Rault-Berthelot, J. Dialkylimidazolium Ionic Liquids as
Electrolytes for Hydrogen Production from Water Electrolysis. Electrochem. Commun. 2006, 8, 211–216.
[CrossRef]
59. Zhao, Y.; Zhao, J.; Huang, Y.; Zhou, Q.; Zhang, X.; Zhang, S. Toxicity of Ionic Liquids: Database and
Prediction via Quantitative Structure–Activity Relationship Method. J. Hazard. Mater. 2014, 278, 320–329.
[CrossRef]
60. Zhao, Y.; Gani, R.; Afzal, R.M.; Zhang, X.; Zhang, S. Ionic Liquids for Absorption and Separation of Gases:
An Extensive Database and a Systematic Screening Method. AIChE J. 2017, 63, 1353–1367. [CrossRef]
61. Zhao, Y.; Pan, M.; Kang, X.; Tu, W.; Gao, H.; Zhang, X. Gas Separation by Ionic Liquids: A Theoretical Study.
Chem. Eng. Sci. 2018, 189, 43–55. [CrossRef]

25
Processes 2020, 8, 248

62. De Souza, R.F.; Padilha, J.C.; Gonçalves, R.S.; de Souza, M.O.; Rault-Berthelot, J. Electrochemical
Hydrogen Production from Water Electrolysis Using Ionic Liquid as Electrolytes: Towards the Best Device.
J. Power Sources 2007, 164, 792–798. [CrossRef]
63. Schalenbach, M.; Lueke, W.; Stolten, D. Hydrogen Diffusivity and Electrolyte Permeability of the Zirfon
PERL Separator for Alkaline Water Electrolysis. J. Electrochem. Soc. 2016, 163, F1480–F1488. [CrossRef]
64. Trinke, P.; Haug, P.; Brauns, J.; Bensmann, B.; Hanke-Rauschenbach, R.; Turek, T. Hydrogen Crossover
in PEM and Alkaline Water Electrolysis: Mechanisms, Direct Comparison and Mitigation Strategies.
J. Electrochem. Soc. 2018, 165, F502–F513. [CrossRef]
65. Roy, A.; Watson, S.; Infield, D. Comparison of Electrical Energy Efficiency of Atmospheric and High-Pressure
Electrolysers. Int. J. Hydrog. Energy 2006, 31, 1964–1979. [CrossRef]
66. Ursúa, A.; Sanchis, P. Static–Dynamic Modelling of the Electrical Behaviour of a Commercial Advanced
Alkaline Water Electrolyser. Int. J. Hydrog. Energy 2012, 37, 18598–18614. [CrossRef]
67. Dobó, Z.; Palotás, Á.B. Impact of the Voltage Fluctuation of the Power Supply on the Efficiency of Alkaline
Water Electrolysis. Int. J. Hydrog. Energy 2016, 41, 11849–11856. [CrossRef]
68. Dobó, Z.; Palotás, Á.B. Impact of the Current Fluctuation on the Efficiency of Alkaline Water Electrolysis.
Int. J. Hydrog. Energy 2017, 42, 5649–5656. [CrossRef]
69. Speckmann, F.W.; Bintz, S.; Birke, K.P. Influence of Rectifiers on the Energy Demand and Gas Quality of
Alkaline Electrolysis Systems in Dynamic Operation. Appl. Energy 2019, 250, 855–863. [CrossRef]
70. De Fátima Palhares, D.D.; Vieira, L.G.M.; Damasceno, J.J.R. Hydrogen Production by a Low-Cost Electrolyzer
Developed through the Combination of Alkaline Water Electrolysis and Solar Energy Use. Int. J. Hydrog.
Energy 2018, 43, 4265–4275. [CrossRef]
71. Hug, W.; Bussmann, H.; Brinner, A. Intermittent Operation and Operation Modeling of an Alkaline
Electrolyzer. Int. J. Hydrog. Energy 1993, 18, 973–977. [CrossRef]
72. Firak, M.; Djukic, A. An Investigation into the Effect of Photovoltaic Module Electric Properties on Maximum
Power Point Trajectory with the Aim of Its Alignment with Electrolyzer U-I Characteristic. Therm. Sci. 2010,
14, 729–738. [CrossRef]
73. Ðukić, A. Autonomous Hydrogen Production System. Int. J. Hydrog. Energy 2015, 40, 7465–7474. [CrossRef]
74. Dai, J.; Liu, D.; Wen, L.; Long, X. Research on Power Coefficient of Wind Turbines Based on SCADA Data.
Renew. Energy 2016, 86, 206–215. [CrossRef]
75. Chin, V.J.; Salam, Z.; Ishaque, K. Cell Modelling and Model Parameters Estimation Techniques for
Photovoltaic Simulator Application: A Review. Appl. Energy 2015, 154, 500–519. [CrossRef]
76. Vergura, S. A Complete and Simplified Datasheet-Based Model of PV Cells in Variable Environmental
Conditions for Circuit Simulation. Energies 2016, 9, 326. [CrossRef]
77. Kovač, A.; Marciuš, D.; Budin, L. Solar Hydrogen Production via Alkaline Water Electrolysis. Int. J. Hydrog.
Energy 2019, 44, 9841–9848. [CrossRef]
78. Bhattacharyya, R.; Misra, A.; Sandeep, K. Photovoltaic Solar Energy Conversion for Hydrogen Production
by Alkaline Water Electrolysis: Conceptual Design and Analysis. Energy Convers. Manag. 2017, 133, 1–13.
[CrossRef]
79. Khalilnejad, A.; Riahy, G. A Hybrid Wind-PV System Performance Investigation for the Purpose of Maximum
Hydrogen Production and Storage Using Advanced Alkaline Electrolyzer. Energy Convers. Manag. 2014,
80, 398–406. [CrossRef]
80. Badwal, S.P.S.; Giddey, S.S.; Munnings, C.; Bhatt, A.I.; Hollenkamp, A.F. Emerging Electrochemical Energy
Conversion and Storage Technologies. Front. Chem. 2014, 2. [CrossRef]
81. Kolli, A.; Gaillard, A.; De Bernardinis, A.; Bethoux, O.; Hissel, D.; Khatir, Z. A Review on DC/DC Converter
Architectures for Power Fuel Cell Applications. Energy Convers. Manag. 2015, 105, 716–730. [CrossRef]
82. Ursúa, A.; Barrios, E.L.; Pascual, J.; San Martín, I.; Sanchis, P. Integration of Commercial Alkaline Water
Electrolysers with Renewable Energies: Limitations and Improvements. Int. J. Hydrog. Energy 2016,
41, 12852–12861. [CrossRef]
83. Zini, G.; Tartarini, P. Wind-Hydrogen Energy Stand-Alone System with Carbon Storage: Modeling and
Simulation. Renew. Energy 2010, 35, 2461–2467. [CrossRef]
84. Akpinar, E.K.; Akpinar, S. An Assessment on Seasonal Analysis of Wind Energy Characteristics and Wind
Turbine Characteristics. Energy Convers. Manag. 2005, 46, 1848–1867. [CrossRef]

26
Processes 2020, 8, 248

85. Babu, N.R.; Arulmozhivarman, P. Wind Energy Conversion Systems-A Technical Review. J. Eng. Sci. Technol.
2013, 8, 493–507.
86. Douglas, T.G.; Cruden, A.; Infield, D. Development of an Ambient Temperature Alkaline Electrolyser for
Dynamic Operation with Renewable Energy Sources. Int. J. Hydrog. Energy 2013, 38, 723–739. [CrossRef]
87. Kiaee, M.; Cruden, A.; Infield, D.; Chladek, P. Utilisation of Alkaline Electrolysers to Improve Power System
Frequency Stability with a High Penetration of Wind Power. IET Renew. Power Gener. 2014, 8, 529–536.
[CrossRef]
88. Varone, A.; Ferrari, M. Power to Liquid and Power to Gas: An Option for the German Energiewende.
Renew. Sustain. Energy Rev. 2015, 45, 207–218. [CrossRef]
89. Parra, D.; Swierczynski, M.; Stroe, D.I.; Norman, S.; Abdon, A.; Worlitschek, J.; O’Doherty, T.; Rodrigues, L.;
Gillott, M.; Zhang, X.; et al. An Interdisciplinary Review of Energy Storage for Communities: Challenges
and Perspectives. Renew. Sustain. Energy Rev. 2017, 79, 730–749. [CrossRef]
90. Kiaee, M.; Infield, D.; Cruden, A. Utilisation of Alkaline Electrolysers in Existing Distribution Networks to
Increase the Amount of Integrated Wind Capacity. J. Energy Storage 2018, 16, 8–20. [CrossRef]
91. Hug, W.; Divisek, J.; Mergel, J.; Seeger, W.; Steeb, H. Highly Efficient Advanced Alkaline Electrolyzer for
Solar Operation. Int. J. Hydrog. Energy 1992, 17, 699–705. [CrossRef]
92. Kuroda, Y.; Nishimoto, T.; Mitsushima, S. Self-Repairing Hybrid Nanosheet Anode Catalysts for Alkaline
Water Electrolysis Connected with Fluctuating Renewable Energy. Electrochim. Acta 2019, 323, 134812.
[CrossRef]
93. Djafour, A.; Matoug, M.; Bouras, H.; Bouchekima, B.; Aida, M.; Azoui, B. Photovoltaic-Assisted Alkaline
Water Electrolysis: Basic Principles. Int. J. Hydrog. Energy 2011, 36, 4117–4124. [CrossRef]
94. Artuso, P.; Zuccari, F.; Dell’Era, A.; Orecchini, F. PV-Electrolyzer Plant: Models and Optimization Procedure.
J. Sol. Energy Eng. 2010, 132, 031016. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

27
processes
Article
Pore Network Simulation of Gas-Liquid Distribution
in Porous Transport Layers
Nicole Vorhauer 1, *, Haashir Altaf 1 , Evangelos Tsotsas 1 and Tanja Vidakovic-Koch 2
1 Institute of Process Engineering, Otto von Guericke University Magdeburg, 39106 Magdeburg, Germany
2 Max-Planck-Institute for Dynamics of Complex Technical Systems Magdeburg, 39106 Magdeburg, Germany
* Correspondence: [email protected]; Tel.: +48-391-67-51684

Received: 13 July 2019; Accepted: 20 August 2019; Published: 23 August 2019

Abstract: Pore network models are powerful tools to simulate invasion and transport processes in
porous media. They are widely applied in the field of geology and the drying of porous media, and
have recently also received attention in fuel cell applications. Here we want to describe and discuss
how pore network models can be used as a prescriptive tool for future water electrolysis technologies.
In detail, we suggest in a first approach a pore network model of drainage for the prediction of the
oxygen and water invasion process inside the anodic porous transport layer at high current densities.
We neglect wetting liquid films and show that, in this situation, numerous isolated liquid clusters
develop when oxygen invades the pore network. In the simulation with narrow pore size distribution,
the volumetric ratio of the liquid transporting clusters connected between the catalyst layer and the
water supply channel is only around 3% of the total liquid volume contained inside the pore network
at the moment when the water supply route through the pore network is interrupted; whereas around
40% of the volume is occupied by the continuous gas phase. The majority of liquid clusters are
disconnected from the water supply routes through the pore network if liquid films along the walls
of the porous transport layer are disregarded. Moreover, these clusters hinder the countercurrent
oxygen transport. A higher ratio of liquid transporting clusters was obtained for greater pore size
distribution. Based on the results of pore network drainage simulations, we sketch a new route for
the extraction of transport parameters from Monte Carlo simulations, incorporating pore scale flow
computations and Darcy flow.

Keywords: pore network model; Monte Carlo simulation; drainage invasion; porous transport layer;
clustering effect; water electrolysis

1. Introduction
Pore network models (PNMs) are discrete mathematical models that are basically used to simulate
and predict pore scale processes. Different types of PNMs are generally distinguished. These are
(i) PNMs for quasi-static drainage invasion processes [1–4], (ii) PNMs for quasi-static imbibition
invasion processes [1,3–5], (iii) PNMs of drainage with phase transition and diffusion of the vapor
(especially applied in drying research) [6–8], and (iv) PNMs for the computation of dynamic pore scale
fluid flow [9–13]. While the first three approaches usually assume quasi-static invasion of the pore
space, the fourth approach considers viscous flow of the liquid phase and dynamic invasion of the
pore space. There are several examples that combine (i) and (ii) [3,4,14]. There also exists a number
of PNMs that additionally incorporate pore scale mechanisms, such as liquid film flow [15–17] or
crystallization [18–20]. Only a few models are available that take into account coupled heat and mass
transfer [21,22] or that at least consider the invasion and transport processes under non-isothermal
conditions [23–25].
A wide range of PNM applications related to fuel cells is already available in literature, e.g., [26–33].
Major focus of the PN studies related to fuel cells is essentially on liquid water distribution in

Processes 2019, 7, 558; doi:10.3390/pr7090558 29 www.mdpi.com/journal/processes


Processes 2019, 7, 558

dependence of the pore structure [28], liquid clustering [29], the thickness of the gas diffusion layer [30],
hydrophobicity [27], local variation of wetting properties, condensation-induced liquid water formation
inside the pore structure [31,33], ice formation, model validation by X-ray tomography [32], and
neutron tomography [34,35].
So far, PNMs are only rarely applied in water electrolysis research. Experimental studies
incorporating microfluidic platforms of pore networks (PNs) are presented by Bazylak et al. [36–39].
Their investigations are based on the correlation of the oxygen flow rate with current density.
The microfluidic platforms employed in their studies are based on the geometric information obtained
from micro tomography measurements of different titanium porous transport layers (PTLs). They
showed that the developing gas bubbles penetrate the porous structure of the model PTLs and form
continuous fractal gas branches that cover the PN at the breakthrough point. The network covering gas
fingers at the breakthrough point are assumed as stable as long as the volume flow rates correlating with
the current density are constant (In [38,39], the air flow rates were 1, 5, and 10 μL min−1 with current
densities of 1.4, 7.0, and 14 A cm−2 , and the liquid flow rates were 5 and 10 μL min−1 . The same liquid
flow rates and air flow rate of 1 μL min−1 were applied in the numerical simulations in [37]. In [36],
higher liquid flow rates of 505 μL min−1 and gas flow rates of 10 to 300 μL min−1 were applied.) It was
illustrated that the shape of the penetrating gas fingers and the final saturation of the microfluidic PN
with oxygen are dictated by the pore structure when the invasion occurs in a capillary regime. Higher
gas saturation values were obtained for lower porosity of the PN and smaller pore and throat sizes
(sintered PTL in [39]). However, an explanation for this outcome is not given in the paper. Additionally,
only 2-dimensional (2D) microfluidic experiments are presented; as will be discussed in this paper,
the relationship between pore size distribution and saturation is different in 3-dimensional (3D) PNs
where essentially the interconnectivity of the pore space is higher. However, it is surmised from PN
modeling that the pore size distribution (PSD), especially the standard deviation of pore and throat
sizes, were different in the three cases studied in [39] (felt, sintered, and foam PTL). The presence of
large and small pores results in the competitive invasion of the PTL. This can especially be illustrated
for strongly heterogeneous pore structures, e.g., bimodal PSD, [40]. Principally, the invasion process
follows the path of least resistance. In the hydrophilic micromodels used in experiments in [39], this is
the path that follows the lowest invasion pressure thresholds of the neighboring pores; it is thus the
route of the largest neighbor pores. In more detail, invasion of the PTL at constant current densities is
a process of quasi-static drainage of water. This process occurs in distinct steps. This is designated as
the mechanism ‘one throat at a time’ in [36] and can be simulated with PNMs. To invade any liquid
filled pore or throat with water, the pressure inside the gas phase has to overcome a critical invasion
pressure threshold that depends on the curvature of the gas-liquid interface. If the pore and throat
sizes are distinct, the invasion events occur at distinct pressures. According to this, the pressure curve
follows a trend of alternating pressure increase (to achieve the critical invasion pressure thresholds),
during which the saturation of the microfluidic network remains constant, and the pressure drops
during the sudden invasion (saturation decrease) when the critical invasion pressure is achieved [41].
In [36], it is also shown that the invasion of the PN can continue after the breakthrough of one gas
branch if the oxygen flow rate, and thus the gas pressures are increased accordingly.
In this paper we focus on water electrolysis cells with in the sandwich coordinate countercurrent
flow of water and oxygen (perpendicular to the catalyst layer and the proton exchange membrane) [42].
Additionally, we consider the situation of high oxygen production rates (current density =
0.6–6 A/cm2 [43]; oxygen flux < 4 μL s−1 mm−2 ). Based on this, we assume plug flow of oxygen and
water. In the first approach presented in this paper, we use pore network Monte Carlo simulations
(PNMCSs) for the prediction of the invasion of an initially fully water saturated PN with oxygen. We
furthermore assume that a stationary distribution of oxygen and water inside the PTL is achieved when
the oxygen flow paths connect the catalyst layer, where the oxygen is produced, with the water supply
channel, while the water connectivity between both sides is maintained, presuming constant oxygen
production rates and constant pressure and temperature along the PTL. As a result of the PNMCSs, we

30
Processes 2019, 7, 558

show and discuss the invasion patterns in the moment when the water supply route is disconnected,
as well as the impact of isolated single liquid clusters on the relative permeability of oxygen and water.
Following parameter estimation concepts, e.g., in drainage of soils [1] or drying [44,45], we propose to
estimate relative permeability from the stationary final gas-liquid distribution obtained from the PN
drainage simulation.

2. Pore Network Model


A schematic sketch of the PNM is given in Figure 1. Here we consider idealized 3D PNs with
PSDs in the range of typical PTLs usually applied for water electrolysis (Figure 2). The application of
idealized lattice structures is commonly practiced if the physical mechanisms and pore scale effects are
analyzed. More advanced studies base the PN simulations on the real structure of the porous medium
(e.g., [36,46]). This is also foreseen in our studies in the next step. However, in this paper we present
and discuss the basics of the method and the basic outcomes of the drainage simulation and thus
disregard the more expensive (concerning time and computational effort) method.

(a) (b)

Figure 1. (a) Schematic illustration of the pore network model (PNM) of drainage with pore and throat
numbering (pore numbers in blue; throat numbers in black). (b) Breakthrough path of oxygen from the
catalyst side to the water supply channel. Note the periodic boundary conditions that allow connection
of edge throats and pores with each other.

As can be seen in Figure 1a, the PN consists of circular pores, i.e., the sites and cylindrical throats,
i.e., the bonds. Both pores and throats can in principal be occupied by either liquid water or gaseous
oxygen. The liquid occupied pores are shown in black in Figure 1a; the gas occupied pores are shown
in red. Similarly, the liquid occupied throats are shown in blue and the empty throats are shown in red.
The pores at the bottom side of the network can be interpreted as water sources as they represent the
connection links to the water supply channel. All of these pores are initially saturated with water (black
pores in Figure 1). In contrast, pores at the top side of the PN are connected to the oxygen sources inside
the catalyst layer at the anode side of the water electrolysis cell and directly supplied with gaseous
oxygen (red pores in Figure 1). The water supply route from the water supply channel connected to the
bottom of the PN towards the catalyst layer (at the top side) is given by the interconnected blue throats,
while the red pores and throats provide distinct routes for the gas phase. Figure 1b shows the path of
oxygen through the PN. As can be seen, the oxygen path covers the PN from the catalyst side to the
water supply channel in this example. This situation is called a breakthrough. The breakthrough of the
gas phase is achieved when one of the bottom pores is occupied with oxygen. On the other hand, the

31
Processes 2019, 7, 558

water supply is interrupted once the blue cluster is either completely split up into numerous smaller
single clusters or disconnected from either side.
The geometric arrangement of pores and throats and the neighbor relations are specified by the
different matrices pnp, tnt, tnp, and pnt. Following the example in Figure 1a, the neighbor relations of
the 3D PN as illustrated here are:
⎡ .. .. .. .. .. .. ⎤
⎢⎢ ⎥⎥
⎢⎢ . . . . . . ⎥⎥
⎢ ⎥⎥
pnp(pore 10) = ⎢⎢⎢⎢ 1 11 12 13 16 19 ⎥⎥
⎥⎥ (1)
⎢⎢ ⎥⎦
⎣ .. .. .. .. .. ..
. . . . . .
⎡ ⎤
⎢⎢ 10 11 ⎥⎥
tnp(throat 1) = ⎢⎢⎢⎣ . ..
⎥⎥
⎥⎦ (2)
.. .
⎡ . . .. .. .. .. ⎤
⎢⎢ .. .. ⎥⎥
⎢⎢⎢ . . . . ⎥⎥
⎥⎥

pnt(pore 10) = ⎢⎢⎢ 1 3 28 34 55 64 ⎥⎥ (3)
⎢⎢ . . ⎥⎥
⎣ . . .. .. .. .. ⎥⎦
. . . . . .
⎡ ⎤
⎢⎢ 2 3 28 29 34 35 55 56 64 65 ⎥⎥
tnt(throat 1) = ⎢⎢⎢⎣ . . . .. .. .. .. .. .. ..
⎥⎥
⎥⎦ (4)
.. .. .. . . . . . . .
From this, it follows that the coordination number of the PN is 6. Note that we apply periodic
boundary conditions. This means that the pores at all lateral sides are connected to each other in order
to reduce confinement effects. While the radius distribution of the throats and pores stochastically
vary in the PNMCSs, the other geometrical parameter, such as throat length and the neighbor relations,
are kept constant (Table 1). The PSDs of pores and throats in the PN simulations presented below are
illustrated in Figure 2.

Figure 2. Representative pore size distribution (PSD) of the pore networks (PNs) studied in the pore
network Monte Carlo simulations (PNMCSs) in Section 3.1. The PN with low standard deviation (STD)
of pore and throat sizes is denoted STD 0.5 μm, and the PN with a higher standard variation is denoted
STD 1 μm (refer to discussions below in Section 3.1). The first peaks correspond to the sizes of throats
and the second peaks to the pore sizes, respectively. Note that the overlap of pore and throat sizes is
greater for a higher standard deviation. The porosity of both PNs is 21%.

32
Processes 2019, 7, 558

Table 1. Simulation parameters.

Parameter Value
Network size 30 × 30 × 10
Pore number 9000
Throat number 22,500
PTL temperature T 50 ◦ C
Cell pressure P 10 bar
Contact angle θ 0◦
Throat length Lt 27 μm
Lattice spacing L 50 μm
Thickness of the PN 450 μm
Porosity 21%

We assume that the oxygen is homogeneously distributed along the surface of the PN and neglect
spatial and temporal pressure fluctuations. Instead, we assume a constant oxygen supply rate. Based
on this, we anticipate plug flow of oxygen and water and compute the invasion of the PN based on the
Young–Laplace equation, following the concepts in [8,47,48]:

2σ cos θ
Pl,t,p = P − (5)
rt,p

The order of invasion thus follows the radius distribution of the throats (with index t) and pores
(with index p). Viscous flow is disregarded for the drainage invasion computation. The liquid clusters
are labeled based on the throat saturation, taking into account the saturation in neighboring pores.
This means that any neighbor throats that contain water belong to the same liquid cluster if the pore
between them is also saturated with water (Figure 1). The gas phase is not labeled as it is continuous.
Note that liquid clusters that are disconnected from the cluster spanning the PTL from the catalyst side
to the water channel side (all blue throats in Figure 1a) are not further invaded. These clusters remain
in their original size and can be regarded as transport barriers for oxygen flow (Figure 3). In regard to
an efficient PTL mass transfer, it would be affordable to avoid clustering of the liquid phase; relevant
issues are also intensively studied e.g., in hydrodynamics of porous geological structures and soils [49].
However, in regard to the optimization of the PTL, several aspects interact with each other, including
mass transfer, heat transfer, and electrical conductivity [50].
The PNMCSs presented here are restricted to the computation of the point when the fragmented
clusters are not further invaded. In most cases, the fragmented clusters do not span the network as
they are either connected to only one of its open sides (top or bottom) or isolated in the center of the
PN. The simulation can result in several gas branches penetrating the network from top to bottom.
(Note that the gas phase always forms a continuum). This is in good agreement with experimental
findings reported in [36].

33
Processes 2019, 7, 558

continuous liquid cluster

isolated clusters

(a) (b)

Figure 3. Liquid clustering in a 3-dimensional (3D) PN on the example of a small 10 × 10 × 10 network


with homogeneous PSD and the parameters specified in Table 1. The transport barrier clusters (isolated
and discrete) are shown in grey (a,b); the liquid transporting clusters are shown in blue (a,b); and the
continuous gas phase is shown in magenta (a). Pores are not shown for reasons of readability.

The computation of the quasi-static invasion profile does not incorporate the solution of mass
transfer equations because invasion percolation with trapping is assumed here. Instead, the pore
scale fluid transport equations are set-up and solved in the second step based on the stationary
invasion patterns from step 1 (Figure 4). For the computation in step 2, the breakthrough invasion
patterns, i.e., at disconnection of water transport routes, are used [29,30,37–39]. Only the liquid cluster
spanning the PN from top to bottom in the moment before disconnection and the continuous gas
phase are considered. The mass transfer through the spanning clusters and the gas phase is computed
pore-to-pore, based on the Hagen–Poiseuille equation:

. ρl,g πrt 4 
Ml,g = Pl,g,1 − Pl,g,2 , (6)
8ηl,g L

for the liquid phase l and the gas phase g. Incompressible, Newtonian viscous flow is assumed
(the compression factor of oxygen at operation conditions of P = 10 bar and T = 50 ◦ C is roughly 1).
Note that the assumption of the Hagen–Poiseuille flow is a strong model simplification for the pore
flow because L  d. A more advanced approach would account for the radial velocity of the flow [51].
In Equation (6), Pl,1 and Pg,1 are the liquid and vapor pressures in neighbor pore 1 of a throat (compare
with Equation (2)) and Pl,2 and Pg,2 are the liquid and vapor pressures in neighbor pore 2 of that throat,
respectively. The resulting set of linear equations is then transferred into the matrix notation:

Pl,g = Al,g gl,g \bl,g (7)

In these equations, Al and Ag represent the matrices of liquid and vapor conductivities of the
throats (e.g., refer to [52] for further details), with

ρl,g πrt 4
gl,g = . (8)
8ηl,g L

34
Processes 2019, 7, 558

Following discussions in [36], conductivities inside pores are not computed as the pores are
not interpreted as hydraulic conductors. Validation of this assumption could be further studied by
Lattice–Boltzmann simulation, e.g., [53,54].
The boundary conditions for each throat are given in bl and bg , which are specified for the pore
neighbors 1 and 2 of a throat. Generally:

bl,g = gl,g Pl,g (9)

The boundary conditions for the PN simulation are P1 at the bottom side of the PN and P2 at the
top side (Figure 4). Solving Equation (7) yields the vapor and liquid pressure fields in the PN. With
this, Equation (6) can be solved for each throat. Once the liquid and vapor flow rates through the
liquid throats in spanning liquid clusters and the gas throats are known, the overall mass flow rates
are computed in step 3 (Figure 4c). From this follows the relative permeability for either the liquid (l)
or gas phase (g):
.
ηl,g Ml,g Δz
K · kl,g = . (10)
ρl,g A ΔPl,g
The absolute permeability K is obtained for the same computations but a totally empty
(gas permeability) or totally saturated (liquid permeability) network.

 
(a) (b) (c)

Figure 4. Extraction of efficient transport parameters based on PNMCSs. (a) Step 1: Computation of
the steady state invasion patterns at the disconnection of the water supply route by PNMCS. (b) Step 2:
Computation of the pore scale fluid flow based on the patterns obtained from step 1. (c) From the
flow rates in step 2, the relative and absolute permeabilities of the PN are computed in step 3 on the
Darcy scale.

Note that high pressures up to P = 30 bar are postulated as usual operating conditions of water
electrolysis cells. It is remarked that the evaporation of water, even at prevailing temperature of
50–80 ◦ C, can be disregarded at such high pressures. Additionally, in the simulations presented here,
we generally assumed hydrophilic conditions with cosθ = 1 and constant temperature and pressure
as well.

3. PN Simulations and Results

3.1. 3D PNMCS of Drainage


We present here the simulation of one realization of a 3D PN with square lattice and 30 × 30 × 10
pores (Figure 5). The simulation parameters are summarized in Table 1. The overall pressure, as well
as the temperature, were kept constant. The lattice spacing between the pores was L = 50 μm and the
throat length was kept constant with Lt = 27 μm. The pore and throat sizes varied in the range given in

35
Processes 2019, 7, 558

Figure 2, with standard deviations 0.5 μm and 1 μm. The simulations were repeated 20 times with
randomly distributed pore and throat radii within the range given in Figure 2.
According to [55] or [36,46], porosity of the PTL usually varies between 54% and 85%, depending
on the kind of the material. Usually, sintered powder has a lower porosity than felt PTLs or foam
PTLs. However, the porosity in our simulations was only around 21%. Larger porosities would only be
achieved in our PN with greater variation of the pore and throat sizes (i.e., by higher standard deviation)
and much shorter throat lengths. To reach the high porosities given in the literature, overlapping of the
pore volumes and negligible throat length would have been necessary. This would lead to different
invasion effects than currently underlying in the proposed PNM, namely site invasion instead of bond
invasion and different flow regimes than those postulated above. (Exemplarily, the porosity could be
increased to 43% for throat lengths of 7 μm. Lt < rt would then require a revision for the assumption of
a developed Poiseuille flow inside the throats). Instead, we expect to achieve higher porosities in PNs
based on the real porous structure, with various heterogeneities of the pore space (see discussions in
Section 4 below). Additionally, due to the relatively long throat length applied in our PN, the thickness
is greater than the given values in literature [36,55] where the 10 pore rows correspond to a thickness
between 170 μm and 300 μm. Independent of this, the PN simulations presented below nicely illustrate
the invasion mechanisms that drive the drainage process. The presented method can be easily adapted
to the simulation of the real porous structure.

Figure 5. 3D PN under study. Pores are shown in black and throats in blue. The top and bottom pore
rows are only vertically connected to their neighbor throats. All pores and throats are initially saturated
with liquid.

In these simulations, pores were invaded independently of their throat neighbors. Note that the
drainage process is principally a bond invasion process wherefore pores could be invaded together
with the throat neighbors [4]. This is also revealed by the experiments presented in [36]. However,
as shown in Figure 2 and also represented in Figure 6, the pores are comparably large in our PN
with invasion pressure thresholds in the range of the throats. The curves of pore and throat capillary
pressures partly overlap and thus indicate the competitive invasion. Apart from that, the invasion
pressure thresholds vary linearly with the size of pores and throats if all other parameter are kept
constant (Equation (5), Table 1). Figure 6 indicates that the capillary pressures are rather small in the
given range of pore sizes, which results in the decrease of liquid pressure (against the gas pressure) of
only a few millibars and thus a concave gas-liquid interface.

36
Processes 2019, 7, 558

Figure 6. Capillary pressure variation in the studied range of pore sizes (standard deviation of 0.5 μm,
Figure 2). Due to the overlapping of the curves the pore invasion is computed independently of the
throat invasion.

The simulation results of one PN simulation (with standard deviation 0.5 μm, Figure 2) are
summarized below and in Figures 7–9. At first, the capillary pressure curves of the PN simulation are
shown in Figure 7. In the hydrophilic drainage process studied here, the available meniscus pore or
throat with the lowest capillary pressure is first invaded. Since all surface pores are available with
liquid menisci at the start of the simulation, the blue curve initially follows the radius distribution of
the surface pores. Once all surface pores are invaded, the capillary curve passes into the trend of the
throats, whereas the pore invasion becomes random depending on the distribution of the available
meniscus pores. Note that more and more pores become available for invasion when the drainage front
widens. However, the blue cloud only represents the pores in liquid clusters spanning the PN, since all
other clusters are stationary and not invaded. This holds for the throats analogously. The black curve
follows the invasion of the largest throats in the spanning clusters, while the randomly distributed
points below the curve correspond again to the PSD of throats in the PN.

Figure 7. Capillary pressure curves of the drainage invasion process initiated in all surface pores. Note
that the small pores and throats with higher capillary pressures (Figure 6) are not invaded.

Figure 8 summarizes the gas-liquid distributions of the PN simulation. At first, the situation a few
moments before loss of the liquid connectivity between the vertical throats in the top and bottom row
is illustrated in Figure 8a. As can be seen, the liquid phase is multiply disconnected and penetrated by
gas branches (in white). The gas phase forms a fractal structure that moves downwards towards the
water supply channel. The invasion of the gas phase leads to the formation of isolated liquid clusters

37
Processes 2019, 7, 558

that remain behind the front. The size of these clusters is stable as they are not further invaded. It is
remarked that the invasion process occurs in all three dimensions in the homogeneous pore structure
underlying this simulation (isotropic invasion). This is explained with the equal distribution of liquid
pressure (associated with PSDs) throughout the PN, constant wettability, temperature, and pressure
conditions. However, it is noted that the horizontal breakthrough occurs much earlier than the vertical
breakthrough. This is explained with the oxygen accessibility of the surface pores. In detail, all of the
surface pores are initially connected to one liquid-filled throat (downwards) and the gas bulk phase
(upwards). Essentially, based on the invasion percolation algorithm and due to the fact that the peak of
the PSD of pores is shifted towards larger radii, all pores empty initially before any throat is invaded.
This results in the formation of an oxygen front that spans the PN horizontally (the size is roughly
NixNj) (Figure 8c,d). According to this, the liquid connectivity between the boundary pores at the
top side and the bottom side is already lost at the start of the process. A different situation would be
expected if the surface pores were smaller and the invasion pressure were higher.

(a) (b)

(c) (d)

Figure 8. (a) PN shortly before loss of connectivity of the liquid supply route through the blue cluster.
(b) Only the liquid transporting cluster (main cluster) is shown here. Note that due to the applied
periodic boundary conditions, the cluster appears at two sides of the PN, while it is disconnected in the
center. (c) Empty surface pores at the top side (catalyst side). Pores connected to a liquid-filled throat
of the spanning cluster are shown in light blue. (d) Surface pores at the bottom side (water channel
side). Pores connected to the spanning (main) cluster in blue; all other pores in red.

At the moment when the liquid supply route is interrupted (or shortly before that moment),
the overall saturation with liquid is S = 0.5996. However, as shown in Figure 8, there exists only one
liquid transporting cluster (main cluster). The other liquid-filled throats are shown in grey and are
either single throats or part of isolated clusters. The number of clusters at the interruption of liquid
connectivity between the surface throats (first and last row of vertical throats) is 933. The maximum
cluster size is 624 throats; this is the blue cluster in Figure 8a,b. The mean size of the other clusters
is 9.5 throats, thus approximately two orders of magnitude smaller. The liquid transporting cluster
contains approximately 5.7417 × 10−3 μL of water. Its volume referred to the total liquid volume
contained in the PN at the interruption of the liquid path is only 3%. This reveals the relevance of
the pore scale information about the liquid connectivity for the calculation of the liquid permeability.
A different situation is expected in presence of wetting liquid films, as will be discussed below.
It is also noted that higher permeabilities would be obtained if the drainage simulation would be
interrupted already at a higher overall saturation. However, in the simulation presented in Figures 7–9,

38
Processes 2019, 7, 558

a breakthrough of the gas phase occurred shortly before disconnection of the main cluster from the
top side.
Furthermore, the saturation of the exchange interfaces plays a vital role. The surface and bottom
saturations are highlighted in Figure 8c,d. Note that the light blue pores at the top side are already
empty. These pores are connected to liquid-filled vertical throats right below the surface. The bottom
pores shown in blue are liquid filled. The red pores in these images are either empty (filled with
oxygen) or belong to isolated clusters. The ratio of the wetted surface (taking into account the vertical
throats inside the first and last layer of the PN) is around 2% at the top surface and around 1% at the
bottom side, thus very low. This results in low liquid permeabilities, as will be shown below.
From the distributions shown in Figure 8, the overall liquid saturation Sl of the pore network can
be calculated from:

22500

9000
Vt · S t + Vp · Sp
Nt = 1 Np = 1
Sl =

, (11)
Vt + Vp
with pore and throat saturation Sp and St . The overall gas saturation Sg is:

Sg = 1 − Sl , (12)

accordingly. The liquid and gas saturations are calculated for each network slice (Nk = 1:10) and
illustrated as a function of the vertical position ((Nk − 1)·L, see Figure 1) in Figure 9a,b. Figure 9a,b
shows the transient saturation profiles of one PN simulation. It reveals that the PN is basically invaded
by oxygen from the top side. If one disregards the top and bottom surface layers of the PN (as they have
a different overall volume, Figure 1), the saturation of liquid is evenly distributed. This means that
isolated liquid clusters homogeneously cover the PN. The liquid confined in the liquid transporting
main cluster spanning the PN is illustrated in Figure 9c. The liquid saturation of the main cluster
is calculated analog to Equation (11) by only taking into account the liquid inside it. Two different
options are compared with the liquid saturation profiles in Figure 9c. These are option 1:

Vt · St,MC + Vp · Sp,MC
Nt ( Nk ) Np (Nk )
SMC,Vtot =

, (13)
Vt + Vp
Nt ( Nk ) Np (Nk )

saturation on the base of the total void volume contained in slice Nk ; and option 2:

Vt · St,MC + Vp · Sp,MC
Nt (Nk ) Nt ( N k )
SMC,Vl =

, (14)
Vt · St + Vp · Sp
Nt ( Nk ) Np ( Nk )

saturation on the base of the total liquid volume contained in that slice. Note that only the profiles
for S = 1 till S = 0.62 are shown here, because the main cluster splits into two clusters at S = 0.62.
The agreement of the curves in Figure 9a,c is good at the start of drainage (when the overall saturation
S is high), indicating that initially nearly all the liquid is contained in one cluster. At later stages
of the invasion process, a gradient develops. Comparison of the blue and black lines in Figure 9c
with the red curves reveals that the center of the volume of the main cluster is located at the bottom
side of the PN, thus closer to the water supply channel, whereas connectivity to the top side is lost
already at the very beginning of the drainage (when all surface pores are invaded). The difference
between the black curves and the red curves is related to the liquid volume contained in single clusters.
As can be seen, more single clusters are found inside the upper region of the pore network (i.e.,
at higher values of (Nk − 1)·L). This reveals that the main cluster is traveling through the PN, leaving
the isolated clusters behind its front. From this it can be expected that the relative permeability of the

39
Processes 2019, 7, 558

liquid phase continuously decreases with progress of invasion, while the oxygen permeability is 0 as
long as the breakthrough of the gas phase is not observed (Section 3.2). This outcome might be an
indicator for the relevance of the thickness of the PTL for water exchange processes in water electrolysis
cells [56]. According to this, thinner PTLs might be more convenient in terms of the water supply
routes. However, further simulations are necessary to proof the consistency of this assumption. This
finding is essentially an explanation for the low permeability of the liquid phase, as discussed below.

(a) (b) (c)

Figure 9. (a) Transient liquid saturation profiles of the one PN simulation shown in Figures 7 and 8.
(b) The according gas saturation profiles. The catalyst side is found on the right (corresponding to the
top of PN) and the water supply channel is found on the left (corresponding to the bottom of PN).
The saturation profiles are shown for overall liquid saturation S = [1:0.61] (the final saturation is S = 0.6).
(c) Saturation profiles of the liquid transporting clusters for S = [1:0.62].

It is remarked that different simulation results are obtained for a higher standard deviation (1 μm)
(Figure 10). The liquid saturation is still homogeneously distributed throughout the network, but the
level of the final slice saturation is lower, namely S = 0.575 (in Figure 9 S = 0.6). Moreover, the main
liquid cluster is larger and appears more dense. It contains 8% of the total liquid volume. This reflects
the importance of studying the impact of the PSD rather than the impact of porosity, which is around
21% in both PNs [57].

(a) (b)

Figure 10. (a) Transient saturation profiles of PN with a standard deviation of 1 μm. The saturation
profiles are shown for overall liquid saturation S = [1:0.58] (the final saturation is S = 0.575).
(b) Main cluster.

Consistency of these results is proven by each 20 repetitions of the PN simulation (PNMCS).


The results are illustrated in Figure 11, which shows the steady state saturation profiles at the moment
of interruption of liquid transport. It is highlighted that the average saturation in the center of the
PN is slightly higher in the network with lower standard deviation. This means that more liquid is
contained in isolated clusters in this case, as previously observed.

40
Processes 2019, 7, 558

(a) (b)

Figure 11. (a) Steady state saturation profiles of 20 PN realizations with standard deviation 0.5 μm.
(b) Steady state saturation profiles of 20 PN realizations with standard deviation 1 μm.

Similar simulation results are obtained if the PN surface is invaded at only one sight. In the
simulation shown in Figure 12, the one open pore at the surface is highlighted in light blue. Note that
the PN is identical to the one shown in Figures 7–9. The invasion follows a similar route as before
and the main cluster spanning the PN at breakthrough is almost identical. The saturation profiles in
Figure 12 are obviously different because the surface pore row basically remains saturated. However,
the average saturation in the center of the PN is again 0.6 at the end of the process. Isotropy of the
invasion process can be shown for such a situation, i.e., when only one pore in the center of the surface
area is accessible for oxygen (Figure 13).

(a) (b)

Figure 12. (a) PN invasion from one surface pore (shown is the complete PN with gas phase in white,
disconnected clusters in grey, and the liquid transporting cluster at the breakthrough of the gas phase
in blue). (b) Saturation profiles for S = [1:0.64].

Figure 13a plots the label of that in the invasion event invaded vertical (Nk -direction) or horizontal
(Ni and Nj -direction) throat (also refer to Figure 1). The three clouds at the start of the process indicate
that throats are invaded in each direction (red circles). The breakthrough point is achieved when the
throat with the lowest label in each horde is invaded. This is highlighted in Figure 13a. The plot

41
Processes 2019, 7, 558

furthermore reveals that, afterwards, horizontal and vertical throats are equally invaded. Figure 13b
shows the invasion velocity:
dI Nver,hor
= , (15)
dItot Ntot
i.e., the number of invasions in the horizontal and vertical throats (Nhor and Nver , respectively) I’
related to the total number of invasions (pores and throats) Itot . According to Figure 13b, more vertical
throats are initially invaded, whereas, at later stages of the drainage process, the invasion occurs more
often in horizontal throats. This is explained with the PN geometry in Figure 5. It is the main reason
for the clustering effect discussed above. In order to prevent the clustering effect, it would be more
convenient if the invasion process was anisotropic and occurred mainly in vertical throats. Taking the
invasion pressure curve in Figure 6 as a reference, this could be achieved by a gradient in throat sizes
(small at the top side, large at the bottom side).

(a) (b)

Figure 13. (a) Order of invasion on the example of the PN discussed above. According to Figure 1,
higher throat labels are associated with the top side of the PN. (b) Invasion velocities in vertical and
horizontal direction.

It is highlighted that a different invasion behavior would occur if also temperature, pressure,
or wettability would spatially vary along the vertical or the horizontal direction. Depending on
the situation, anisotropic invasion can also be expected. We plan to illustrate such situations in a
forthcoming paper.

3.2. Estimation of Relative Permeabilities


The final gas-liquid phase distribution of the PN drainage simulation is employed for the
calculation of gas and liquid permeability. Note that due to the trapping of clusters, the relative
permeabilities of the liquid and gas phases cannot be related to each other (kl  1 − kg ).
For calculation of the absolute and relative permeabilities, we have considered the liquid flow
through pores in Nk = 2 and Nk = 9 (Figures 1 and 5) because all surface pores are initially invaded.
The pressure gradient is imposed between the pores in these rows. The relative permeability of the
liquid phase is related to the total surface area, although only 1–2% of the interfaces are wetted by
the liquid spanning cluster. For the PN simulation presented in Figures 7–9, we found an absolute
permeability of K = 1.9639 × 10−12 m2 (liquid phase) and K = 6.5612 × 10−12 m2 (gas phase) (in [50],
similar values are obtained for a porosity of 50%). The relative permeabilities are kl = 7.8439 × 10−4
for the liquid phase and kg = 0.0068 for the gas phase. The relatively low oxygen permeability is
associated to the hindrance by isolated liquid clusters. Higher liquid permeabilities are obtained
if the calculations are repeated for higher liquid saturations of the PN, i.e., if the PN is not fully

42
Processes 2019, 7, 558

drained (Figure 14). (Note that the gas phase is not penetrating the PN for liquid saturations shown in
Figure 14). These values are in good agreement with previous results in [45] and they clearly illustrate
the different relative permeabilities of 2D and 3D PNs, which are related to the higher connectivity of
the 3D PN and the associated clustering effect. This situation can change if wetting liquid films are
considered in the PNM [57]; however, their development and extension depend on the wettability of
the surface, temperature, pressure, and process conditions. (In [57] it is mentioned that the wettability
alteration of titanium PTLs occurs due to the oxidation of its surface.) Additionally, a heterogeneous
pore structure with interpenetrating large and small pores can lead to higher relative permeabilities if
the breakthrough occurs in the large pores, whereas the small pores remain saturated and provide
the pathway for water transport. The relative permeabilities are thus a function of PSD [57]. The
absolute permeability depends on the ratio of pore space to solid, i.e., on the porosity (e.g., [1,50]).
However, according to Equation (10), higher absolute permeability does not necessarily increase
relative permeabilities.

Figure 14. Permeability variation for simulations with low and high standard deviation of PSD.

3.3. Oxygen Production Rate


The decomposition of water occurs if a voltage higher than the thermoneutral voltage is supplied
to the electrolysis cell, e.g., U = 1.5–2.2 V [43]. It is anticipated that, at low voltages, oxygen can be
solved inside the water that is transported through the PTL towards the catalyst layer. However,
if higher oxygen production rates are achieved, the gas can invade the PTL if the pressure in the gas
phase is high enough. In [58,59], the different transport regimes for oxygen through the PTL that
correlate with the current density of the electrolyzer are shown. In this, dispersed bubbly flow (#1),
plug flow (#2), slug flow (#3), churn flow (#4), and annular flow (#5) are distinguished. The latter
flow regime is based on wetting liquid films forming along the surface of the pores and sustaining
liquid transport between the water flow channel and catalyst layer. For simplicity, we assumed that
the current density is always high enough to allow for a plug flow of the gas phase without formation
of liquid films; this is option #6 (Figure 15a,b). This assumption is reasonable in the face of the total
volume of the pore network of only a few μL (see below). However, if good wettability of the PTL
surface is anticipated (hydrophilic properties), liquid films are likely to form along the solid pore
surface. These films can enhance liquid transport through invaded pores as they can connect the
water supply channel with liquid clusters at the top side of the PTL [15,16,60]. Basically, in presence of
liquid films, the previously (and above discussed) isolated clusters would not occur, but rather the
complete liquid phase in the PN would be interconnected and participate in liquid transport. This
could enhance liquid transport properties significantly. (Exemplarily, in drying processes, liquid films
can reduce the overall drying time by orders of magnitude, e.g., [15,17]). The presence of these films
strongly depends on the surface roughness, the geometry of pore corners, wettability, temperature,
and pressure, as previously mentioned.

43
Processes 2019, 7, 558

(a) (b)

Figure 15. (a) Plug flow invasion of the gas phase (invasion regime #6) completely separating the gas
transporting from the liquid transporting pore space. (b) Annular invasion of the gas phase according
to [58] (invasion regime #5), leading to annularly wetted pore space transporting gas towards the water
supply channel in the center and liquid films transporting water towards the catalyst layer, as well as
fully saturated pores and throats.

The drainage algorithm presented above in Section 2 works independently of the current oxygen
production rate. We postulate that the oxygen production rate is always high enough to (i) exceed the
solubility threshold of oxygen in water (around 40 mg/kg water at 25 ◦ C and 1 bar up to 1 g/kg water
at the same temperature and around 25 bar [61]) and to (ii) simultaneously flood the drained channels
with oxygen. If this is fulfilled, the invasion process can be seen as fast and is expected to take place in
a very short time period.
The oxygen production rate can be calculated using Faraday’s law, assuming that the current
efficiency of oxygen is equal to 1 and stochiometry

2H2 O  4H+ + 4e− + O2 . (16)

Based on Faraday’s law:


Q = I · t = F · z · NO2 , (17)

where z is the number of exchanged electrons (z = 4) and F is Faraday’s constant


(F = 9.64853 × 104 A s mol−1 ) the moles of electrolyzed water can be correlated with the current
of the electrolyzer. From this follows:

dNO2 . I
= NO2 = , (18)
dt F·z
the formed molar amount of oxygen per unit time. The volume flux of oxygen can be further
calculated from:
3 .
. m NO2 R·T J
R·T
vO2 2 = · = · , (19)
m s A P F·z P
with current density
I
J= (20)
A
in A m−2 . The relationship in Equation (19) is illustrated in Figure 16a; the linear dependence is in
agreement with values reported in [39]. Taking the PN from Section 3.1 as a basis, the surface area
connected to the catalyst layer is roughly 0.22 mm2 . If it is furthermore assumed that the complete
surface area is electrolytically reactive, the oxygen flux in Figure 16a can be converted into a volumetric
flow rate: .    .  
VO2 μL min−1 = 0.22 mm2 vO2 mm min−1 · 103 . (21)

44
Processes 2019, 7, 558

(a) (b)

Figure 16. (a) Linear dependence of the oxygen volume flux on current density. (b) Required invasion
time for the constant PN volume VPN (slope) in dependence of the reversed volumetric production
flow rate.

Figure 16b illustrates the time dependence of the PN invasion on the reversed volumetric flow
. .
1/VO2 = 1/vO2 A [sμL−1 ], with the slope being the total volume of the PN in Section 3.1,

1
t = VPN . , (22)
vO2 A

and A being the total cross section of the PN surface. The total volume of the PN discussed in Section 3.1
is approximately VPN = 0.19 μL; the pore volume is 0.0234 μL and the throat volume is 0.17 μL. In detail,
at a current density of 2 A cm−2 , the complete PN could be invaded within 0.0675 s, whereas reduction
of the current density by factor 1000 increases the invasion time to roughly 1 min. This is in agreement
with experiments reported in [36], where, at postulated current densities of 7 A cm−2 , the according
volume flow rates of the gas phase resulted in invasion times <1 min. Note that always only the dry
part of the PN is available for the gas phase. Since this region is much smaller than the total volume,
the breakthrough of the gas phase can occur in a shorter time. Additionally, the available cross section
for gas invasion is usually lower than the given value in the presence of some liquid pores at the
surface (Figure 8).

4. Summary and Discussions


We have presented a method to study the pore scale transport of oxygen and water through
the PTL of water electrolysis cells. The method is based on Monte Carlo pore network simulations
(PNMCS). It allows simulation of the distribution of the two fluids inside the pore space on a physical
base, i.e., without incorporating effective parameters. We have shown that this can be a useful and
reliable tool to numerically estimate the pore scale distribution of gas and liquid and the transport
coefficients of PTLs, such as relative and absolute permeabilities. More clearly, if these are estimated
more accurately based on the proposed method, transport characteristics of PTLs could be predicted
more precisely in the future. Moreover, it is highlighted that the discrete method allows correlation of
specific transport characteristics with the individual structure of the pore space. The proposed method,
therefore, opens up a new route for designing powerful PTLs on customized demand in the future.
The developed PNM basically incorporates invasion percolation rules for hydrophilic drainage
with trapping of liquid clusters. Pores and throats are separately invaded in our PN because of
their similar invasion pressure thresholds. We assumed plug flow of the liquid and gas phases
and negligible film flow. We also assumed that the oxygen production rate is always high enough
to homogeneously invade the PN through all surface pores; pressure gradients were disregarded.

45
Processes 2019, 7, 558

The correlation with Faraday’s law showed that the small PN can be invaded during less than one
minute if the current density is higher than 2 mA cm−2 . The invasion process was simulated until the
disconnection of the spanning liquid clusters, i.e., when the transport route for water from the top
(assumed to be connected to the catalyst layer) to the bottom (assumed to be connected to water supply
channel) was interrupted. Due to the discrete character of the model, it was possible to distinguish
between isolated clusters, which do not contribute to liquid transport but rather hinder the oxygen
flow, and the liquid transporting clusters, which span the network. The possibility to separate single
liquid pores and throats depending on their designation as transporting or non-transporting elements
particularly reveals the strength of PN modeling and the perspectives for future applications. Different
situations were studied. At first, we illustrated relevance of the pore size distribution (PSD). Secondly,
we changed the invasion rules of surface pores. For the latter, we allowed invasion of all surface
pores in the first situation and restricted invasion to only one pore in the center of the PN in the
second situation. We have shown that the final saturation of the PNs with liquid as well as the liquid
volume contained in the spanning liquid cluster depends primarily on the PSD; the porosity was kept
constant in the situations studied in this paper. Additionally, our simulations revealed that the liquid
transporting clusters cover only a very small percentage of the total volume of the PN, whereas most of
the liquid phase is disconnected in isolated single clusters if the drainage simulation is continued until
disconnection of the water supply route. Additionally, the spanning cluster was travelling through
the PN during the drainage invasion and had its center of mass at the bottom side at the moment
when liquid connectivity was interrupted. The clustering effect and the travelling spanning cluster
are major outcomes of the PN drainage simulation with a significant impact on the liquid and gas
permeabilities, which were very low in the studied cases. Furthermore, we found that the spanning
liquid clusters providing the transport route for water cover less than 2% of both surfaces (top, bottom).
Though this low value can be attributed to the idealized PN structure, the associated low porosity, and
the postulated constant boundary conditions, new PN simulations shall further enlighten the major
characteristics of mass transfer in PTLs. Therefore, we plan to incorporate micro tomography scans of
the PTL in the PNM in the next step. The instrument to be used is a Procon CT alpha with maximum
resolution of 0.6 μm, equipped with image processing software, Mavi, from the Fraunhofer Institute
for Industrial Mathematics ITWM Kaiserslautern. Following concepts, e.g., described in [1,2,62], we
aim to run the PN simulation using the extracted real structure of the porous medium. This will allow
more realistic prediction of the gas-liquid distribution for real PTLs as well as study of the role of
heterogeneous pore structures, such as bimodal PSDs. Furthermore, microfluidic experiments will be
necessary to validate the assumptions of the model. Current open questions concern the limits of the
flow regimes (based on discussion in [58,59]), the relevance of liquid flow through corner films [57],
the impact of local temperature and wettability variations, as well as the dynamic invasion in the
presence of current density fluctuations. Based on [36], we plan to illustrate in more detail the impact
of flow regimes (in terms of higher capillary numbers), as well as unsteady operation conditions
related to the fluctuation of the current density. In the latter case, we expect multiple redistribution of
liquid due to pressure (and eventually also temperature) variations. This implies the application of
imbibition invasion rules additionally to drainage rules. In this context, it will also be worth studying
if the disconnected liquid clusters can be reconnected and open up more routes for water transport,
especially in the situation where liquid films support liquid flow through corners.

Author Contributions: Conceptualization, N.V., E.T., and T.V.-K.; methodology, N.V., E.T., and T.V.-K.; software,
N.V. and H.A.; validation, N.V., H.A., and T.V.-K.; formal analysis, N.V. and T.V.-K.; investigation, N.V.; resources,
N.V. and E.T.; data curation, N.V. and H.A.; writing—original draft preparation, N.V.; writing—review and editing,
T.V.-K.; visualization, N.V.; supervision, N.V.; project administration, N.V., E.T., and T.V-K.; funding acquisition,
N.V., E.T., and T.V.-K.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.

46
Processes 2019, 7, 558

Abbreviations
Symbol Parameter (Unit)
A Area (m2 )
A Matrix of conductances (-)
b Vector of boundary conditions (-)
d Diameter (m)
F Faraday constant (A s mol−1 )
g Conductance (m s)
I Current (A)
I’, Itot Number of invasion events (-)
J Area related current (A m−2 )
k Relative permeability (-)
K Absolute permeability (m2 )
L Length (m)
.
M Flow rate (kg s−1 )
N Molar amount (mol), number (-)
Ni , Nj , Nk Room coordinates (-)
P Pressure (Pa)
Pc Capillary pressure (Pa)
pnp Matrix of pore neighbor relations (-)
pnt Matrix of pore and throat neighbor relations (-)
Q Electric charge (C)
r Radius (m)

R Universal gas constant (J mol−1 K−1 )
S Saturation (-)
t Time (s)
tnp Matrix of throat and pore neighbor relations (-)
tnt Matrix of throat neighbor relations (-)
T Temperature (K)
U Voltage (V)
V Volume (m3 )
.
v Velocity (m s−1 )
.
V Volume flow rate (m3 s−1 )
z Valency number (-), room coordinate (m)
η Dynamic viscosity (Pa s)
θ Contact angle (◦ )
ρ Density (kg m−3 )
σ Surface tension (N m−1 )
Subscripts
1,2 Pore 1 or 2
av Average value
g Gas phase
hor Horizontal throats
k Slice index/number
l Liquid phase
p Pore
PN Pore network
MC Main cluster
t Throat
tot Total
V Volume related
ver Vertical throats

47
Processes 2019, 7, 558

Abbreviations
PN Pore network
PNM Pore network model
PNMCS Pore network Monte Carlo simulation
PSD Pore size distribution
PTL Porous transport layer

References
1. Blunt, M.J. Multiphase Flow in Permeable Media, 1st ed.; Cambridge University Press: Cambridge, UK, 2017.
2. Bultreys, T.; Van Hoorebeke, L.; Cnudde, V. Multi-scale, micro-computed tomography-based pore network
models to simulate drainage in heterogeneous rocks. Adv. Water Resour. 2015, 78, 36–49. [CrossRef]
3. Niasar, V.J.; Hassanizadeh, S.M.; Pyrak-Nolte, L.J.; Berentsen, C. Simulating drainage and imbibition
experiments in a high-porosity micromodel using an unstructured pore network model. Water Resour. Res.
2009, 45, W02430.
4. Patzek, T.W. Verification of a complete network simulator of drainage and imbibition. SPE J. 2000. [CrossRef]
5. Sun, Y.; Kharaghani, A.; Tsotsas, E. Micro-model experiments and pore network simulations of liquid
imbibition in porous media. Chem. Eng. Sci. 2016, 150, 41–53. [CrossRef]
6. Metzger, T. A personal view on pore network models in drying technology. Dry. Technol. 2019, 37, 497–512.
[CrossRef]
7. Vorhauer, N.; Tsotsas, E.; Prat, M. Drying of thin porous disks from pore network simulations. Dry. Technol.
2017, 36, 651–663. [CrossRef]
8. Prat, M. Percolation model of drying under isothermal conditions in porous media. Int. J. Multiph. Flow
1993, 1, 691–704. [CrossRef]
9. Börnhorst, M.; Walzel, P.; Rahimi, A.; Kharaghani, A.; Tsotsas, E.; Nestle, N.; Besser, A.; Kleine Jäger, F.;
Metzger, T. Influence of pore structure and impregnation: Drying conditions on the solid distribution in
porous support materials. Dry. Technol. 2016, 34, 1964–1978. [CrossRef]
10. Segura, L.A. Modeling at pore-scale isothermal drying of porous materials: Liquid and vapor diffusivity.
Dry. Technol. 2007, 25, 1677–1686. [CrossRef]
11. Metzger, T.; Tsotsas, E. Viscous stabilization of drying front: Three-dimensional pore network simulations.
Chem. Eng. Res. Des. 2008, 86, 739–744. [CrossRef]
12. Yiotis, A.G.; Stubos, A.K.; Boudouvis, A.G.; Tsimpanogiannis, I.N.; Yortsos, Y.C. Pore-network modeling of
isothermal drying in porous media. Transp. Porous Med. 2005, 58, 63–86. [CrossRef]
13. Prat, M.; Bouleux, F. Drying of capillary porous media with a stabilized front in two dimensions. Phys. Rev. E
1999, 60, 5647–5656. [CrossRef] [PubMed]
14. Fenwick, D.H.; Blunt, M. Three-dimensional modeling of three phase imbibition and drainage. Adv. Water
Res. 1998, 21, 121–143. [CrossRef]
15. Prat, M. On the influence of pore shape, contact angle and film flows on drying of capillary porous media.
Int. J. Heat Mass Tran. 2002, 50, 1455–1468. [CrossRef]
16. Yiotis, A.G.; Boudouvis, A.G.; Stubos, A.K.; Tsimpanogiannis, I.N.; Yortsos, Y.C. Effect of liquid films on the
drying of porous media. AIChE J. 2004, 50, 2721–2737. [CrossRef]
17. Vorhauer, N.; Wang, Y.; Kharaghani, A.; Tsotsas, E.; Prat, M. Drying with formation of capillary rings in a
model porous medium. Transp. Porous Med. 2015, 110, 197–223. [CrossRef]
18. Rahimi, A.; Metzger, T.; Kharaghani, A.; Tsotsas, E. Discrete modelling of ion transport and crystallization in
layered porous media during drying. In Proceedings of the 21th International Drying Symposium (IDS2018),
Valencia, Spain, 11–14 September 2018.
19. Rahimi, A.; Kharaghani, A.; Metzger, T.; Tsotsas, E. Pore network modelling of a salt solution droplet on
a porous substrate: Imbibition, evaporation and crystallization, In Proceedings of the 20th International
Drying Symposium (IDS2016), Gifu, Japan, 7–10 August 2016.
20. Veran-Tissoires, S.; Prat, M. Evaporation of a sodium chloride solution from a saturated porous medium
with efflorescence formation. J. Fluid Mech. 2014, 749, 701–749. [CrossRef]

48
Processes 2019, 7, 558

21. Surasani, V.; Metzger, T.; Tsotsas, E. Drying simulations of various 3D pore structures by a nonisothermal
pore network model. Dry. Technol. 2010, 28, 615–623. [CrossRef]
22. Plourde, F.; Prat, M. Pore network simulations of drying of capillary porous media: Influence of thermal
gradients. Int. J. Heat Mass Tran. 2003, 46, 1293–1307. [CrossRef]
23. Huinink, H.P.; Pel, L.; Michels, M.A.J.; Prat, M. Drying processes in the presence of temperature gradients:
Pore-scale modelling. Eur. Phys. J. 2002, 9, 487–498. [CrossRef]
24. Vorhauer, N.; Tsotsas, E.; Prat, M. Temperature gradient induced double-stabilization of the evaporation
front within a drying porous medium. Phys. Rev. Fluids 2018, 3, 114201. [CrossRef]
25. Vorhauer, N.; Tran, Q.T.; Metzger, T.; Tsotsas, E.; Prat, M. Experimental investigation of drying in a model
porous medium: Influence of thermal gradients. Dry. Technol. 2013, 31, 920–929. [CrossRef]
26. Hinebaugh, J.; Bazylak, A. Condensation in PEM fuel cell gas diffusion layers: A pore network modeling
approach. J. Electrochem. Soc. 2010, 157, 1382–1390. [CrossRef]
27. Shahraeeni, M.; Hoorfar, M. Pore-network modeling of liquid water flow in gas diffusion layers of proton
exchange membrane fuel cells. Int. J. Hydrog. Energy 2014, 39, 10697–10709. [CrossRef]
28. Lee, K.J.; Kang, J.H.; Nam, J.H. Liquid water distribution in hydrophobic gas diffusion layers with interconnect
rib geometry: An invasion-percolation pore network analysis. Int. J. Hydrog. Energy 2014, 39, 6646–6656.
[CrossRef]
29. Lee, K.J.; Nam, J.H.; Kim, C.J. Steady saturation distribution in hydrophobic gas-diffusion layers of polymer
electrolyte membrane fuel cells: A pore-network study. J. Power Sources 2010, 195, 130–141. [CrossRef]
30. Lee, K.J.; Kang, J.H.; Nam, J.H.; Kim, C.J. Steady liquid water saturation distribution in hydrophobic
gas-diffusion layers with engineered pore paths: An invasion-percolation pore-network analysis. J. Power
Sources 2010, 195, 3508–3512. [CrossRef]
31. Straubhaar, B.; Pauchet, J.; Prat, M. Pore network modelling of condensation in gas diffusion layers of proton
exchange membrane fuel cells. Int. J. Heat Mass Tran. 2016, 102, 891–901. [CrossRef]
32. Agaesse, T.; Lamibrac, A.; Büchi, F.N.; Pauchet, J.; Prat, M. Validation of pore network simulations of ex-situ
water distributions in a gas diffusion layer of proton exchange membrane fuel cells with X-ray tomographic
images. J. Power Sources 2010, 331, 462–474. [CrossRef]
33. Aghighi, M.; Gostick, J. Pore network modeling of phase change in PEM fuel cell fibrous cathode. J. Appl.
Electrochem. 2017, 47, 1323–1338. [CrossRef]
34. De Beer, F.; van der Merwe, J.H.; Bessarabov, D. PEM water electrolysis: Preliminary investigations using
neutron radiography. Phys. Procedia 2017, 88, 19–26. [CrossRef]
35. Biesdorf, J.; Oberholzer, P.; Bernauer, F.; Kästner, A.; Vontobel, P.; Lehmann, E.H.; Schmidt, T.J.; Boillat, P.
Dual spectrum neutron radiography: Identification of phase transitions between frozen and liquid water.
Phys. Rev. Lett. 2014, 112, 248301. [CrossRef] [PubMed]
36. Lee, C.H.; Hinebaugh, J.; Banerjee, R.; Chevalier, S.; Abouatallah, R.; Wang, R.; Bazylak, A. Influence of
limiting throat and flow regime on oxygen bubble saturation of polymer electrolyte membrane electrolyzer
porous transport layers. Int. J. Hydrog. Energy 2017, 42, 2724–2735. [CrossRef]
37. Arbabi, F.; Montazeri, H.; Abouatallah, R.; Wang, R.; Bazylak, A. Three-dimensional computational fluid
dynamics modelling of oxygen bubble transport in polymer electrolyte membrane electrolyzer porous
transport layers. J. Electrochem. Soc. 2016, 163, 3062–3069. [CrossRef]
38. Arbabi, F.; Kalantarian, A.; Abouatallah, R.; Wang, R.; Wallace, J.; Bazylak, A. Visualizing bubble flows in
electrolyzer GDLs using microfluidic platforms. ECS Trans. 2013, 58, 907–918. [CrossRef]
39. Arbabi, F.; Kalantarian, A.; Abouatallah, R.; Wang, R.; Wallace, J.S.; Bazylak, A. Feasibility study of using
microfluidic platforms for visualizing bubble flows in electrolyzer gas diffusion layers. J. Power Sources 2014,
258, 142–149. [CrossRef]
40. Vorhauer, N.; Metzger, T.; Tsotsas, E.; Prat, M. Experimental investigation of drying by pore networks:
Influence of pore size distribution and temperature. In Proceedings of the 4th International Conference on
Porous Media and its Applications in Science, Engineering and Industry, Potsdam, Germany, 17–22 June
2012.
41. Morrow, N.R. Physics and thermodynamics of capillary action in porous media. Ind. Eng. Chem. 1970, 62,
32–56. [CrossRef]

49
Processes 2019, 7, 558

42. Babic, U.; Suermann, M.; Büchi, F.N.; Gubler, L.; Schmidt, T.J. Critical Review—Identifying critical gaps for
polymer electrolyte water electrolysis development. J. Electrochem. Soc. 2017, 164, 387–399. [CrossRef]
43. Bernt, M.; Gasteiger, H.A. Influence of ionomer content in IrO2/TiO2 electrodes on PEM water electrolyzer
performance. J. Electrochem. Soc. 2016, 163, 3179–3189. [CrossRef]
44. Moghaddam, A.A.; Kharaghani, A.; Tsotsas, E.; Prat, M. Kinematics in a slowly drying porous medium:
Reconciliation of pore network simulations and continuum modeling. Phys. Fluids 2017, 29, 022102.
[CrossRef]
45. Vorhauer, N.; Metzger, T.; Tsotsas, E. Extraction of effective parameters for continuous drying model from
discrete pore network model. In Proceedings of the 17th International Drying Symposium (IDS 2010),
Magdeburg, Germany, 3–6 October 2010; pp. 415–422.
46. Schuler, T.; De Bruycker, R.; Schmidt, T.J.; Büchi, F.N. Polymer Electrolyte Water Electrolysis: Correlating
porous transport layer structural properties and performance: Part I. Tomographic analysis of morphology
and topology. J. Electrochem. Soc. 2019, 166, 555–565. [CrossRef]
47. Metzger, T.; Irawan, A.; Tsotsas, E. Influence of pore structure on drying kinetics: A pore network study.
AIChE J. 2007, 53, 3029–3041. [CrossRef]
48. Blunt, M.J.; Jackson, M.D.; Piri, M.; Valvatne, P.H. Detailed physics, predictive capabilities and macroscopic
consequences for pore-network models of multiphase flow. Adv. Water Resour. 2002, 25, 1069–1089. [CrossRef]
49. Lenormand, R. Flow through porous media: Limits of fractal patterns. Proc. R. Soc. Lond. A 1989, 423, 159–168.
[CrossRef]
50. Zielke, L.; Fallisch, A.; Paust, N.; Zengerle, R.; Thiele, S. Tomography based screening of flow field/current
collector combinations for PEM water electrolysis. RSC Adv. 2014, 4, 58888. [CrossRef]
51. Tang, T.; Yu, P.; Shan, X.; Chen, H.; Su, J. Investigation of drag properties for flow through and around square
arrays of cylinders at low Reynolds numbers. Chem. Eng. Sci. 2019, 199, 285–301. [CrossRef]
52. Metzger, T.; Tsotsas, E.; Prat, M. Pore-network models: A powerful tool to study drying at the pore level
and understand the influence of structure on drying kinetics. In Drying Technology, Modern Computational
Tools at Different Scales; Tsotsas, E., Mujumdar, A.S., Eds.; Wiley-VCH: Weinheim, Germany, 2011; Volume 1,
pp. 57–102.
53. Zachariah, G.T.; Panda, D.; Surasani, V.J. Lattice Boltzmann modeling and simulation of isothermal drying of
capillary porous media. In Proceedings of the 21th International Drying Symposium (IDS2018), Valencia,
Spain, 11–14 September 2018.
54. Yiotis, A.G.; Psihogios, J.; Kainourgiakis, M.E.; Papaioannou, A.; Stubos, A.K. A lattice Boltzmann study of
viscous coupling effects in immiscible two-phase flow in porous media. Colloids Surf. A Physicochem. Eng.
Asp. 2007, 300, 35–49. [CrossRef]
55. Dhanushkodi, S.R.; Capitanio, F.; Biggs, T.; Merida, W. Understanding flexural, mechanical and
physicochemical properties of gas diffusion layers for polymer membrane fuel cell and electrolyzer systems.
Int. J. Hydrog. Energy 2015, 40, 16846–16859. [CrossRef]
56. Siracusano, S.; Di Blasi, A.; Baglio, V.; Brunaccini, G.; Briguglio, N.; Stassi, A. Optimization of components
and assembling in a PEM electrolyzer stack. Int. J. Hydrog. Energy 2011, 36, 3333–3339. [CrossRef]
57. Bromberger, K.; Ghinaiya, J.; Lickert, T.; Fallisch, A.; Smolinka, T. Hydraulic ex situ through-plane
characterization of porous transport layers in PEM water electrolysis cells. Int. J. Hydrog. Energy 2018, 43,
2556–2569. [CrossRef]
58. Ito, H.; Maeda, T.; Nakano, A.; Hwang, C.M.; Ishida, M.; Kato, A.; Yoshida, T. Experimental study on porous
current collectors of PEM electrolyzers. Int. J. Hydrog. Energy 2012, 37, 7418–7428. [CrossRef]
59. Ito, H.; Maeda, T.; Nakano, A.; Hasegawa, Y.; Yokoi, N.; Hwang, C.M.; Ishida, M.; Kato, A.; Yoshida, T. Effect
of flow regime of circulating water on a proton exchange membrane electrolyzer. Int. J. Hydrog. Energy 2010,
35, 9550–9560. [CrossRef]
60. Geistlinger, H.; Ding, Y.; Apelt, B.; Schlüter, S.; Küchler, M.; Reuter, D.; Vorhauer, N.; Vogel, H.J. Evaporation
study based on micromodel-experiments: Comparison of theory and experiment. Water Resour. Res. 2019.
[CrossRef]

50
Processes 2019, 7, 558

61. Kolev, N.I. Multiphase Fluid Dynamics 4: Turbulence, Gas Adsorption and Release, Diesel Fuel Properties, 1st ed.;
Springer: Berlin, Germany, 2011.
62. Pashminehazar, R.; Ahmed, S.J.; Kharaghani, A.; Tsotsas, E. Spatial morphology of maltodextrin agglomerates
from X-ray microtomographic data: Real structure evaluation vs. spherical primary particle model. Powder
Technol. 2018, 331, 204–217. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

51
processes
Article
Steady-State Water Drainage by Oxygen in Anodic
Porous Transport Layer of Electrolyzers: A 2D Pore
Network Study
Haashir Altaf 1 , Nicole Vorhauer 1, *, Evangelos Tsotsas 1 and Tanja Vidaković-Koch 2
1 Institute of Process Engineering, Otto von Guericke University, 39106 Magdeburg, Germany;
[email protected] (H.A.); [email protected] (E.T.)
2 Electrochemical Energy Conversion, Max Planck Institute for Dynamics of Complex Technical Systems,
39106 Magdeburg, Germany; [email protected]
* Correspondence: [email protected]; Tel.: +49-391-67-51684

Received: 12 February 2020; Accepted: 19 March 2020; Published: 21 March 2020

Abstract: Recently, pore network modelling has been attracting attention in the investigation of
electrolysis. This study focuses on a 2D pore network model with the purpose to study the drainage
of water by oxygen in anodic porous transport layers (PTL). The oxygen gas produced at the anode
catalyst layer by the oxidation of water flows counter currently to the educt through the PTL. When
it invades the water-filled pores of the PTL, the liquid is drained from the porous medium. For
the pore network model presented here, we assume that this process occurs in distinct steps and
applies classical rules of invasion percolation with quasi-static drainage. As the invasion occurs in the
capillary-dominated regime, it is dictated by the pore structure and the pore size distribution. Viscous
and liquid film flows are neglected and gravity forces are disregarded. The curvature of the two-phase
interface within the pores, which essentially dictates the invasion process, is computed from the
Young Laplace equation. We show and discuss results from Monte Carlo pore network simulations
and compare them qualitatively to microfluidic experiments from literature. The invasion patterns of
different types of PTLs, i.e., felt, foam, sintered, are compared with pore network simulations. In
addition to this, we study the impact of pore size distribution on the phase patterns of oxygen and
water inside the pore network. Based on these results, it can be recommended that pore network
modeling is a valuable tool to study the correlation between kinetic losses of water electrolysis
processes and current density.

Keywords: pore network model; drainage invasion; pore size distribution; porous transport
layer; electrolysis

1. Introduction
With innovations in the energy sector and a need for clean fuel, research is in progress to exploit
the potential of hydrogen as an efficient energy source. Exemplarily, fuel cells can utilize hydrogen to
produce electricity, and it can also serve as a fuel for internal combustion engines [1]. For the production
of hydrogen, electrolyzer technology serves as a very promising and viable option. The purity of
the produced hydrogen can be almost 100 vol % [2]. This way, it can be integrated with other
renewable resources to offer a broad range of ecologically clean methods for hydrogen production [3–5].
Shortly, electrolyzers and fuel cells will be able to help alleviate the effects of fossil and nuclear fuel
consumption [2,6,7]. This, however, implies efficient performance of electrolyzers.
A trade-off between the production rates and the efficiency of an electrolyzer is still to be met. For
this technology to be commercial, the cost and hence the efficiency needs to be optimized. Among the
electrolyzer technologies, polymer electrolyte membrane (PEM) electrolyzers have an edge over the

Processes 2020, 8, 362; doi:10.3390/pr8030362 53 www.mdpi.com/journal/processes


Processes 2020, 8, 362

other varieties because of high energy efficiency and compact design [8–10]. Power needed for the
electrolyzer operation can be obtained from renewable sources too and such systems have also been
gaining a lot of attention recently [11]. Polymer electrolyte membrane electrolytic cells (PEMECs) are
very common when coupling the technology with other renewable sources like solar or wind energy.
High current densities are necessary in order to obtain high production rates. The problem
is that the high current density operation results in a decrease of electrical efficiency. This is for
example, reported in [12–16] and it is mainly explained with the kinetic losses associated with the mass
transfer resistances through the porous transport layers (PTL) [17]. The counter-flow of the two phases
causes hindrance against each other, and this mass transfer resistance causes a decrease in the overall
electrolyzer performance [18,19].
It is strongly assumed that the high oxygen production rates achieved at the high current densities
allow for the formation of gas bubbles that can invade the PTL and partially drain the water [20,21].
The size of gas bubbles increases with the increase in current density [12,15,22]. This leads to the
formation of gas fingers penetrating the PTL from the anode catalyst layer, where the oxygen is
produced, to the water supply channel from where the oxygen is removed [8]. The development of
these gas fingers obviously results in the reduction of the overall water loading of the PTL. This can
severely affect the water permeability, especially if the surface saturation of the PTL with water is
significantly reduced by gas-filled pores [23]. On the other hand, the efficiency of the oxygen transport
in the opposite direction depends mainly on the tortuosity of the evolving gas branches.
According to the work done by Suerman et al. [24], 25% of the total cell overpotential could be
contributed to the mass transport losses and these losses are mostly credited to the oxygen withdrawal
from the catalyst surface and the PTLs. Yigit et al. [25] reported that at current densities less than
0.7 A/cm2 , the hydrogen production rates were very low and at values above 1 A/cm2 the electrical
efficiency decreased. Larger pores or high porosity values could easily mitigate this mass transport
problem on the side of the gas phase, but it would also decrease the electron transport and affect
the efficiency [8]. In contrast, small porosity values would hinder the gas removal and increase the
entrapped water amount within the catalyst layer, and thus, decrease the rate of reaction [8]. For this
purpose, current research aims at the structural optimization of PTLs with respect to efficient mass and
electron transfer.
Various experimental methods [12,14–16,20,26–28] and modeling approaches [29–33] are already
established to analyze the key factors of the PTL, such as flow regimes, structure, porosity, pore size
distribution (PSD), permeability, corrosion resistance, and electrical conductivity. From these studies,
the mass transfer limitations discussed before are generally either assigned to the flow regimes of the
gas phase (e.g., in Dedigama et al. [12] and Zhang et al. [34]) or to the structure of the PTL. The latter
has been studied in [14,15,26,29] (Table 1). In Ito et al. [15] an experimental study on a 27 cm2 PEM
electrolyzer cell was presented that investigates the influence of porosity and pore diameter. In this
study, Ti-felt and Ti-powder prepared PTLs with different porosities and different pore structures were
used. According to this study, the optimum pore diameter would be 10 μm, whereas no significant
effect was seen for a porosity value greater than 50%. Grigoriev et al. [14] estimated an optimum value
of 12–13 μm and 30%–50% as the optimum value of porosity using polarization curves for Ti-sintered
PTLs with different properties. Findings presented in Kang et al. [26] in an experimental study based
on thin/well tunable PTLs in a conventional cell suggested high porosity values and small pore sizes.
Ojung et al. [29] used a semi-empirical model in their investigations to study PEM cell without flow
channels. They varied pore sizes between 5–30 μm. They observed a decrease in performance at
5 μm and greater than 11 μm there was no significant improvement shown by their model. They also
concluded that in a system without flow channels, porosity would not influence the performance at
fixed pore diameter and an optimum value of 60% was observed. Hwang et al. [35] also showed with
the experiments on reversible fuel cells using Ti-felt that for mean pore sizes around 30–50 μm porosity
is an insignificant factor.

54
Processes 2020, 8, 362

Table 1. Study of the interrelation of mass transfer limitations and porous transport layers
(PTL) structure.

Estimated Estimated Operated


Reference Type of Technique
Optimum Pore Optimum Current
No. Material Used to Study
Size Porosity Density
[14] Ti-sintered Experimental 12–13 μm 30–50% 0–1.0 A/cm2
Ti (felt +
[15] Experimental 10 μm <50% 0–2.0 A/cm2
sintered)
Thin/well-tunable
[26] Experimental 400 μm 70% 0–2.0 A/cm2
Ti
Semi-empirical
[29] Ti 5–11 μm 60% 0–5.0 A/cm2
model

Besides this, explanations for the structure dependence of the electro activity can also be derived
from the consideration of flow regimes. Dedigama et al. [12] studied the flow regime within the PTL
using electrochemical impedance spectroscopy (EIS) and thermal imaging. They found a reduction in
the mass transfer limitation when passing from the bubble to the slug flow regime. In agreement with
that, Zhang et al. [34] observed a decrease in the efficiency with an increase of the mass flow rate of
water. Aubras et al. [31] showed that the porosity of the anode PTL affects the non-coalesced bubble
regime. According to this study, higher porosity can enhance coalescence of oxygen bubbles and
increase the performance of electrolyzer. Han et al. [36] also showed an increase in performance linked
to an increase in porosity. In addition to that, more recent studies [33,37] revealed the importance
of the interaction of the two involved hierarchical porous structures at the anode electrode, namely
the PTL and the catalyst layer. Exemplarily, it is demonstrated by Lee et al. [37] using micromodel
experiments that the pore sizes control the burst velocity of gas resulting in the application of a thin,
low porosity region at the inlet in order to reduce the gas saturations in the PTL.
The majority of the data suggests that a relatively lower value of pore size (around 10 μm) is
favorable and no significant conclusion can be drawn about the porosity value. Some authors suggest
a high porosity value but others suggest that higher porosity would result in a slug flow, which can
lead to inefficient mass transfer and a decrease in efficiency. On the contrary, coalesced bubble regime
is also suggested to enhance the performance. In our view, other properties besides mean pore size and
porosity might influence the invasion process more significantly. In this paper, we aim at approaching
open questions by means of pore network (PN) modeling. In detail, we will consider the role of PSD,
which has a higher significance for the invasion in PTL than porosity.
From the above discussions, it can be concluded that advanced manufacturing processes are
required to tune the PTL performance. The reader may refer to [26,38–40] for various examples of PTL
production techniques that optimize the structure in terms of electrical efficiency and also in terms of
material consumption and costs. According to Lettenmeier et al. [39] vacuum plasma spraying can be
used to manufacture a PTL with a gradient in pore size along the thickness. It is possible to obtain an
average pore size of 10 μm close to the bipolar plate and 5 μm at the electrode side, with the help of
this technique (Figure 1). This coating technique can also be used to alter other properties of the PTLs
suiting to the need. In a very recent study, Lee et al. [33] showed the effect of porosity gradient on the
performance of the electrolyzer. The low porosity region was towards the membrane side and the
high porosity region on the water inlet side (Figure 1). They observed high water permeation despite
high oxygen saturations. Mo et al. [38] used electron beam melting (EBM) to mitigate the cost and
manufacturing issues of the PTL. They showed 8% improvement in the performance compared to the
conventionally woven PTLs. They obtained smooth surfaces at both ends of the PTL, thereby reducing
the contact resistance between PTL and catalyst layer. Schuler et. al. [41] verified this impact of the
interface between PTL and the catalyst layer clearly in their work.

55
Processes 2020, 8, 362

Figure 1. (a) Schematic representation of different porosity regions in PTL as studied in [33]. (b) Pore
size gradient within PTL as in the study from [39].

2. Pore Network Models


The available methods for studying two-phase flow can be divided into continuum and discrete
models. The continuum models are usually formulated for macroscale continuous phases employing
effective parameters and are thus not suitable for explaining the discrete processes that occur at pore
scale. So, in order to gain a deeper understanding of the invasion processes under the action of
capillarity, viscous or gravity forces, pore scale models are usually preferred. Pore network models
(PNMS) are a type among various pore scale models available, e.g., Lattice Boltzmann models which
are mostly available on the scale of one pore. PNMs are discrete models that represent the pore
space by a lattice of pores and throats (e.g., [42–46]). In comparison to other methods, which are
computationally more expensive in terms of discretization of the physical domain and solving of the
governing equations, PNMs can be used to study larger systems computationally more efficiently.
Besides fundamental physical studies, PNMs are also used for the parameterization of macroscopic
models, e.g., to predict the capillary pressure curve, relative permeability curves, as well as saturation
curves [46].
PNMs are generally distinguished between quasi-static and dynamic models [47]. For the
simulation of the steady-states in capillary-dominated applications [33,48–51], quasi-static models
are used. In the absence of dynamic effects, e.g., driven by viscous forces, the entry capillary
pressure of pores and throats controls the displacement of liquid (drainage) or gas (imbibition) in such
applications [52]. In this work, the quasi-static model from [46] is applied for the simulation of the
drainage of water by oxygen.
The objective of this study is to achieve a fundamental understanding of gas and liquid transport
within the PTLs and the pore structure dependence of the invasion patterns. A regular 2D network of
pores and throats is used to illustrate the porous PTL. The void space of this network comprises of
spherical pores and cylindrical throats. Invasion percolation rules for quasi-static capillary invasion are
employed to simulate the displacement of water by oxygen. The displacement mechanism is controlled
by the capillary pressure curves of water that are obtained from the Young–Laplace equation:

2σcosθ
Pl = Patm − (1)
r

where Pl is the liquid pressure, Patm is the atmospheric pressure, σ in kg s−2 is the surface tension, θ is
the wetting angle, and r is the radius of the channel.
Invasions or displacements occur when they are energetically more favorable. More clearly,
the pressure difference between the wetting fluid (water) and the non-wetting fluid (oxygen) leads to
the formation of a curvature with a radius depending on the pressure difference. In the case of drainage,
the non-wetting phase is the one with higher pressure to be forced through the porous structure.

56
Processes 2020, 8, 362

The incremental increase of the pressure inside the gas phase results in the invasion of the liquid filled
pores and throats, with liquid pressures depending on their radius and wettability (Equation (1)). This
means, that the stepwise invasion of the interconnected pore space results in the formation of distinct
invasion patterns that depend on the PSD and the connectivity of pores and throats.
In such networks, the porosity can be increased basically in two ways, namely by (i) increasing
the number of throats of the same dimensions, and (ii) changing the distribution of the throat sizes
(Figure 2). In the example illustrated in Figure 3, porosity and mean throat diameter are kept constant
and only the distribution varies following the invasion pressure curve computed using Equation (1).
As can be seen, the differences in the structural organization of the PN are expected to result in different
overall liquid saturations and different gas–liquid distributions [53]. The larger pores and throats are
invaded by the gas phase while the smaller ones remain liquid saturated.

Figure 2. Different throat size distributions with same porosity but different mean throat diameter and
standard deviation of the throat size. Solid in white and liquid saturated void space (i.e., pores and
throats) in blue.

ȱ
Figure 3. Different saturations for constant porosity and constant mean throat diameter but different
organization of the pore network (PN). Solid and empty pores and throats in white and liquid saturated
pores and throats in blue.

3. Model Description
The model is comprised of two parts; part one includes the determination of the data structure
for the definition of the geometry of the void and solid space, and the second part contains the
equations of the drainage algorithm and the cluster labeling (Figure 4). The data structure contains
the information about the connections between the pores and throats in the network (Figure 5).
This information is used in the drainage algorithm for the stepwise calculation of the successive
invasion. The Hoshen–Kopelman algorithm [54,55] is then applied to identify invading and isolated
liquid clusters.

57
Processes 2020, 8, 362

ȱ
Figure 4. Scheme of the algorithm.

ȱ
Figure 5. Pore and throat numbering in a 2D PN. The dashed lines illustrate the periodicity at the
lateral boundaries.

3.1. Network Generation


Pore and throat radii (rp , rt ) are randomly distributed around a given average value with defined
standard deviations. The geometrical arrangement of throats and neighbor pores with the relevant
geometry parameters is illustrated in Figure 6.

Figure 6. Geometric information about pores and throats.

3.2. Invasion Algorithm


After the geometric parameters are specified, active, i.e., invading clusters with their menisci are
identified and the maximum liquid pressure is computed within the active clusters using Equation (1).
The algorithm then selects the largest accessible throat or pore for gas invasion following the rules
specified in Vorhauer et al. [46]. As invasion proceeds, entrapped clusters are formed, which are
permanently isolated due to the incompressibility of the fluids.

58
Processes 2020, 8, 362

3.3. Cluster Labeling


During this process, labeling of liquid clusters is used to identify the pores and throats that are
connected to each other and form a pathway for liquid and gas transport. At the start of the invasion
process, the network is completely occupied by liquid, and there is only one cluster spanning the
whole network and conducting the liquid phase. With initiation of invasion, numerous liquid clusters
can form that are distinguished into liquid-conducting clusters (connected to bottom and top side of
the PN) and isolated clusters. The Hoshen–Kopelman algorithm [54,55] is used for the labeling of
these clusters. Clusters are reviewed and relabeled after each invasion step to update the connections
between the pores and throats.

3.4. Model Assumptions


1. Quasi-static drainage invasion in the capillary dominated regime.
2. PN initially saturated with water.
3. No phase transition occurs.
4. Oxygen is injected at the top side and water is removed from the bottom side.
5. There is no mixing or diffusion between the two phases.
6. Viscous, gravity and liquid film flow are neglected.
7. Piston type throat invasion computed based on the Young–Laplace equation.
8. No further invasion occurs after breakthrough of the gas phase.

4. Pore Network Simulation of Microfluidic Experiments


A pore and throat network was conceived using the parameters of microfluidic experiments
from Arbabi et al. [20]. The PNM was constructed based on the image data extracted from Figure 7.
The experimental image in Figure 7 is then used to compare the flow path of gas within the micromodel
qualitatively with our own simulation results. In Figure 7, gas pores and throats are highlighted in
blue (pores) and yellow (throats) while liquid pores are in red and liquid throats in black. In this
investigation, we are interested if the PNM introduced above is able to predict the experimentally
estimated invasion path. It is remarked that invasion is dictated by the interface curvature of pore
and throat menisci, wherefore the pore and throat sizes are of interest here. They were determined
from the experimental image and transferred into the data structure of the PNM. Although the pore
sizes are significantly larger than the throat sizes in this example, pore invasion pressures were not
matched with the pore sizes. Instead the pore sizes were randomly adjusted. As can be seen below, the
simulation leads already to a very good agreement of the results. However, in a future study, it would
be preferable to track experimentally the different invasion pressure thresholds of pores and throats
based on the interface curvature of liquid menisci.

ȱ
Figure 7. Microfluidic drainage experiment from Arbabi et al. [20] (Reprinted with the permission from
Elsevier, 2014). Solids and gas invaded area in black, liquid in white. The PN is identified by the circles
and throats. The image shows the steady-state invasion pattern after breakthrough of the gas phase
from inlet (at the bottom) to water channel (at the top).

59
Processes 2020, 8, 362

The data of pore and throat sizes was implemented in the PNM to compute the successive invasion
of the PN. The result of the simulation is shown in Figure 8. It is observed that the invasion pattern
simulated with the PNM is identical with the experimental image (Figure 7). The perfect agreement
reveals the suitability of the model structure and model assumptions and the ability of PNMs to predict
the quasi-static drainage patterns. Though, in a future study it would be of interest, if also the stepwise
invasion of the PN can be accurately predicted, not only for idealized 2D structures but also in larger
and more realistic 3D structures.

Figure 8. (a) PNM simulation result, (b) invasion pattern comparison of simulation and experiment
result. Liquid-filled throats are shown in black, invaded pores in red, and invaded throats in blue.
Liquid-filled pores are not shown.

5. Monte Carlo Simulations

5.1. Impact of Pore Size Distribution in Monomodal PNs


The pore size properties of sintered PTL were used to study the effect of PSD on the invasion
patterns and the steady-state saturation of the PTL at breakthrough of the gas phase. For this purpose,
a pore network (PN) with the properties summarized in Table 2 were used.

Table 2. Simulation Parameters.

Parameter Value
Network size (columns and rows) 80 × 30
Temperature 80 ◦ C
Contact Angle 60◦
Surface Tension of water 0.0627 N/m
Avg. pore diameter 23 μm
Avg. throat diameter 17 μm
Lattice spacing 50 μm
Avg. throat length 27 μm
Porosity 63%

The standard deviation values were varied from 0.5 μm to 3 μm for a mean throat size of 17 μm
(Figure 9), and pore sizes were used with a constant standard deviation value of 2 μm. Monte Carlo
simulations were done for each data point so that the given gas saturation values in Figure 10 and
Table 3 are an average of 20 simulations. In general, it is observed that the final gas saturation increases
at breakthrough with widening of the radius distribution.

60
Processes 2020, 8, 362

Figure 9. Histograms of pore size distribution (PSD) with varying standard deviation in μm as indicated
in the legend.

Figure 10. Gas saturation at breakthrough of the gas phase for different PSDs. The mean value of throat
sizes is 17 μm.

Table 3. Total void volumes and porosities associated with the gas saturation at breakthrough.

Standard Deviation 0.5 μm 1.0 μm 1.5 μm 2.0 μm 2.5 μm 3.0 μm


Breakthrough gas saturation (%) 22.9 24.3 24.6 26.4 25.6 29
Porosity (%) 63.01 63.05 63.06 63.13 63.14 63.21
Total Void Volume (μL) 0.0178 0.0179 0.0180 0.0185 0.0191 0.0194

The simulation results clearly reveal the impact of PSD on the final distribution of liquid and
gas phase. While the average throat size was kept constant, the porosity of the PN increased slightly
with increasing standard deviation of throat sizes (Table 3). Thus, following the literature discussions
summarized above, it might be anticipated that higher porosities at constant mean throat or pore sizes
are a result of an increasing standard deviation of pore and throat sizes in the referenced situations. As
shown in Figure 10 and further analyzed below, the variation of PSD affects the invasion and thus the
gas saturation. In detail, a higher gas saturation is obtained for broader PSDs. It is to be noted that
higher PSDs than presented here can only be studied with a greater mean value of the throat size as
will be discussed below.

61
Processes 2020, 8, 362

5.2. Bimodal Pore Size Distributions


The previous results analyze the influence of the PSD on the example of monomodal PNs, i.e., only
one peak in the histograms in Figure 9. As can be seen from Table 3, the porosity is only marginally
affected in these cases. Due to this, the phase distribution patterns change only slightly in Figure 10. In
reality, pore structures often obey bimodal PSDs, i.e., with two peaks in the histogram (Figure 11) and
with much stronger impact on porosity. The influence of macro pores is highlighted in Figure 3.

Figure 11. Bimodal throat size distribution with standard deviation 2.0 μm for smaller (the first peak)
and 2.5 μm for larger (the second peak) throats.

As can be seen, for the monomodal PN in Figure 12, widespread invasion patterns with a high
number of invasions is not achieved. This is in contrast to the bimodal PN in Figure 13. Monte Carlo
simulations yielded an average gas saturation of 26% for the monomodal network and 38% for the
bimodal network. This shows that widening of the PSD (Figure 9) and the introduction of macro
pores (Figure 11), both, result in a change of the invasion process. This change is more significant in
the second situation. While capillary fingering with narrow single gas branches is rather favored by
narrow PSDs, widening of the invasion front with higher gas saturations is obtained by larger PSDs,
and larger porosities. However, it can also be shown that the widening of the front can also be achieved
when the porosity is kept constant and only the PSD is adjusted (Figure 13). The monomodal and
bimodal networks used in simulations have a constant porosity of 71%. Figures 12 and 13 also show
the saturation profiles of these simulations. They reveal the importance of the consideration of the PSD
for characterization of the invasion process.

ȱ
(a)ȱ
Figure 12. Cont.

62
Processes 2020, 8, 362

(b)ȱ
Figure 12. (a) Exemplary invasion patterns of the monomodal PN with porosity 71%. Liquid in blue,
gas in white and solid in gray. The arrow indicates the direction of gas invasion. (b) Saturation profiles
for different overall number of invaded throats and pores achieved during one drainage simulation of
the PN with randomly distributed pore and throat sizes.


(a)ȱ


(b)ȱ

Figure 13. (a) Exemplary invasion patterns of the bimodal PN with porosity 71%. Liquid in blue, gas in
white and solid in gray. Macro-pores are represented by thicker lines. The arrow indicates the direction
of gas invasion. (b) Saturation profiles for different overall number of invaded throats and pores from
one drainage simulation.

These findings are in very good agreement with literature predictions on drainage invasion
regimes [48,56,57]. According to [57], the invasion occurs in the capillary dominated regime, i.e., with

63
Processes 2020, 8, 362

rather low injection rate and negligible viscous forces in both fluids. The viscosity ratio of water/air is
around 50, and the capillary number is calculated from the ratio of viscous over capillary forces:

ΔPl
Ca = (2)
ΔPc

with liquid pressure difference ΔPl and capillary pressure gradient ΔPc . It depends on the viscosity
of the liquid phase η, the wetting curvature by means of surface tension σ, and the velocity of the
invading menisci v, as
8ηvLt
ΔPl = (3)
r2
And
2σσ0
ΔPc = 2 (4)
r
Ref. [58] considering the PSD with mean throat radius r and standard deviation σ0 . (In Equation
(3), Lt denotes the length of a throat). The conditions for capillary fingering are fulfilled, if the capillary
number is below 10−4 [48,59]. With the given values for viscosity, surface tension and PSD, this would
be achieved if the invasion velocity is below 3.3 mm/min (with liquid viscosity of water η = 10−3 Pa s
and surface tension σ = 0.073 N/m at 20 ◦ C, throat length Lt = 50 μm and standard deviation = 1.5 μm).
This is in good agreement with correlations of the invasion time and current density estimated in [46].
Transition to another invasion regime, e.g., stable displacement or viscous fingering [48] would occur
if the invasion velocity would be increased or also if the standard deviation of the throat size would be
decreased. Such effects are observed in the research of drying of PNs, where sufficiently narrow PSDs
can result in stable, i.e., viscosity-dominated invasion [58]. When viscosity comes into play, the width
of the invasion front scales with capillary number. However, in the absence of viscosity, the probability
of the throat invasion depends on the PSD. In the limit of identical throat sizes, all throats would be
equally selected for invasion. In the case of our simulation, where in a such a limit the selection would
not be stochastically distributed but ordered by the throat number, the invasion would always occur in
the throat with the lowest number (also refer to Figure 4) wherefore consequently this special situation
(i.e., no distribution of throat and pore sizes) could not be simulated with the current algorithm. In
contrast, when the throat size distribution becomes broader, the invasion follows the path of the least
resistance, which results in a more random distribution of the phase patterns, provided that the throat
sizes are randomly distributed. Following [60] and [61], the occupation probability decreases with
growing Δr.

5.3. Pore Network Simulations of Real Porous Structures


The above discussions are referred to rather artificial porous structures aiming at a more
fundamental understanding of the influence of the pore structure on the invasion behavior. In
the following, we would like to compare the results for realistic pore structures, extracted from felt,
foam and sintered PTLs. The simulation results are compared to literature values from Arbabi et al. [20].
For this purpose, the PN simulations were repeated with a single-entry point (similar as in Figure 7). In
more detail, the invasion starts from a single gas pore while all other surface pores are not connected to
the oxygen supply. Simulations for each type of PTL were repeated five times with the PSDs shown in
Figure 14. From these simulations, representative patterns, i.e., which were most frequently observed,
are shown in Figure 15. It is seen that felt and foam PTL show more constricted patterns, while the
sintered PTL allows relatively broad patterns with more gas invasions, which is explained well enough
by the assumed PSDs for each type (Figure 14).
It is remarked that the PSDs are usually not provided in literature [20]. Due to this, we have
selected a standard deviation of throat width so as to catch the gas–liquid phase distributions found
for microfluidic experiments in the literature. The agreement of the simulated and the experimental
invasion behavior is very well. Also, the results are in line with the discussion of the impact of PSDs

64
Processes 2020, 8, 362

above. Namely, the foam exhibits one capillary finger that invades the PN almost straightly. The felt
PTL, with a slightly broader PSD (Figure 15), instead reveals a slight widening of the invasion front in
horizontal direction. In contrast to that, the sintered PTL allows for a broadening of the invasion front
and a significantly higher gas saturation at breakthrough.
In addition to that, it becomes clear from Figure 15 that the effect of liquid clustering occurs in all
types of the PTL. The greatest number of isolated liquid clusters is observed in the sintered PTL where
the number of invaded pores is higher and the invasion front appears more ramified. In the case of felt
and foam PTL, isolated clusters are not seen on the experimental images. This could be because of the
small size of the PN and the overall lower number of pore invasions.

Figure 14. PSDs used for different PTL types with standard deviations: 1.0 μm for foam, 1.7 μm for felt
and 2.5 μm for sintered.

ȱ
Figure 15. Invasion pattern comparison: (a) foam PTL, (b) felt PTL, (c) sintered PTL (Experimental
images from [20] are reprinted with the permission from Elsevier, 2014).

6. Summary and Conclusions


In this study, the applicability of PMNs for the simulation of water drainage from PTL was
discussed. The PNM applies invasion percolation rules for hydrophilic drainage. It was shown that the
PNM can reliably predict the invasion patterns of 2D microfluidic devices related to water electrolysis

65
Processes 2020, 8, 362

studies. In addition to that, the structure dependence of the invasion process was studied using various
types of PTLs (foam, felt, sintered). In agreement with experimental data from literature, we could
predict different invasion patterns for the three cases. Felt and foam PTLs showed narrow gas fingers
while sintered PTL showed widespread invasion front patterns with a higher number of trapped
clusters. This observation matches the theoretical investigations based on a Monte Carlo study of
PSDs and invasion probability. Based on this, it could furthermore be shown, that the PSD affects
the invasion patterns more significantly than the porosity. More clearly, it was revealed that the PSD
could be adjusted independently of the porosity and that this resulted in different invasion patterns.
The impact of PSD can also be extended towards porous media in fuel cells.
As a next step, the PN will be further enhanced to replicate the local coordination numbers in
real PTLs with non-uniform distribution of pores and throats. The purpose would be to simulate
mass transfer within the real structure of porous media rather than in an idealized network. These
simulations would then be much closer to reality compared to the ones with a fixed coordination
number in the network. The effective transport parameters, e.g., permeability, could also then be
extracted by solving mass transfer equations pore by pore in a real structure. It would also be important
to study the effect of local temperature changes in the system and liquid flow through corner films.
Unsteady changes in the current density could also lead to pressure changes. For this reason, the
application of imbibition rules along with drainage will become important.

Author Contributions: Conceptualization, N.V. and T.V.-K.; methodology, N.V.; software, H.A.; validation,
H.A.; formal analysis, N.V. and T.V.-K.; investigation, H.A.; resources, N.V. and E.T.; data curation, H.A.;
writing—original draft preparation, H.A.; writing—review and editing, N.V.; visualization, H.A.; supervision, N.V.
and T.V.-K.; project administration, N.V., T.V.-K. and E.T.; funding acquisition, N.V., T.V.-K. and E.T. All authors
have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.

Symbols
L Distance between nodes
Lt Length of throat
Pc Capillary pressure
Pl Liquid pressure
rp Pore radius
rt Throat radius
r Mean radius
v Velocity of invading menisci
Vp Volume of pore
Vt Volume of throat
z PTL space coordinate
η Viscosity of liquid phase
θ Contact angle
σ Surface tension
σ0 Standard deviation
Abbreviations
Ca Capillary number
PN Pore network
PNM Pore network model
PSD Pore size distribution
PTL Porous transport layer
Ti Titanium

66
Processes 2020, 8, 362

References
1. Das, L.M. Hydrogen-fueled internal combustion engines. In Compendium of Hydrogen Energy; Elsevier:
Amsterdam, The Netherlands, 2016; pp. 177–217.
2. Ursua, A.; Gandia, L.M.; Sanchis, P. Hydrogen Production From Water Electrolysis: Current Status and
Future Trends. Proc. IEEE 2012, 100, 410–426. [CrossRef]
3. Balat, M. Potential importance of hydrogen as a future solution to environmental and transportation problems.
Int. J. Hydrog. Energy 2008, 33, 4013–4029. [CrossRef]
4. Chen, Y.; Hu, X.; Liu, J. Life Cycle Assessment of Fuel Cell Vehicles Considering the Detailed Vehicle
Components: Comparison and Scenario Analysis in China Based on Different Hydrogen Production Schemes.
Energies 2019, 12, 3031. [CrossRef]
5. Elgowainy, A.; Gaines, L.; Wang, M. Fuel-cycle analysis of early market applications of fuel cells: Forklift
propulsion systems and distributed power generation. Int. J. Hydrog. Energy 2009, 34, 3557–3570. [CrossRef]
6. Felder, F.A.; Hajos, A. Using Restructured Electricity Markets in the Hydrogen Transition: The PJM Case.
Proc. IEEE 2006, 94, 1864–1879. [CrossRef]
7. Shaw, S.; Peteves, E. Exploiting synergies in European wind and hydrogen sectors: A cost-benefit assessment.
Int. J. Hydrog. Energy 2008, 33, 3249–3263. [CrossRef]
8. Carmo, M.; Fritz, D.L.; Mergel, J.; Stolten, D. A comprehensive review on PEM water electrolysis. Int. J.
Hydrog. Energy 2013, 38, 4901–4934. [CrossRef]
9. Zoulias, E.I.; Glockner, R.; Lymberopoulos, N.; Tsoutsos, T.; Vosseler, I.; Gavalda, O.; Mydske, H.J.; Taylor, P.
Integration of hydrogen energy technologies in stand-alone power systems analysis of the current potential
for applications. Renew. Sustain. Energy Rev. 2006, 10, 432–462. [CrossRef]
10. Degiorgis, L.; Santarelli, M.; Calì, M. Hydrogen from renewable energy: A pilot plant for thermal production
and mobility. J. Power Sources 2007, 171, 237–246. [CrossRef]
11. Kosonen, A.; Koponen, J.; Huoman, K.; Ahola, J.; Ruuskanen, V.; Ahonen, T.; Graf, T. Optimization strategies
of PEM electrolyser as part of solar PV system. In Proceedings of the IEEE 18th European Conference
on Power Electronics and Applications (EPE’16 ECCE Europe), Karlsruhe, Germany, 5–9 September 2016;
pp. 1–10.
12. Dedigama, I.; Angeli, P.; Ayers, K.; Robinson, J.B.; Shearing, P.R.; Tsaoulidis, D.; Brett, D.J.L. In situ diagnostic
techniques for characterisation of polymer electrolyte membrane water electrolysers—Flow visualisation
and electrochemical impedance spectroscopy. Int. J. Hydrog. Energy 2014, 39, 4468–4482. [CrossRef]
13. Nieminen, J.; Dincer, I.; Naterer, G. Comparative performance analysis of PEM and solid oxide steam
electrolysers. Int. J. Hydrog. Energy 2010, 35, 10842–10850. [CrossRef]
14. Grigoriev, S.A.; Millet, P.; Volobuev, S.A.; Fateev, V.N. Optimization of porous current collectors for PEM
water electrolysers. Int. J. Hydrog. Energy 2009, 34, 4968–4973. [CrossRef]
15. Ito, H.; Maeda, T.; Nakano, A.; Hwang, C.M.; Ishida, M.; Kato, A.; Yoshida, T. Experimental study on porous
current collectors of PEM electrolyzers. Int. J. Hydrog. Energy 2012, 37, 7418–7428. [CrossRef]
16. Mo, J.; Kang, Z.; Yang, G.; Li, Y.; Retterer, S.T.; Cullen, D.A.; Toops, T.J.; Bender, G.; Pivovar, B.S.;
Green, J.B., Jr.; et al. In situ investigation on ultrafast oxygen evolution reactions of water splitting in proton
exchange membrane electrolyzer cells. J. Mater. Chem. A 2017, 5, 18469–18475. [CrossRef]
17. Lee, C.H.; Banerjee, R.; Arbabi, F.; Hinebaugh, J.; Bazylak, A. Porous Transport Layer Related Mass
Transport Losses in Polymer Electrolyte Membrane Electrolysis: A Review. In Proceedings of the ASME 14th
International Conference on Nanochannels, Microchannels, and Minichannels, Washington, DC, USA, 10–14
July 2016. [CrossRef]
18. Ito, H.; Maeda, T.; Nakano, A.; Kato, A.; Yoshida, T. Influence of pore structural properties of current
collectors on the performance of proton exchange membrane electrolyzer. Electrochim. Acta 2013, 100,
242–248. [CrossRef]
19. Lee, C.; Hinebaugh, J.; Banerjee, R.; Chevalier, S.; Abouatallah, R.; Wang, R.; Bazylak, A. Influence of limiting
throat and flow regime on oxygen bubble saturation of polymer electrolyte membrane electrolyzer porous
transport layers. Int. J. Hydrog. Energy 2017, 42, 2724–2735. [CrossRef]
20. Arbabi, F.; Kalantarian, A.; Abouatallah, R.; Wang, R.; Wallace, J.S.; Bazylak, A. Feasibility study of using
microfluidic platforms for visualizing bubble flows in electrolyzer gas diffusion layers. J. Power Sour. 2014,
258, 142–149. [CrossRef]

67
Processes 2020, 8, 362

21. Nie, J.; Chen, Y. Numerical modeling of three-dimensional two-phase gas–liquid flow in the flow field plate
of a PEM electrolysis cell. Int. J. Hydrog. Energy 2010, 35, 3183–3197. [CrossRef]
22. Ito, H.; Maeda, T.; Nakano, A.; Hasegawa, Y.; Yokoi, N.; Hwang, C.M.; Ishida, M.; Kato, A.; Yoshida, T. Effect
of flow regime of circulating water on a proton exchange membrane electrolyzer. Int. J. Hydrog. Energy 2010,
35, 9550–9560. [CrossRef]
23. Badwal, S.P.S.; Giddey, S.; Munnings, C. Emerging technologies, markets and commercialization of
solid-electrolytic hydrogen production. WIRE Energy Environ. 2018, 7, e286. [CrossRef]
24. Babic, U.; Suermann, M.; Büchi, F.N.; Gubler, L.; Schmidt, T.J. Critical Review—Identifying Critical Gaps for
Polymer Electrolyte Water Electrolysis Development. J. Electrochem. Soc. 2017, 164, F387–F399. [CrossRef]
25. Yigit, T.; Selamet, O.F. Mathematical modeling and dynamic Simulink simulation of high-pressure PEM
electrolyzer system. Int. J. Hydrog. Energy 2016, 41, 13901–13914. [CrossRef]
26. Kang, Z.; Mo, J.; Yang, G.; Retterer, S.T.; Cullen, D.A.; Toops, T.J.; Green, J.B., Jr.; Mench, M.M.; Zhang, F.-Y.
Investigation of thin/well-tunable liquid/gas diffusion layers exhibiting superior multifunctional performance
in low-temperature electrolytic water splitting. Energy Environ. Sci. 2017, 10, 166–175. [CrossRef]
27. Sadeghi Lafmejani, S.; Olesen, A.C.; Kaer, S.K. Analysing Gas-Liquid Flow in PEM Electrolyser
Micro-Channels. ECS Trans. 2016, 75, 1121–1127. [CrossRef]
28. Selamet, O.F.; Pasaogullari, U.; Spernjak, D.; Hussey, D.S.; Jacobson, D.L.; Mat, M. In Situ Two-Phase Flow
Investigation of Proton Exchange Membrane (PEM) Electrolyzer by Simultaneous Optical and Neutron
Imaging. ECS Trans. 2011, 41, 349–362.
29. Ojong, E.T.; Kwan, J.T.H.; Nouri-Khorasani, A.; Bonakdarpour, A.; Wilkinson, D.P.; Smolinka, T. Development
of an experimentally validated semi-empirical fully-coupled performance model of a PEM electrolysis cell
with a 3-D structured porous transport layer. Int. J. Hydrog. Energy 2017, 42, 25831–25847. [CrossRef]
30. Abdol Rahim, A.H.; Tijani, A.S.; Kamarudin, S.K.; Hanapi, S. An overview of polymer electrolyte membrane
electrolyzer for hydrogen production: Modeling and mass transport. J. Power Sour. 2016, 309, 56–65.
[CrossRef]
31. Aubras, F.; Deseure, J.; Kadjo, J.-J.A.; Dedigama, I.; Majasan, J.; Grondin-Perez, B.; Chabriat, J.-P.; Brett, D.J.L.
Two-dimensional model of low-pressure PEM electrolyser: Two-phase flow regime, electrochemical modelling
and experimental validation. Int. J. Hydrog. Energy 2017, 42, 26203–26216. [CrossRef]
32. Olesen, A.C.; Rømer, C.; Kær, S.K. A numerical study of the gas-liquid, two-phase flow maldistribution in
the anode of a high pressure PEM water electrolysis cell. Int. J. Hydrog. Energy 2016, 41, 52–68. [CrossRef]
33. Lee, J.K.; Lee, C.; Bazylak, A. Pore network modelling to enhance liquid water transport through porous
transport layers for polymer electrolyte membrane electrolyzers. J. Power Sour. 2019, 437, 226910. [CrossRef]
34. Zhang, H.; Lin, G.; Chen, J. Evaluation and calculation on the efficiency of a water electrolysis system for
hydrogen production. Int. J. Hydrog. Energy 2010, 35, 10851–10858. [CrossRef]
35. Hwang, C.M.; Ishida, M.; Ito, H.; Maeda, T.; Nakano, A.; Hasegawa, Y.; Yokoi, N.; Kato, A.; Yoshida, T.
Influence of properties of gas diffusion layers on the performance of polymer electrolyte-based unitized
reversible fuel cells. Int. J. Hydrog. Energy 2011, 36, 1740–1753. [CrossRef]
36. Han, B.; Mo, J.; Kang, Z.; Yang, G.; Barnhill, W.; Zhang, F.-Y. Modeling of two-phase transport in proton
exchange membrane electrolyzer cells for hydrogen energy. Int. J. Hydrog. Energy 2017, 42, 4478–4489.
[CrossRef]
37. Lee, C.; Zhao, B.; Abouatallah, R.; Wang, R.; Bazylak, A. Compressible-Gas Invasion into Liquid-Saturated
Porous Media: Application to Polymer-Electrolyte-Membrane Electrolyzers. Phys. Rev. Appl. 2019, 11, 054029.
[CrossRef]
38. Mo, J.; Dehoff, R.R.; Peter, W.H.; Toops, T.J.; Green, J.B.; Zhang, F.-Y. Additive manufacturing of liquid/gas
diffusion layers for low-cost and high-efficiency hydrogen production. Int. J. Hydrog. Energy 2016, 41,
3128–3135. [CrossRef]
39. Lettenmeier, P.; Kolb, S.; Sata, N.; Fallisch, A.; Zielke, L.; Thiele, S.; Gago, A.S.; Friedrich, K.A. Comprehensive
investigation of novel pore-graded gas diffusion layers for high-performance and cost-effective proton
exchange membrane electrolyzers. Energy Environ. Sci. 2017, 10, 2521–2533. [CrossRef]
40. Kang, Z.; Mo, J.; Yang, G.; Li, Y.; Talley, D.A.; Retterer, S.T.; Cullen, D.A.; Toops, T.J.; Brady, M.P.; Bender, G.; et al.
Thin film surface modifications of thin/tunable liquid/gas diffusion layers for high-efficiency proton exchange
membrane electrolyzer cells. Appl. Energy 2017, 206, 983–990. [CrossRef]

68
Processes 2020, 8, 362

41. Schuler, T.; Schmidt, T.J.; Büchi, F.N. Polymer Electrolyte Water Electrolysis: Correlating Performance and
Porous Transport Layer Structure: Part II. Electrochemical Performance Analysis. J. Electrochem. Soc. 2019,
166, F555–F565. [CrossRef]
42. Prat, M.; Agaësse, T. Thin Porous Media. In Handbook of Porous Media, 3rd ed.; Vafai, K., Ed.; CRC Press: Boca
Raton, FL, USA, 2015; pp. 89–112.
43. Raeini, A.Q.; Yang, J.; Bondino, I.; Bultreys, T.; Blunt, M.J.; Bijeljic, B. Validating the Generalized Pore Network
Model Using Micro-CT Images of Two-Phase Flow. Transp. Porous. Med. 2019, 130, 405–424. [CrossRef]
44. Tranter, T.G.; Gostick, J.T.; Burns, A.D.; Gale, W.F. Pore Network Modeling of Compressed Fuel Cell
Components with OpenPNM. Fuel Cells 2016, 16, 504–515. [CrossRef]
45. Metzger, T. A personal view on pore network models in drying technology. Dry. Technol. 2019, 37, 497–512.
[CrossRef]
46. Vorhauer, N.; Altaf, H.; Tsotsas, E.; Vidakovic-Koch, T. Pore Network Simulation of Gas-Liquid Distribution
in Porous Transport Layers. Processes 2019, 7, 558. [CrossRef]
47. Joekar-Niasar, V.; Hassanizadeh, S.M. Analysis of Fundamentals of Two-Phase Flow in Porous Media Using
Dynamic Pore-Network Models: A Review. Crit. Rev. Environ. Sci. Technol. 2012, 42, 1895–1976. [CrossRef]
48. Sinha, P.K.; Wang, C.-Y. Pore-network modeling of liquid water transport in gas diffusion layer of a polymer
electrolyte fuel cell. Electrochim. Acta 2007, 52, 7936–7945. [CrossRef]
49. Sinha, P.K.; Wang, C.-Y. Liquid water transport in a mixed-wet gas diffusion layer of a polymer electrolyte
fuel cell. Chem. Eng. Sci. 2008, 63, 1081–1091. [CrossRef]
50. Lee, K.-J.; Nam, J.H.; Kim, C.-J. Pore-network analysis of two-phase water transport in gas diffusion layers of
polymer electrolyte membrane fuel cells. Electrochim. Acta 2009, 54, 1166–1176. [CrossRef]
51. Bultreys, T.; Singh, K.; Raeini, A.Q.; Ruspini, L.C.; Øren, P.-E.; Berg, S.; Rücker, M.; Bijeljic, B.;
Blunt, M.J. Verifying pore network models of imbibition in rocks using time-resolved synchrotron imaging.
EarthArXiv 2019. [CrossRef]
52. Vorhauer, N.; Tran, Q.T.; Metzger, T.; Tsotsas, E.; Prat, M. Experimental Investigation of Drying in a Model
Porous Medium: Influence of Thermal Gradients. Dry. Technol. 2013, 31, 920–929. [CrossRef]
53. Metzger, T.; Irawan, A.; Tsotsas, E. Influence of pore structure on drying kinetics: A pore network study.
AIChE J. 2007, 53, 3029–3041. [CrossRef]
54. Al-Futaisi, A.; Patzek, T.W. Extension of Hoshen–Kopelman algorithm to non-lattice environments. Phys. A
Stat. Mech. Appl. 2003, 321, 665–678. [CrossRef]
55. Hoshen, J. On the application of the enhanced Hoshen–Kopelman algorithm for image analysis. Pattern
Recognit. Lett. 1998, 19, 575–584. [CrossRef]
56. Xu, B.; Yortsos, Y.C.; Salin, D. Invasion percolation with viscous forces. Phys. Rev. E 1998, 57, 739–751.
[CrossRef]
57. Lenormand, R. Liquids in porous media. J. Fluid Mech. 1990, 2, SA79–SA88. [CrossRef]
58. Metzger, T.; Tsotsas, E.; Prat, M. Pore-Network Models: A Powerful Tool to Study Drying at the Pore Level
and Understand the Influence of Structure on Drying Kinetics. In Modern Drying Technology: Tools at Different
Scales; Tsotsas, E., Mujumdar, A.S., Eds.; Wiley: Weinheim, Germany, 2007; pp. 57–102.
59. Lenormand, R.; Touboul, E.; Zarcone, C. Numerical models and experiments on immiscible displacements in
porous media. J. Fluid Mech. 1988, 189, 165–187. [CrossRef]
60. Prat, M.; Bouleux, F. Drying of capillary porous media with a stabilized front in two dimensions. Phys. Rev.
E 1999, 60, 5647–5656. [CrossRef]
61. Vorhauer, N.; Tsotsas, E.; Prat, M. Temperature gradient induced double stabilization of the evaporation
front within a drying porous medium. Phys. Rev. Fluids 2018, 3, 217. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

69
processes
Article
Extreme Learning Machine-Based Model
for Solubility Estimation of Hydrocarbon Gases
in Electrolyte Solutions
Narjes Nabipour 1 , Amir Mosavi 2,3,4,5 , Alireza Baghban 6 , Shahaboddin Shamshirband 7,8, *
and Imre Felde 9
1 Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam;
[email protected]
2 Kando Kalman Faculty of Electrical Engineering, Obuda University, 1034 Budapest, Hungary;
[email protected]
3 School of Built the Environment, Oxford Brookes University, Oxford OX30BP, UK
4 Faculty of Health, Queensland University of Technology, 130 Victoria Park Road,
Brisbane, QLD 4059, Australia
5 Institute of Structural Mechanics, Bauhaus Universität-Weimar, D-99423 Weimar, Germany
6 Chemical Engineering Department, Amirkabir University of Technology, Mahshahr Campus,
Mahshahr, Iran; [email protected]
7 Department for Management of Science and Technology Development, Ton Duc Thang University,
Ho Chi Minh City, Vietnam
8 Faculty of Information Technology, Ton Duc Thang University, Ho Chi Minh City, Vietnam
9 John von Neumann Faculty of Informatics, Obuda University, 1034 Budapest, Hungary; [email protected]
* Correspondence: [email protected]

Received: 26 November 2019; Accepted: 7 January 2020; Published: 9 January 2020

Abstract: Calculating hydrocarbon components solubility of natural gases is known as one of


the important issues for operational works in petroleum and chemical engineering. In this work,
a novel solubility estimation tool has been proposed for hydrocarbon gases—including methane,
ethane, propane, and butane—in aqueous electrolyte solutions based on extreme learning machine
(ELM) algorithm. Comparing the ELM outputs with a comprehensive real databank which has
1175 solubility points yielded R-squared values of 0.985 and 0.987 for training and testing phases
respectively. Furthermore, the visual comparison of estimated and actual hydrocarbon solubility
led to confirm the ability of proposed solubility model. Additionally, sensitivity analysis has been
employed on the input variables of model to identify their impacts on hydrocarbon solubility. Such a
comprehensive and reliable study can help engineers and scientists to successfully determine the
important thermodynamic properties, which are key factors in optimizing and designing different
industrial units such as refineries and petrochemical plants.

Keywords: hydrocarbon gases; solubility; natural gas; extreme learning machines; electrolyte solution;
prediction model; big data; data science; deep learning; chemical process model; machine learning

1. Introduction
Solubility of hydrocarbon and nonhydrocarbon gases—i.e., mixtures of methane, ethane, propane,
CO2 , and N2 in aqueous phases—is known as one of the important practical and theoretical challenges
in petroleum, geochemical, and chemical engineering. This property has an effective role in different
processes, such as achieving optimum conditions for oil and gas transportation, gas hydrate formation,
designing thermal separation processes, gas sequestration for protecting environment, and coal
gasification. Petroleum reservoirs normally have some natural gases with aqueous solution at

Processes 2020, 8, 92; doi:10.3390/pr8010092 71 www.mdpi.com/journal/processes


Processes 2020, 8, 92

high-pressure and high-temperature conditions so that the solubility of gas becomes attractive for
engineers [1–8]. In production and transportation of hydrocarbons, it is possible that water content
of gas undergoes an alteration in phase from vapor to ice and gas hydrates. The crystalline solid
phases called gas hydrates are created when small-sized gas molecules are trapped in lattice of water
molecules. Creation of hydrates can cause major flow assurance problems during production and
transportation of hydrocarbons steps such as pipeline blockage, corrosion, and many other issues
resulted from the two-phase flow [1,9–11].
In the recent years, investigations on CO2 solubility in aqueous electrolyte solutions have
grown significantly as well as they are related to CO2 capture and storage. It is a clear fact that
the dominant cause of global warming is emission of CO2 gas generated from fossil fuels so its
sequestration and disposal in the ocean have been known as a reasonable choice to overcome global
warming problems [12–14]. Simulation of enhanced oil recovery, design of supercritical extraction,
and optimization of CO2 dissolution in the ocean need a comprehensive knowledge about carbon
dioxide solubility in aqueous electrolytes solutions [13–15].
Investigation of natural gas phase behavior in aqueous solutions in different operational conditions
is known one of the important issues in the industry, which has wide applications for avoiding
problems in designing and optimization of gas processing. In the literature, there are different solubility
datasets for various gas–liquid systems. These datasets mostly include hydrocarbons’ dissolution
in water/brine systems [1,4,5,9,16–20] and non-hydrocarbons such as CO2 and N2 dissolution in
water/brine systems [7,12–14,18,21–24]. A brief summary of the hydrocarbon systems datasets is
shown in Table 1 for hydrocarbons. The experimental data of water content of hydrocarbons and
non-hydrocarbons are limited because of difficulties in measurement of the low water content gases
at high pressure and low temperature. Mohammadi and coworkers expressed that an accurate
estimation of water content can be obtained by gas solubility data, therefore, they overcame the
complexities of experimental determination of the water content in natural gases [1]. Due to limited
number of measurement data, wide attempts have been made to model and describe the gas-liquid
equilibrium in aqueous electrolyte solutions. There are several thermodynamic models which uses the
Henry’s constant, activity coefficient, and cubic equations of state to obtain more information about
the equilibrium conditions. The changes of Henry’s constant for the pressure lower than 5 MPa are
negligible and it is dominantly affected by temperatures [19]. The high dependency on temperature
is obvious at low temperature and also the nonlinear decreasing relationship is observed at high
temperatures [25]. Furthermore, there is just a limited number of Henry’s constants for hydrocarbon
systems at low temperature. According to this fact, there are several drawbacks in applying the
Henry’s law, whereas it has great ability for accurate prediction of solubility. As an example, it is
suitable for dilute solutions or near-ideal solutions [26]. Additionally, this method is correct for
single compounds in no chemical reaction conditions for aqueous phase. Another method is cubic
EOS which has several advantages such as small number of parameters, computational efficiency,
and ease of performance [3,4,21]. The EOSs were proposed originally for pure fluids, after that, their
applications were expanded for mixtures by combining the constants from different pure components.
This extension can be done by different methods such as Dalton’s law of additive partial pressures
and Amagat’s rule of additive volumes [5]. For complex compounds, there are some limitations in
accuracy of EOS which highlight the importance of empirical adjustments by dealing with the binary
interaction parameters. In order to determine these parameters, a reliable source of experimental data
for vapor-liquid equilibrium is required which induces some uncertainty into EOSs [7].
Due to above discussions, development of an accurate and reliable approach for estimation of
solubility of hydrocarbons and non-hydrocarbons in aqueous electrolyte solutions has been highlighted.
Nowadays, machine learning approaches have shown extensive applications in different topics [27–35].
This work organizes a novel artificial intelligence method called extreme learning machine (ELM) to
estimate solubility of hydrocarbons in aqueous electrolyte mixtures in terms of types of gas, mole
fractions of gases, pressure, temperature, and ionic strength.

72
Processes 2020, 8, 92

Table 1. Details of experimental hydrocarbons solubility in aqueous electrolyte solutions.

Mole Fraction of the Components


Author P (Mpa) T (◦ C) Composition
in the Gaseous Phase
Culberson et al. 0.8–69.61 37.78–171.11 Pure water C1: 0.0000698–0.0033
Pure water, LiBr, KBr,
Kiepe et al. 0.304–10.23 40–100.14 C1: 0.00003–0.00154
LiCl, KCl
C1: 0.000204–0.002459
C2: 0.0000147–0.0000674
Chapoy et al. 0.357–18 1.98–95.01 Pure water
C3: 0.0000321–0.0002694
C4: 0.00000387–0.00001121
C1: 0.00099–0.00282
Marinakis et al. 6.22–20.1 1.4–25.98 Pure water, NaCl C2: 0.000038–0.000249
C3: 0.000006–0.000042
Crovetto et al. 1.327–6.451 24.35–245.15 Pure water C1: 0.0002124–0.0010337
C1: 0.000563–0.004049
Wang et al. 1–40.03 2.5–30.05 Pure water
C2: 0.0000986–0.000864
C1: 0.00045–0.0037
Amirjafari 4.66–56.16 54.44–104.44 Pure water C2: 0.000119–0.001768
C3: 1.9 × 10−5 –0.001863
C1: 0.000805–0.0043
O’Sullivan et al. 10.2–62 51.5–125 Pure water, NaCl
C2: 0.000825–0.001438
Pure water, NaCl, LiCl,
Michels et al. 4.09–45.89 25–150 C1: 0.000173–0.00269
NaBr, NaI, CaCl2
Mohammadi et al. 1.14–31.1 4.65–24.75 Pure water C1: 0.000313–0.00311
Vul’fson et al. 2.53–60.8 19.95–79.95 Pure water C1: 0.000361–0.004328
C1: 0.000127–0.005085
Dhima 2.5–100 71 Pure water C2: 0.000821–0.001398
C4: 0.000021–0.000103

2. Methodology

2.1. Experimental Dataset Collection


In order to construct a highly accurate and comprehensive model capable of estimating the
solubility of mixtures of hydrocarbons in aqueous electrolyte solutions, a comprehensive databank
was provided based on existing experimental data in Table 1. This databank contains total number of
1175 solubility points for hydrocarbons (881 and 294 points for training and testing phases, respectively)
(see Table S1 of data set in Supplementary Materials). According to the literature [1,4,5,9,16–20],
the solubility of gases in these systems is highly function of aqueous solutions, pressure, temperature,
and gaseous phase composition. The aqueous phase composition was change into ionic strength (I)
from salt concentrations to reduce dimensions of modeling process. The following equation presents
the relationships between ionic strength, valance of charged ions (zi ), and molar concentration of each
ion (mi ).
1
I= mi |zi |2 (1)
2
In this study, the solubility of hydrocarbons is predicted in terms of concentration of components
in gaseous mixture, ionic strength of solution, temperature, and pressure.

ηh = f (C1, C2, C3, C4, I, P, T, idx) (2)

In which, ηh represents the hydrocarbon solubility in aqueous phase; C(1–4) are known as the
methane, ethane, propane, and butane mole fraction in gas phase (0–99.99); I denotes the ionic strength
based on molarity (0–37.35); T denotes the temperature in terms of ◦ C (1.4–245.15); P shows the pressure
in MPa (0.3–100), and idx symbolizes the index of fraction whose solubility is to be determined (1,2,3,4).

73
Processes 2020, 8, 92

2.2. Extreme Learning Machine


Huang proposed a new intelligence method based on single-layer feedforward neural network
(SLFFNN) called extreme learning machine to satisfy the drawbacks of gradient-based algorithms
such low training speed and low learning rate. In the ELM algorithm, the hidden nodes are selected
randomly and the weights of output of the SLFFNN are calculated by applying Moore–Penrose
generalized inverse [36,37].
The scheme of ELM algorithm is demonstrated in Figure 1. By assuming N training sets such as
(xi , yi ) ∈ Rn × Rm for L hidden nodes, the SLFFNN algorithms can be written as

L 

i=1 βi fi x j = Li=1 βi fi (ai .bi .x) j = 1, . . . , N (3)

In which, ai = [ai1 , . . . , ain ]T points to input weights matrix which is related to hidden nodes,
βi = [βi1 , . . . , βim ]T represents the output weights matrix which is related to hidden nodes, and bi
symbolizes the hidden layer bias. 

L
i=1 βi fi x j = Hβ (4)

In which, β = [β1 , . . . , βL ] and h(x) = [h1 (x), . . . hL (x)] are known as the hidden layer output matrix
and the output weight matrix.
The first step of this model is the random calculation of input weight and the bias of hidden
layer for the training phase. Then, for determining these values, the hidden layer matrix is obtained
by utilization of input variables. Then, the SLFFNN training is changed to a least-square problem.
The ELM algorithms implement regularization theory to define a target function as [38–40]

1  2 c  2


minLELM = β + T − Hβ (5)
2 2

Figure 1. Structure of ELM algorithm.

3. Results and Discussion


In this study, the solubility of hydrocarbons in the aqueous electrolyte phase is determined based
on ELM algorithm. To this end, the sigmoid function is set as activation function and the input weights
were initialized randomly in range of (−1, 1). Additionally, the number of nodes in the hidden layers
was estimated as 30 based on the lowest value of RMSE as determined in Figure 2. As shown, after

74
Processes 2020, 8, 92

30 nodes, by increasing complexity of model, the testing error increased so the optimum structure of
the algorithm has 30 nodes to prohibit overfitting.

Figure 2. Obtaining optimum structure of proposed algorithm.

In the following, the statistical results of the estimation of hydrocarbon solubility are inserted in
Table 2. The following equations are used to achieve this end:

 actual − Xpredicted
100
N Xi i
Mean relative error MRE = ( (6)
N i=1 Xactual i

1 N  actual
predicted 2
Root mean square error (RMSE) = ( Xi − Xi ) (7)
N i=1

1  actual
N
predicted 2
Mean squared error (MSE) = Xi − Xi (8)
N
i=1

N 
predicted 2
Xiactual − Xi
i=1
R − squared (R2) = 1 − (9)

N  actual 2
i = 1 Xi − Xactual

As shown in Table 2, the MRE, MSE, and RMSE are determined as 22.049, 1.33285 × 10−8 , and 0.0001
for training phase respectively. Moreover, for testing phase, MRE = 22.054, MSE = 1.05351 × 10−8 and
RMSE = 0.0001 are calculated. The estimated R2 values are 0.985, 0.987, and 0.985 for training, testing,
and overall datasets respectively. These results give the knowledge about the high degree of accuracy
for proposed ELM algorithm.

75
Processes 2020, 8, 92

Table 2. Statistical analyses of developed model.

Dataset R2 MRE (%) MSE RMSE


Training 0.985 22.049 1.33285 × 10−8 0.0001
Testing 0.987 22.054 1.05351 × 10−8 0.0001
Overall 0.985 22.050 1.26295 × 10−8 0.0001

On the one hand, the comparison between the estimated and real hydrocarbons solubility in
aqueous electrolyte solutions are shown in Figure 3. This depiction demonstrates an excellent agreement
between estimated and real solubility values. Figure 4 also represents the regression plot of actual
hydrocarbons solubility versus estimated one. A light cloud of data near the 45◦ line expresses the
validity and accuracy of ELM algorithm. Additionally, Figure 5 also shows the distribution of relative
deviations between forecasted and actual hydrocarbons solubility in aqueous solutions. It can be seen
that the ELM outputs deviate slightly from the real solubility and most of relative deviations are near
to zero. Furthermore, Figure 6 shows the histograms of relative deviations for training and testing
phases. In this demonstration, frequency diagram confirms that most of the error points are close to
zero and also cumulative axis express the fact that range of deviation is very limited and the highest
slope of the cumulative curve occurred near the zero point.

Figure 3. Comparison of actual and estimated solubility of hydrocarbons.

Figure 4. Cross plot of actual and estimated solubility of hydrocarbons.

76
Processes 2020, 8, 92

Figure 5. Relative deviation between actual and estimated solubility of hydrocarbons.

Figure 6. Histogram diagram of relative deviations.

77
Processes 2020, 8, 92

The ELM algorithm implemented in the current work shows an excellent ability in calculation of
solubility of hydrocarbons in aqueous phases. One of the important factors which can influence the
validation of model is degree of precision of utilized data. In order to clarify the accuracy of solubility
databank, the leverage mathematical method is recruited. This method has some rules to identify the
suspected solubility data so that a matrix which is known as hat matrix, should be constructed based
on formulation [41–45]
 −1
H = U UT U UT (10)

In which, U symbolizes a matrix of i × j dimensional. i and j are known as the number of algorithm
parameter and training points which are used for determination of critical leverage limit as

H∗ = 3( j + 1)/i (11)

In order to detect the reliable zone, there are two standard residual indexes (−3 and 3) which
are used in the leverage method. As shown in Figure 7, the reliable area is bound by these two
residual indexes and critical leverage limit. The critical straight lines are shown by red and green
colors. This plot is known as William’s plot. In this plot, normalized residual is depicted versus hat
value which is determined from the main diagonal of aforementioned matrix. It is obvious that the
major number of solubility data are located in this area which expresses validation of the hydrocarbon
solubility databank.

Figure 7. Detection of suspected data for hydrocarbon solubility dataset.

In the most of parametric studies, it is a valuable attempt to identify the effectiveness of all
inputs on the target. According to this fact, the sensitivity analysis is employed to investigate effect
of concentration of components in gaseous mixture; ionic strength of solution; and temperature and
pressure on the solubility of hydrocarbons in aqueous electrolyte systems. To this end, the relevancy
factor should be determined as follows for each input parameter [46–54]:

n 
i=1 (Xk,i − Xk ) Y i − Y
r=  (12)

n 2
n 2
i=1 (Xk,i − Xk ) i=1 (Yi − Y)

78
Processes 2020, 8, 92

In which Yi and Y denote the ‘i’ th output and output average. Xk,i and Xk are known as ‘k’
th of input and average of input. Figure 8 shows the relevancy factor for each effective variable of
hydrocarbon solubility. It is necessary to explain that the relevancy factor lies in range of −1 to 1 so
that the higher absolute value has more impact on hydrocarbon solubility. Furthermore, the positive
relevancy factor shows the straight relationship between input and target. The relevancy factors for
pressure, temperature, the index of fraction, ionic strength, methane, ethane, propane, and butane mole
fraction in gas phase are 0.52, 0.20, −0.48, −0.16, 0.11, 0.06, −0.19, and −0.07 respectively. According
to this explanation and results, as pressure, temperature, and mole fraction of methane and ethane
increase, the solubility of investigated hydrocarbon increases. Moreover, pressure and mole fraction of
ethane in gaseous phase are the most and least effective parameters for determination of solubility
of hydrocarbons.

Figure 8. Sensitivity analysis for solubility of hydrocarbons.

4. Conclusions
The hydrocarbon solubility in aqueous electrolyte phases at high temperature and pressure
conditions is known as a major effective parameter in variety of applications for petroleum industries
and chemical engineering. Numerous attempts have been made in the current study to suggest a highly
accurate and comprehensive predicting tool on the basis of extreme learning machine to calculate
hydrocarbons solubility in wide ranges of operational conditions. Comparing the ELM outputs with
a comprehensive real databank which has 1175 solubility points concluded to R-squared values of
0.985 and 0.987 for training and testing phases respectively. The excellent agreements of ELM and
real hydrocarbon solubility values express that the ELM algorithm is a valuable tool for design and
optimization of various processes that are related to vapor-liquid equilibrium. Furthermore, this study
gives more information about the intensity of each input parameter on solubility of hydrocarbons.
Due to the aforementioned results, this work has potential application in commercial software packages
such as CMG and ECLIPSE for simulation of fluid flow in porous media.

Supplementary Materials: The following are available online at http://www.mdpi.com/2227-9717/8/1/92/s1,


Table S1: data set.
Author Contributions: Conceptualization, A.B. and S.S.; Methodology, N.N., A.M., and A.B.; Software, N.N.,
A.M., A.B., S.S., and I.F.; Validation, N.N., A.M., A.B., S.S., and I.F.; Formal analysis, N.N., A.M., A.B., S.S., and I.F.;
Investigation, N.N., A.M., A.B., S.S., and I.F.; Resources, N.N., A.M., A.B., S.S., and I.F.; Data curation, N.N., A.M.,
A.B., S.S., and I.F.; Writing—Original draft preparation, N.N., A.M., and A.B.; Writing—Review and editing, N.N.,

79
Processes 2020, 8, 92

A.M., A.B., S.S., and I.F.; Visualization, N.N., A.M., A.B., S.S., and I.F.; Project administration and submission,
A.M., Supervision, I.F. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: We acknowledge the financial support of this work by the Hungarian State and the European
Union under the EFOP-3.6.1-16-2016-00010 project.
Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Mohammadi, A.H.; Chapoy, A.; Tohidi, B.; Richon, D. Gas solubility: A key to estimating the water content
of natural gases. Ind. Eng. Chem. Res. 2006, 45, 4825–4829. [CrossRef]
2. Chapoy, A.; Haghighi, H.; Tohidi, B. Development of a Henry’s constant correlation and solubility
measurements of n-pentane, i-pentane, cyclopentane, n-hexane, and toluene in water. J. Chem. Thermodyn.
2008, 40, 1030–1037. [CrossRef]
3. Chapoy, A.; Mokraoui, S.; Valtz, A.; Richon, D.; Mohammadi, A.H.; Tohidi, B. Solubility measurement and
modeling for the system propane–water from 277.62 to 368.16 K. Fluid Phase Equilibria 2004, 226, 213–220.
[CrossRef]
4. Chapoy, A.; Mohammadi, A.H.; Richon, D.; Tohidi, B. Gas solubility measurement and modeling for
methane–water and methane–ethane–n-butane–water systems at low temperature conditions. Fluid Phase
Equilibria 2004, 220, 111–119. [CrossRef]
5. Dhima, A.; de Hemptinne, J.-C.; Moracchini, G. Solubility of light hydrocarbons and their mixtures in pure
water under high pressure. Fluid Phase Equilibria 1998, 145, 129–150. [CrossRef]
6. Kiepe, J.; Horstmann, S.; Fischer, K.; Gmehling, J. Experimental determination and prediction of gas solubility
data for methane+ water solutions containing different monovalent electrolytes. Ind. Eng. Chem. Res. 2003,
42, 5392–5398. [CrossRef]
7. Bamberger, A.; Sieder, G.; Maurer, G. High-pressure (vapor+ liquid) equilibrium in binary mixtures of
(carbon dioxide+ water or acetic acid) at temperatures from 313 to 353 K. J. Supercrit. Fluids 2000, 17, 97–110.
[CrossRef]
8. Mohammadi, A.H.; Chapoy, A.; Tohidi, B.; Richon, D. Water content measurement and modeling in the
nitrogen+ water system. J. Chem. Eng. Data 2005, 50, 541–545. [CrossRef]
9. Marinakis, D.; Varotsis, N. Solubility measurements of (methane+ ethane+ propane) mixtures in the aqueous
phase with gas hydrates under vapour unsaturated conditions. J. Chem. Thermodyn. 2013, 65, 100–105. [CrossRef]
10. Kondori, J.; Zendehboudi, S.; Hossain, M.E. A review on simulation of methane production from gas hydrate
reservoirs: Molecular dynamics prospective. J. Pet. Sci. Eng. 2017, 159, 754–772. [CrossRef]
11. Kondori, J.; Zendehboudi, S.; James, L. Evaluation of gas hydrate formation temperature for gas/water/
salt/alcohol systems: Utilization of extended UNIQUAC model and PC-SAFT equation of state. Ind. Eng.
Chem. Res. 2018, 57, 13833–13855. [CrossRef]
12. Tong, D.; Trusler, J.M.; Vega-Maza, D. Solubility of CO2 in aqueous solutions of CaCl2 or MgCl2 and in a
synthetic formation brine at temperatures up to 423 K and pressures up to 40 MPa. J. Chem. Eng. Data 2013,
58, 2116–2124. [CrossRef]
13. Teng, H.; Yamasaki, A. Solubility of liquid CO2 in synthetic sea water at temperatures from 278 K to 293 K
and pressures from 6.44 MPa to 29.49 MPa, and densities of the corresponding aqueous solutions. J. Chem.
Eng. Data 1998, 43, 2–5. [CrossRef]
14. Lucile, F.; Cézac, P.; Contamine, F.; Serin, J.-P.; Houssin, D.; Arpentinier, P. Solubility of carbon dioxide in
water and aqueous solution containing sodium hydroxide at temperatures from (293.15 to 393.15) K and
pressure up to 5 MPa: Experimental measurements. J. Chem. Eng. Data 2012, 57, 784–789. [CrossRef]
15. Nighswander, J.A.; Kalogerakis, N.; Mehrotra, A.K. Solubilities of carbon dioxide in water and 1 wt.%
sodium chloride solution at pressures up to 10 MPa and temperatures from 80 to 200.◦ C. J. Chem. Eng. Data
1989, 34, 355–360. [CrossRef]
16. Dhima, A.; de Hemptinne, J.-C.; Jose, J. Solubility of hydrocarbons and CO2 mixtures in water under high
pressure. Ind. Eng. Chem. Res. 1999, 38, 3144–3161. [CrossRef]
17. Michels, A.; Gerver, J.; Bijl, A. The influence of pressure on the solubility of gases. Physica 1936, 3, 797–808.
[CrossRef]

80
Processes 2020, 8, 92

18. O’Sullivan, T.D.; Smith, N.O. Solubility and partial molar volume of nitrogen and methane in water and in
aqueous sodium chloride from 50 to 125.deg. and 100 to 600 atm. J. Phys. Chem. 1970, 74, 1460–1466.
19. Vul’fson, A.; Borodin, O. A thermodynamic analysis of the solubility of gases in water at high pressures and
supercritical temperatures. Russ. J. Phys. Chem. A 2007, 81, 510–514. [CrossRef]
20. Wang, L.-K.; Chen, G.-J.; Han, G.-H.; Guo, X.-Q.; Guo, T.-M. Experimental study on the solubility of natural
gas components in water with or without hydrate inhibitor. Fluid Phase Equilibria 2003, 207, 143–154.
[CrossRef]
21. Chapoy, A.; Mohammadi, A.H.; Tohidi, B.; Richon, D. Gas solubility measurement and modeling for the
nitrogen+ water system from 274.18 K to 363.02 K. J. Chem. Eng. Data 2004, 49, 1110–1115. [CrossRef]
22. Prutton, C.; Savage, R. The solubility of carbon dioxide in calcium chloride-water solutions at 75, 100, 120 and
high pressures1. J. Am. Chem. Soc. 1945, 67, 1550–1554. [CrossRef]
23. Bando, S.; Takemura, F.; Nishio, M.; Hihara, E.; Akai, M. Solubility of CO2 in aqueous solutions of NaCl at
(30 to 60) C and (10 to 20) MPa. J. Chem. Eng. Data 2003, 48, 576–579. [CrossRef]
24. Smith, N.O.; Kelemen, S.; Nagy, B. Solubility of natural gases in aqueous salt solutions—II: Nitrogen
in aqueous NaCl, CaCl2, Na2SO4 and MgSO4 at room temperatures and at pressures below 1000 psia.
Geochim. Cosmochim. Acta 1962, 26, 921–926. [CrossRef]
25. Crovetto, R.; Fernández-Prini, R.; Japas, M.L. Solubilities of inert gases and methane in H2O and in D2O in
the temperature range of 300 to 600 K. J. Chem. Phys. 1982, 76, 1077–1086. [CrossRef]
26. Battino, R.; Clever, H.L. The solubility of gases in liquids. Chem. Rev. 1966, 66, 395–463. [CrossRef]
27. Kang, X.; Liu, C.; Zeng, S.; Zhao, Z.; Qian, J.; Zhao, Y. Prediction of Henry’s law constant of CO2 in ionic
liquids based on SEP and Sσ-profile molecular descriptors. J. Mol. Liq. 2018, 262, 139–147. [CrossRef]
28. Kang, X.; Qian, J.; Deng, J.; Latif, U.; Zhao, Y. Novel molecular descriptors for prediction of H2S solubility in
ionic liquids. J. Mol. Liq. 2018, 265, 756–764. [CrossRef]
29. Kang, X.; Zhao, Y.; Li, J. Predicting refractive index of ionic liquids based on the extreme learning machine
(ELM) intelligence algorithm. J. Mol. Liq. 2018, 250, 44–49. [CrossRef]
30. Qiao, W.; Huang, K.; Azimi, M.; Han, S. A Novel Hybrid Prediction Model for Hourly Gas Consumption in
Supply Side Based on Improved Machine Learning Algorithms. IEEE Access 2019, 7, 88218–88230. [CrossRef]
31. Qiao, W.; Lu, H.; Zhou, G.; Azimi, M.; Yang, Q.; Tian, W. A hybrid algorithm for carbon dioxide emissions
forecasting based on improved lion swarm optimizer. J. Clean. Prod. 2020, 244, 118612. [CrossRef]
32. Hemmati-Sarapardeh, A.; Hajirezaie, S.; Soltanian, M.R.; Mosavi, A.; Nabipour, N.; Shamshirband, S.;
Chau, K.W. Modeling natural gas compressibility factor using a hybrid group method of data handling.
Eng. Appl. Comput. Fluid Mech. 2020, 14, 27–37. [CrossRef]
33. Qiao, W.; Yang, Z.; Kang, Z.; Pan, Z. Short-term natural gas consumption prediction based on Volterra adaptive
filter and improved whale optimization algorithm. Eng. Appl. Artif. Intell. 2020, 87, 103323. [CrossRef]
34. Choubin, B.; Abdolshahnejad, M.; Moradi, E.; Querol, X.; Mosavi, A.; Shamshirband, S.; Ghamisi, P. Spatial
hazard assessment of the PM10 using machine learning models in Barcelona, Spain. Sci. Total Environ. 2020,
701, 134474. [CrossRef]
35. Zhao, Y.; Gao, J.; Huang, Y.; Afzal, R.M.; Zhang, X.; Zhang, S. Predicting H 2 S solubility in ionic liquids by
the quantitative structure–property relationship method using S σ-profile molecular descriptors. RSC Adv.
2016, 6, 70405–70413. [CrossRef]
36. Huang, G.-B.; Zhu, Q.-Y.; Siew, C.-K. Extreme learning machine: A new learning scheme of feedforward
neural networks. Neural Netw. 2004, 2, 985–990.
37. Huang, G.-B.; Zhu, Q.-Y.; Siew, C.-K. Extreme learning machine: Theory and applications. Neurocomputing
2006, 70, 489–501. [CrossRef]
38. Bengio, Y. Learning deep architectures for AI. Found. Trends®Mach. Learn. 2009, 2, 1–127. [CrossRef]
39. Liu, Q.; Yin, J.; Leung, V.C.; Zhai, J.-H.; Cai, Z.; Lin, J. Applying a new localized generalization error model
to design neural networks trained with extreme learning machine. Neural Comput. Appl. 2016, 27, 59–66.
[CrossRef]
40. Rao, C.R.; Mitra, S.K. Further contributions to the theory of generalized inverse of matrices and its applications.
Sankhyā Indian J. Stat. Ser. A 1971, 33, 289–300.
41. Bemani, A.; Baghban, A.; Mohammadi, A.H. An insight into the modeling of sulfur solubility of sour gases
in supercritical region. J. Pet. Sci. Eng. 2019, 184, 106459. [CrossRef]

81
Processes 2020, 8, 92

42. Razavi, R.; Bemani, A.; Baghban, A.; Mohammadi, A.H.; Habibzadeh, S. An insight into the estimation of
fatty acid methyl ester based biodiesel properties using a LSSVM model. Fuel 2019, 243, 133–141. [CrossRef]
43. Mesbah, M.; Soroush, E.; Azari, V.; Lee, M.; Bahadori, A.; Habibnia, S. Vapor liquid equilibrium prediction
of carbon dioxide and hydrocarbon systems using LSSVM algorithm. J. Supercrit. Fluids 2015, 97, 256–267.
[CrossRef]
44. Rousseeuw, P.J.; Leroy, A.M. Robust Regression and Outlier Detection; John Wiley & Sons: Hoboken, NJ, USA, 2005.
45. Razavi, R.; Sabaghmoghadam, A.; Bemani, A.; Baghban, A.; Chau, K.-W.; Salwana, E. Application of ANFIS
and LSSVM strategies for estimating thermal conductivity enhancement of metal and metal oxide based
nanofluids. Eng. Appl. Comput. Fluid Mech. 2019, 13, 560–578. [CrossRef]
46. Baghban, A.; Kahani, M.; Nazari, M.A.; Ahmadi, M.H.; Yan, W.-M. Sensitivity analysis and application of
machine learning methods to predict the heat transfer performance of CNT/water nanofluid flows through
coils. Int. J. Heat Mass Transf. 2019, 128, 825–835. [CrossRef]
47. Baghban, A.; Kardani, M.N.; Mohammadi, A.H. Improved estimation of Cetane number of fatty acid methyl
esters (FAMEs) based biodiesels using TLBO-NN and PSO-NN models. Fuel 2018, 232, 620–631. [CrossRef]
48. Baghban, A.; Adelizadeh, M. On the determination of cetane number of hydrocarbons and oxygenates using
Adaptive Neuro Fuzzy Inference System optimized with evolutionary algorithms. Fuel 2018, 230, 344–354.
[CrossRef]
49. Zarei, F.; Rahimi, M.R.; Razavi, R.; Baghban, A. Insight into the experimental and modeling study of process
intensification for post-combustion CO2 capture by rotating packed bed. J. Clean. Prod. 2019, 211, 953–961.
[CrossRef]
50. Baghban, A.; Pourfayaz, F.; Ahmadi, M.H.; Kasaeian, A.; Pourkiaei, S.M.; Lorenzini, G. Connectionist
intelligent model estimates of convective heat transfer coefficient of nanofluids in circular cross-sectional
channels. J. Therm. Anal. Calorim. 2018, 132, 1213–1239. [CrossRef]
51. Bemani, A.; Baghban, A.; Shamshirband, S.; Mosavi, A.; Csiba, P.; Varkonyi-Koczy, A.R. Applying ANN,
ANFIS, and LSSVM Models for Estimation of Acid Solvent Solubility in Supercritical CO2. Preprints 2019.
[CrossRef]
52. Shamshirband, S.; Hadipoor, M.; Baghban, A.; Mosavi, A.; Bukor, J.; Várkonyi-Kóczy, A.R. Developing
an ANFIS-PSO Model to Predict Mercury Emissions in Combustion Flue Gases. Mathematics 2019, 7, 965.
[CrossRef]
53. Shabani, S.; Samadianfard, S.; Sattari, M.T.; Mosavi, A.; Shamshirband, S.; Kmet, T.; Várkonyi-Kóczy, A.R.
Modeling Pan Evaporation Using Gaussian Process Regression K-Nearest Neighbors Random Forest and
Support Vector Machines; Comparative Analysis. Atmosphere 2020, 11, 66. [CrossRef]
54. Ouaer, H.; Hosseini, A.H.; Nait Amar, M.; El Amine Ben Seghier, M.; Ghriga, M.A.; Nabipour, N.;
Andersen, P.Ø.; Mosavi, A.; Shamshirband, S. Rigorous Connectionist Models to Predict Carbon Dioxide
Solubility in Various Ionic Liquids. Appl. Sci. 2020, 10, 304. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

82
processes
Article
Acid-Base Flow Battery, Based on Reverse
Electrodialysis with Bi-Polar Membranes:
Stack Experiments
Jiabing Xia, Gerhart Eigenberger, Heinrich Strathmann and Ulrich Nieken *
Institute of Chemical Process Engineering, University of Stuttgart, Boeblinger Strasse 78, D-70199 Stuttgart,
Germany; [email protected] (J.X.); [email protected] (G.E.);
[email protected] (H.S.)
* Correspondence: [email protected]

Received: 19 November 2019; Accepted: 9 January 2020; Published: 11 January 2020

Abstract: Neutralization of acid and base to produce electricity in the process of reverse electrodialysis
with bipolar membranes (REDBP) presents an interesting but until now fairly overlooked flow battery
concept. Previously, we presented single-cell experiments, which explain the principle and discuss
the potential of this process. In this contribution, we discuss experiments with REDBP stacks at lab
scale, consisting of 5 to 20 repeating cell units. They demonstrate that the single-cell results can be
extrapolated to respective stacks, although additional losses have to be considered. As in other flow
battery stacks, losses by shunt currents through the parallel electrolyte feed/exit lines increases with
the number of connected cell units, whereas the relative importance of electrode losses decreases
with increasing cell number. Experimental results are presented with 1 mole L−1 acid (HCl) and base
(NaOH) for open circuit as well as for charge and discharge with up to 18 mA/cm2 current density.
Measures to further increase the efficiency of this novel flow battery concept are discussed.

Keywords: electrical energy storage; acid-base neutralization flow battery; reverse electrodialysis
with bipolar membranes; stack test results

1. Introduction
The efficient storage and recovery of electrical energy is a key issue in the transition of electric energy
generation from fossil fuels towards sustainable sources like solar and wind power. Flow batteries
present an attractive solution, in particular for stationary and decentralized applications, since the
amount of energy stored is separated from energy conversion. Presently, vanadium redox flow batteries
are the most important and best-studied type of flow batteries. In redox flow batteries, the electric
potential is generated by the electrochemical reactions at the electrodes.
As an alternative, the reverse electrodialysis (RED) has been proposed, where an electrical
potential is generated across ion exchange membranes, which separate salt solutions of different ionic
concentrations. In most cases studied, seawater and river water have been used as the two solutions [1].
Such a RED-stack consists of a sequence of cells containing one compartment for seawater and one for
river water, separated by cation and anion exchange membranes. However, the open voltage of such
a cell is only about 0.15 V, which requires a large number of subsequent cells to obtain a reasonable
stack voltage.
An interesting alternative RED-concept, which will be considered in the following, is based on the
neutralization reaction of acid and base at the internal interface of a bipolar membrane, the reverse
electrodialysis with bipolar membranes (REDBP).

Processes 2020, 8, 99; doi:10.3390/pr8010099 83 www.mdpi.com/journal/processes


Processes 2020, 8, 99

In [2] we presented an overview of previous publications on REDBP [3–5], together with our
experimental result for a single REDBP cell unit, showing the potential advantages and discussing the
present shortcomings of this—so far fairly overlooked—concept.
In the present contribution we extend our experimental results to REDBP stacks, consisting of
5 to 20 cell units. Figure 1 from [2] is a sketch of the stack under charge (lower half) and discharge
conditions (upper half). Only one of several repeating cell units between the electrode chambers is
shown and only the required main ionic fluxes and the electrode reactions are depicted.

Figure 1. Schematic of main ionic fluxes and electrode reactions of the REDBP (reverse electrodialysis
with bipolar membranes) stack used, in the lower part during charge and in the upper part during
discharge. Only one of several repeating cell units is shown between the electrode compartments. The
respective electrode reactions during charge and discharge are indicated at the electrodes. CM: cation
exchange membrane, AM: anion exchange membrane, BM: bipolar membrane.

Each repeating cell unit consists of an acid (HCl) and a base chamber (NaOH) at both sides of
the bipolar membrane (BM). A salt compartment (NaCl), separated from the acid or base chamber by
an anion exchange membrane (AM) or a cation exchange membrane (CM), provides or accepts the
respective Cl− and Na+ ions to ensure electroneutrality. To form a REDBP stack, several repeating cell
units have to be assembled, with an electrode compartment at each end of the stack.
The main element of each repeating cell unit is the bipolar membrane. Under charge (lower half
of Figure 1), a sufficiently high electrical potential has to be applied over the electrodes such that
water is split inside the bipolar membrane into H+ and OH− . H+ and OH− combine with Cl− and
Na+ from the adjacent salt chambers to produce HCl and NaOH, which can be stored in separate
storage tanks. This corresponds to the well-established process of acid and base generation from salt
solutions by electrodialysis with bipolar membranes [6]. The respective electrode reactions are water
splitting under generation of OH− and of H2 at the negative electrode and with H+ and O2 generation
at the positive electrode. The applied DC power supply transports electrons e− from the positive to the
negative electrode.

84
Processes 2020, 8, 99

If the DC power supply is removed (“open circuit conditions”) a potential will establish across
each of the bipolar membranes in the stack. It prevents the further recombination of H+ and OH−
inside each bipolar membrane. The sum of all resulting membrane potentials can be measured between
the electrodes of the stack as open circuit voltage (OCV).
If the electrodes are connected via an external load (upper part of Figure 1) the unit can be
used as a battery. Now H+ from HCl and OH− from NaOH recombine to water inside each bipolar
membrane and generate electric energy; the ionic fluxes in each repeating cell unit reverse. The resulting
electric potential across all repeating cell units drives the depicted electrode reactions. The electrode
reactions transform the ionic into electronic current, which flows through the load of the external
circuit. The reaction details inside the bipolar membrane during charge and discharge have been
discussed in [2]. They depend on the permselectivities and on the fixed charge densities of the ion
exchange membranes used.
Commercially available membranes currently limit acid and base concentrations to 2 M, but
long-term stability allows only for operation up to 1 M [2]. Using 1 M acid and base, the theoretical
power density is 11 kWh/m3 . Energy densities of vanadium redox flow batteries are in the order of
25 kWh/m3 [7].

2. Experimental
In our stack experiments hydrochloric acid (HCl) and sodium hydroxide solutions (NaOH)of
different concentrations have been used as the respective acid and base, with 0.5 mole/L NaCl as
the resulting salt solution. As shown in Figure 2, the respective solutions were circulated through
the corresponding compartments from external storage tanks. Through the electrode compartments,
a 0.25 mole/L sodium sulphate solution (Na2 SO4 was circulated instead of NaCl, to avoid chlorine
formation in the electrode reactions. In the set-up of Figure 1, tested in this contribution, an additional
salt compartment (NaCl) was placed between the sequence of repeating cell units and each electrode
compartment, separated from the electrode compartments by a cation exchange membrane (CM).
This should prevent the penetration of Cl− into the electrode compartments, which would lead to a
release of Cl2 in the electrode reactions. Prime advantages of the chosen electrolytes are their low price,
ample availability and low ecological concern.

Figure 2. Parallel flow concept for acid, base and salt solution through four repeating cell units in the
middle of a stack. The black arrows indicate the flow directions of acid (HCl), of base (NaOH) and of
salt solution (NaCl), and the colored arrows indicate the main fluxes of specific ions during discharge,
also at OCV.

Table 1 shows the expected electrode reactions during charge and discharge. Their standard
potentials represent a voltage sink of 1.23 + 0.83 = 2.06 V, both at charge and at discharge. This loss

85
Processes 2020, 8, 99

could be reduced by modified electrode reactions with gas-consuming electrodes, but this was not in
the scope of the present study.

Table 1. Electrode reactions during charge and discharge.

Electrochemical Reaction Standard Potential


+ − −1.229 V
2 H2 O → H + 4 O2 ↑ +e
Left electrode (negative): Anode 1 1
Discharge
Right electrode (positive): Cathode e− + H2 O → OH− + 12 H2 ↑ +0.8277 V
Left electrode (negative): Cathode e− + H2 O → OH− + 12 H2 ↑ −0.8277 V
Charge + −
Right electrode (positive): Anode 2 H2 O → H + 4 O2 ↑ +e
1 1 +1.229 V

Since the open circuit voltage (OCV) obtained in the single cell experiments with 1 M acid and
base was 0.76 V, only a stack with three and more cell units would be able to compensate the electrode
losses and deliver a net voltage during discharge. The influence of the electrode losses obviously
decreases with an increasing number of cell units in the stack. However, the stack voltage is not fully
proportional to the number of cell units. This is partly due to the so-called shunt currents, which flow
through the feed and exit lines of the different electrolytes, transforming electrical energy into heat.

2.1. Parallel Flow Concept


In our set-up, a parallel flow concept for the electrolytes has been used, as shown schematically
in Figure 2 for four repeating cell units in the middle of a stack. The parallel flow concept has the
advantage of reduced pressure drop, compared to a serial flow concept where all of the acid, base and
salt solution flows consecutively through adjacent cell units. The potential disadvantage is that the
flow through adjacent parallel chambers may differ, if the flow resistances of the chambers are not
equal. In addition, a parallel flow concept leads to larger ionic shunt currents through the feed/exit
lines, as discussed in the next subsection. The colored arrows in Figure 2 indicate the main ionic fluxes
across the membranes at discharge. At a smaller extent, similar fluxes also occur at OCV conditions.
They are caused by a breakthrough of co-ions (Cl− through CM and H+ through AM), as discussed
more detailed in [8]. This breakthrough is responsible for the fact that the open current voltage (OCV)
across a repeating cell unit for 1 mol/L HCl and NaOH at 25 ◦ C decreases from the theoretical value of
0.828 V to a measured value of 0.764 V.
Like conventional electrodialysis stacks, the stack is formed from cell frames with pathways for
the respective electrolytes, grid spacers in the cell frame windows, with the respective membranes and
with rubber gaskets between cell frames and membranes. The cell frames for the stack experiments of
0.5 mm thickness have been obtained from DEUKUM Co, Frickenhausen, Germany (www.deukum.de),
see Figure 3. They proved to be well suited for our lab-scale experiments, since they allow feeding of
four different electrolytes through the holes in the frame. Their active membrane area is about 100 cm2 .
As for the single cell tests, the membranes (fumasep® FKB, fumasep® FAB and fumasep® FBM,) have
been provided by Fumatech Co., Bietigheim-Bissingen, Germany. Their properties are listed in Table 2.

Table 2. Properties of membranes fumasep® FAB, fumasep® FBM with fumasep® PEEK as
reinforcement, provided by Fumatech [9].

Thickness IEC Selectivity Specific Area


Type Reinforcement
[μm] [meqg−1 ] [%] Resistance [Ω/cm2 ]
FAB anion PEEK 100–130 1.0–1.1. 94–97 4–7
FKB cation PEEK 100–130 1.2–1.3 98–99 4–6
FBM bipolar PEEK 100–130 - -

86
Processes 2020, 8, 99

Figure 3. Subsequent cell frames for one repeating cell unit: (a) for up-flow of HCl, (b) for up-flow of
NaOH and (c) for cross-flow of the NaCl solution.

Figure 3 shows the flow through subsequent cell frames, (a) for up-flow of HCl, (b) for up-flow of
NaOH and (c) for cross-flow of the NaCl solution. Together with the respective membranes, separated
by rubber gaskets and grid spacers, the three frames form the repeating cell units shown schematically
in Figure 1. The feed and exit flows of the respective electrolytes pass through the holes in the cell
frames as indicated by the respective colors in Figure 3. All electrolytes are circulated by separate
pumps from reservoirs through the respective compartments as indicated in Figure 2. Further details
of the design are given in [8].

2.2. Voltage Measurements


The most direct information during stack operation can be obtained from voltage measurements
along the stack. In the single repeating cell unit studied in [2], Haber-Luggin capillaries with calomel
electrodes have been used to measure the voltage drop across the respective membranes. To reduce the
ample space required for the Haber-Luggin capillaries, in the present stack experiments the voltages
between different repeating cell units have been measured by thin Pt wires. The insulated Pt wires
were introduced and sealed between two rubber gaskets. To minimize concentration effects, the Pt
wires were only positioned in the acid compartments, where ionic conductivities are high and about
equal throughout the stack.
In the following, voltage measurements will be presented where Pt wire probes were positioned
in the center of the 1st, the 2nd, the 6th, the 11th and the 19th acid cell of the stack. This allows
measurement of the voltages across the 1st repeating cell unit, across the first five-cell unit (next to the
electrode) and across the second five-cell unit (in the middle of the stack). In order to determine the
total voltage of the 20-cell stack, the measured first cell voltage was added to the voltage measured
between the first and the 19th acid compartment, assuming that the voltage across the repeating cell
units next to both electrodes was equal.

2.3. Operating Conditions


In all cases, Na2 SO4 electrolyte of 0.25 M was circulated through the electrode chambers and 0.5 M
NaCl electrolyte was circulated through the salt chambers, while the concentration of HCl and NaOH,
circulated through their chambers, was varied between 0.5 M and 1 M. If not specified differently,
the empty-space flow velocity through the electrolyte chambers was adjusted to a mean value of about
2 to 3 cm/s.

3. Results and Discussion


In this section, the measured voltages in stacks consisting of 5 to 20 repeating cell units are
presented and discussed for different acid and base concentrations, for different charge and discharge
currents and for different charge/discharge periods.

87
Processes 2020, 8, 99

3.1. Open Circuit Voltage (OCV)


With the Pt wire probes, the open circuit voltage can be measured between different cell units as
mentioned above, but OCV for the whole stack can also be measured directly between the electrodes of
the stack. Table 3 presents OCV results for increasing stack size between 1 and 20 repeating cell units if
the electrolytes flow through the stack continuously (constant ion concentrations). In an ideal stack,
OCV should amount to the theoretical single cell voltage (0.828 V for 1 M acid and base at 25 ◦ C) times
the number of repeating cell units. With the measured single cell voltage of 0.764 V, the theoretical
values of the stack are given in the second row. The actually obtained OCV values, measured across
the electrodes, are given in the third row. While the measured values are still close to the calculated
values for up to 10 repeating cell units, an increasing voltage loss is observed if the cell unit number is
raised to 15 and further to 20 cells. Reasons for this trend are discussed in the next section.

Table 3. Calculated and measured open circuit voltage (OCV) for 1 M acid and base at 25 ◦ C, depending
on the number of repeating cell units.

Number of Repeating Cell Units 1 5 10 15 20


OCV calculated from single cell voltage 0.764 V 3.82 V 7.64 V 11.46 V 15.28 V
OCV measured across electrodes 0.764 V 3.8 V 7.6 V 11.2 V 14.4 V

3.2. Self-Discharge at OCV


The experimental results of Table 2 have been obtained with stacks where the required electrolytes
were continuously pumped through the stack, compensating the effect of any side reactions or shortcuts,
which may consume acid and base at OCV. In the OCV experiments reported in Figure 4, the electrolyte
flows have been stopped at the beginning of the experiment. Now the resulting drop of the voltage in
stacks with different numbers of repeating cell units can be analyzed over time. This should provide
indications as to why OCV across stacks with a larger number of repeating cell units increases less than
proportional to the number of units, as shown in Table 2. A self-discharge due to limited membrane
permselectivities, as indicated by the ionic flux arrows at OCV in Figure 2 and discussed in detail in [2],
could contribute to the effect. However, it should not be responsible for the less than proportional
increase of OCV with the number of repeating cell units.
Figure 4a–d show the results of the self-discharging tests of (a) 5-cell, (b) 10-cell, (c) 15-cell and
(d) 20-cell stacks with initial 1 M acid and base. In a 5-cell stack (Figure 4a), the self-discharge is
very low, and OCV remains close to 4V during the whole measurement period. In a 10-cell stack
(Figure 4b), the stack voltage is about doubled and a slow self-discharge is followed by an accelerated
discharge after about 30 min down to a level of 4 V. If the cell number is further increased (Figure 4c,d),
the accelerated drop from the higher stack voltage down to about 4 V occurs at about 20 min for the
15-cell stack, and at about 12 min for the 20-cell stack. This faster decrease of OCV is consistent with the
assumption that ionic shunt current through the feed/exit lines of the electrolytes should be responsible
for the OCV decrease. A similar influence of shunt currents through parallel feed/exit lines of the
electrolytes is well known from redox flow batteries [10–12].

88
Processes 2020, 8, 99

Figure 4. Self-discharge test at 25 ◦ C for different stack sizes and different acid and base concentrations.
(a) 5-cell stack, 1M acid and base, (b) 10-cell stack, 1M acid and base, (c) 15-cell stack, 1M acid and base,
(d) 20-cell stack, 1M acid and base, (e) 20-cell stack with 0.5 M acid/base concentration. OCV_Stack
(black) has been measured between the electrodes, OCV_AllSCs (blue) between the first and the last
repeating cell unit of the stack, OCV_5cells_1st (green solid) and OCV_5cells_2nd (green, dotted) have
been measured across the first and second 5-cell unit.

In stacks with 10 and more cells, OCV shows an accelerated decline to about 4 V after a certain
period of low decline, which becomes shorter with increasing cell number. An explanation can be drawn
from Figure 4d. Here, the difference between OCV over the first 5 cells of the stack (OCV_5cells_1st,
green solid line) and OCV over the second 5-cell package (OCV_5cells_2nd, green dotted line) is shown.
The voltage over the 5 middle cells was smaller than that over the first 5 cells from the very beginning.
It dropped to (almost) zero at the time of the accelerated decline of the total cell voltage, whereas the
voltage over the first 5 cells was less affected. The voltage decline indicates that the acid and base have
been completely consumed (neutralized) in the middle of the stack, but are still present in the cells
next to the electrodes. This can be explained by the ionic shunt currents through the feed/exit lines
of the electrolytes. Since H+ and OH− have much higher ionic mobility than the other ions present,
only the influence of H+ and OH− shunt currents in the acid and base feed/exit lines will be considered
in the following, using the simplified picture in Figure 5.

89
Processes 2020, 8, 99

Figure 5. (a) Simplified sketch of ionic shunt currents of H+ (red arrows) and OH− (blue arrows)
through the feed/exit lines of acid and base, and compensating ionic currents inside of four repeating
cell units in the middle of the stack. The different arrow thickness indicates the increased shunt currents
in the stack center. (b) Corresponding electric potential profiles ϕ across the repeating cell units (black)
and in the acid (red) and the base feed/exit lines (blue). The arrows indicate the direction of the ion
fluxes between cell compartments and feed/exit lines.

In Figure 5a only four repeating cell units in the middle of the stack are shown, together with their
feed/exit lines for acid (red) and base (blue). At OCV, a voltage profile ϕ (black) develops across the
stack as shown in Figure 5b. It mainly consists of the voltage steps in each of the bipolar membranes.
The voltage gradient across the stack causes a counter-movement of H+ ions and of OH− ions along
their feed/exit lines as indicated by the red and blue arrows in Figure 5a. The ions originate from the
acid/base compartments at one side of the stack and turn back into the respective compartments at the
opposite side. This leads to an ion flux maximum in the stack center. To ensure electroneutrality, the
ion fluxes through the feed/exit lines have to be compensated by fluxes of opposite direction through
the cell compartments. This leads to the indicated fluxes with neutralization of H+ and OH− inside the
bipolar membranes, which is maximal in the middle of the stack. Comparable to a simple mixing of
acid and base, this neutralization energy is converted into heat, representing a self-discharge with loss
of electrical energy. The fluxes of H+ and OH− into the bipolar membranes have to be compensated by
equivalent fluxes of Cl− and Na+ into the adjacent salt compartments.
Arrows in Figure 5b qualitatively depict the driving potentials between the feed/exit lines and
the potentials on both sides of the respective bipolar membranes. This explains why the ion fluxes
of H+ and OH- between the stack cells and the feed/exit lines change direction in the middle of the
stack, as indicated by the arrows in Figure 5a,b. The fluxes of H+ and OH− through the feed/exit lines
accumulate in the stack center as indicated by the arrow thickness. The compensating fluxes inside the
repeating cell units are therefore also strongest in the middle of the stack. This leads to the pronounced
self-discharge and the resulting drop of OCV across the bipolar membranes in the middle of the stack.
In summary, the ionic shunt currents of H+ and OH− ions in the acid/base feed exit lines are driven
by the potential along the acid/base feed/exit lines and result in a neutralization reaction in the bipolar
membranes. Estimation shows that in the stack used, shunt currents due to H+ migration through the

90
Processes 2020, 8, 99

feed/exit lines of HCl amount to more than 60% and due to OH− migration through the feed/exit lines,
NaOH amounts to more than 30% of the total shunt current [8]. This leads to the consumption of acid
and base, primarily in the stack center, which reduces the respective cell voltages. If under OCV the
flow of acid and base is stopped, it leads to the gradual neutralization of all acid and base in the stack
center as can be concluded from the voltage drop to zero of “OCV_5cells_2nd” after about 12 min in
Figure 4d.
In Figure 4e, results are presented where the initial acid/base concentration has been reduced
from 1 to 0.5 M. This reduction of 50% should result in an accelerated discharge in about half the time.
However, since the reduced concentrations also lead to reduced ionic conductivities and hence to a
reduced shunt current, only a small difference between Figure 4d,e can be observed. Again, OCV
across the second 5-cell unit in the stack center drops to zero, now after around 10 min.
The above details of the observed OCV decline after stopping the acid/base feed help to elucidate
the influence of shunt currents. However, it should be mentioned that shunt currents are only of
importance for the REDBP battery during operation. The consumption of acid and base by self-discharge
at OCV is limited to the amount present in the stack and does not affect the concentration of acid and
base in the storage tanks. This is a general advantage of flow batteries.
To determine all the details of ionic flux interactions, a numerical simulation of the whole stack
would be required, where ionic species balances for all ionic species as well as electroneutrality have to
be enforced in the whole stack. But this is out of scope for the present contribution.

3.3. Charge and Discharge Behavior


Figure 6 shows the charge and discharge behavior of a 20-cell stack with 1 M acid/base concentration
at 25 ◦ C. Starting with OCV conditions, after 1 min a charge current density of 9 mA/cm2 is imposed
for 5 min, followed by zero current for 1 min and a discharge current density of 9 mA/cm2 for another
5 min. The voltage measured across the stack (black line) is higher under charging and lower under
discharging than the voltage measured with the Pt wire probes between all repeating cell units, because
the electrode losses need to be added to the blue curve under charge and subtracted under discharge.
The mean voltages measured along the 5 min charge and discharge periods are constant. Under
constant acid and base concentrations, the measured voltages even remain constant during extended
charge and discharge periods, indicating a stable operation. Again, the voltage measured across the
second 5-cell package (V5cells_2nd, green dotted line) is lower than V5cells_1st (green solid line) at
OCV as well as under charge/discharge, indicating the increased influence of the shunt currents in the
stack center.
The absolute difference between the blue and the black curves should, however, be about equal for
charge and discharge and vanish at OCV. This would be the case if the blue lines were shifted upwards
by 0.7 V, leading to an absolute difference between both curves at charge and discharge of about
3.4 V. This seems to be a reasonable estimate of the losses at the electrodes and across the electrode
compartments during charge/discharge with 9 mA/cm2 , since under ideal conditions electrode losses
of 2.07 V have been predicted by Table 1. The required voltage shift of 0.7 V points to a measurement
error of the Pt wire probes (blue curves). It is well known that measurements with Pt wire probes are
less accurate than with Haber-Luggin electrodes as applied in [2]. In the summary of the measured
results, presented in Figure 7, this voltage shift has been considered and compensated.

91
Processes 2020, 8, 99

Figure 6. Charge and discharge behavior of a 20-cell stack with 1 M acid/base at 25 ◦ C and 9 mA/cm2
current density. The black line is the stack voltage measured across the electrodes; the blue line is the
stack voltage across the 20-cell units. The voltages measured across the first repeating cell unit next
to the electrode (red), the first 5-cell unit (green, solid) and the second 5-cell unit (green, dotted) are
also shown.

Figure 7. Discharge voltages and net discharge power densities of the 20-cell stack with 1 M acid/base
at 25 ◦ C, plotted over discharge current density.

In Figure 8, measured voltages are presented for increasing current densities under charge
(increasing voltages) and under discharge (decreasing voltages). At each current density,
the measurement continued until a stable voltage was established. Figure 8a shows the results
of a 20-cell stack and Figure 8b of a 15-cell stack, for 1M acid and base. Both results show the same
general trends. The voltages measured across the electrodes of the whole stack are given by black
dots and the measurements across all repeating cell units (excluding the electrodes) by blue dots.
In addition, measurements across the first 5-cell unit (next to one electrode) are given by green solid

92
Processes 2020, 8, 99

lines and the measurements across the second 5-cell unit (in the middle of the stack) by green broken
lines. Measurements for one repeating cell unit next to one electrode are displayed in red. If a shift of
+0.7 V is imposed on the blue curves, as discussed before, the electrode losses between charge and
discharge are about equal. They increase moderately with increasing current density.

Figure 8. Current density measurements during charge and discharge with 1 M acid and base at
25 ◦ C and 0.1 MPa at steady state. Current-voltage curves of positive slope represent charging, of
negative slope discharging. Results for a 20-cell stack are shown in (a) and for a 15-cell stack in (b).
Displayed are the measured voltages, VSC (across BP (bipolar membranes) of the first single cell, red),
V5cell_2nd (across the second 5-cell unit, green dotted), V5cell_1st (across the first 5-cell unit, green
solid), VAllSCs (across all cell units measured between Pt electrodes, blue) and VStack (stack voltage
between electrodes, black).

93
Processes 2020, 8, 99

A comparison of Figure 8a,b shows that the voltage losses increase with the number of unit cells.
This is a consequence of the fact that the shunt current accumulates in the stack center, as discussed
qualitatively with Figure 5. This shunt current has to be compensated by an increased neutralization
reaction of H+ and OH− in the bipolar membrane. The unit cells in the stack center therefore contribute
less to the stack voltage if the cell number increases. A decrease of the voltage of the second 5-cell
unit, located in the middle of the stack, compared to the first 5-cell voltage is also obvious in Figure 8a,
demonstrating the influence of the increasing shunt current in the middle of the stack, both at charge
and discharge.

3.4. Discharge Power Density and Efficiency


The focus of the present contribution is on the discharge behavior of flow batteries. Here, as in [2],
power density PD, defined as net electric power, divided by stack volume, is useful for characterizing
the discharge. For the stack volume, only the active area of the membranes and the thickness of each
repeating cell unit (1.96 mm) have been considered, while the volume of both electrode chambers
has not been included. In Figure 7, discharge voltage and net discharge power density of the tested
20-cell stack with 1 M acid/base at 25 ◦ C are displayed over discharge current density. As explained in
Section 3.3, the voltages measured across all single cells (the blue dotted line) have been moved up by
0.7 V (compared to Figure 8), to compensate for measurement errors of the Pt probes.
Considering an ideal behavior without losses across the stack (zero resistances) and with the
theoretical single cell voltage of 0.828 V, a constant voltage of 16.56 V would result, leading to a linear
increase of power density over current density, marked by the solid red line. The actually measured
voltages across all 20 cells of the stack over current density lead to the power densities as displayed by
the red line with measurement bullets. The losses partly result from the Ohmic resistances across the
membranes and the electrolyte compartments, as already considered in the single-cell experiments
in [2]. Another part results from the shunt currents, discussed in Section 3.2.
If we include the measured voltage drop over the electrodes, the pink line with bullets results in
the measured power density of the stack. The difference between the red and the pink line is attributed
to additional losses in the electrode compartments. The fact that the total losses are substantially
higher than predicted from the single-cell experiments can be attributed to the shunt currents through
the feed/exit lines. In addition, a nonuniform flow distribution of the electrolytes through adjacent
cells could have contributed. This can be concluded from differences between measured voltages of
adjacent cell units.
The highest power density for the 20-cell stack of 15 mW/cm3 has been reached at 10 mA/cm2 ,
corresponding to a stack power output of 6 W with 36% voltage efficiency. Neglecting the electrode
chamber losses, the highest stack power goes up to 9 W or 24 mW/cm−3 with 50% voltage efficiency.
The power density decreases at higher current densities due to increasing resistive losses.
As observed in the single-cell experiments [2], discharge current densities well above 20 mA/cm2
resulted in unstable behavior, with voltage breakdown due to water accumulation inside BP. However,
this is above the discharge power densities reached in the stack experiments.

4. Conclusions
The experimental results obtained in this study show that the single cell results presented in [2]
can be extrapolated to multicell stacks, but additional losses have to be considered. Most importantly,
the influence of shunt currents has to be taken into account. As demonstrated with open current
experiments under self-discharging conditions in Section 3.2, the influence of shunt currents through the
parallel acid/base feed/exit lines increases with increasing number of cell units. It is most pronounced
in the middle of the stack, where it can substantially reduce the open cell voltage.
The single cell experiments reported in [2] showed that at higher acid/base concentrations,
a breakdown of the discharge voltage occurred, if the current density was raised above a certain value
between 20 and 40 mA/cm2 . This was attributed to a delamination of the bipolar membrane, caused by

94
Processes 2020, 8, 99

the excessive formation of water through the neutralization reaction of H+ and OH− in the reaction
layer. However, these current densities were well above the discharge current densities reached in the
stack experiments of Figures 6 and 8. Delamination of bipolar membranes during discharge was not
observed in the stack results presented.
Figure 7 summarizes the experimental discharge results for the 20-cell stacks and points to
possibilities for improvement. A reduction of the influence of shunt currents would directly reduce
the difference between the (ideal) red solid line and the measured red dotted line. A straightforward
approach would be to change the acid and base flow pattern from parallel to a serial flow, where all of
the acid and base flows consecutively upwards in one and downwards in the next acid/base chamber.
Then, minor shunt currents could only occur between adjacent cells, driven by the single cell voltages
but not by the total stack voltage, as in case of the parallel feed/exit lines. Preliminary experiments
at open current under self-discharge (similar to Figure 4) proved that the self-discharge of a 5-cell
stack with serial flow of acid and base was considerably slower, compared with a similar stack with
parallel flow. However, for the present design the pressure drop across the acid/base feed/exit lines
was substantially increased. This would result in an unacceptable pressure drop for a 20-cell stack
with serial flow.
Since Figure 4 has shown that a 5-cell stack with parallel flow was only marginally affected by
shunt currents, a combination of parallel and serial flow could be a compromise between pressure
drop and shunt current. The parallel flow through the acid and base compartments should change
direction after, for example, every 5 to 10 cells. Electrode losses (the difference between the red and the
pink lines in Figure 7) could be reduced, e.g., by improved, gas-consuming electrodes [13].
With respect to higher power densities, the positive influence of larger concentrations of the
electrolytes has to be balanced against reduced permselectivities of the ion exchange membranes used.
Here, the challenge is to develop monopolar and bipolar membranes with increased fixed-charge
concentrations as already discussed in [2]. In the single-cell experiments [2], water accumulation at the
BP interface resulted in a breakdown of the cell voltage at current densities above about 30 mA/cm2 for
1M acid/base. This was well above the current densities reached in the stack experiments. Nevertheless,
an improved water transport through BP will be a further challenge for BP membranes optimized
for REDBP.

Author Contributions: Conceptualization, G.E., H.S. and U.N.; formal analysis, J.X.; investigation, J.X.;
methodology, J.X. and H.S.; resources, J.X. and H.S.; supervision, G.E., H.S. and U.N. All authors have read and
agreed to the published version of the manuscript.
Funding: Jiabing Xia gratefully acknowledges the Ph.D. scholarship from the GREES program at University of
Stuttgart in Germany.
Conflicts of Interest: The authors declare no conflict of interest.

List of Abbreviations
AM anion exchange membrane
BM bipolar membrane
CM cation exchange membrane
OCV open circuit voltage
REDBP reverse electrodialysis with bipolar membranes

References
1. Post, J.W.; Veerman, J.; Hamelers, H.V.M.; Euverink, G.J.W.; Metz, S.J.; Nymeijer, K.; Buisman, C.J.N.
Salinity-gradient power: Evaluation of pressure retarded osmosis and reverse electrodialysis. J. Membr. Sci.
2007, 288, 218–230. [CrossRef]
2. Xia, J.; Eigenberger, G.; Strathmann, H.; Nieken, U. Flow battery based on reverse electrodialysis with bipolar
membranes: Single cell experiments. J. Membr. Sci. 2018, 565, 157–168. [CrossRef]

95
Processes 2020, 8, 99

3. Walther, J.F. Process for Production of Electrical Energy from the Neutralization of Acid and Base in a Bipolar
Membrane Cell. U.S. Patent 06183483, 19 January 1982.
4. Zholkovskij, E.; Müller, M.; Staude, E. The storage battery with bipolar membranes. J. Membr. Sci. 1998, 141,
231–243. [CrossRef]
5. Pretz, J.; Staude, E. Reverse electrodialysis (RED) with bipolar membranes, an energy storage system.
Phys. Chem. 1998, 102, 676–685. [CrossRef]
6. Huang, C.; Xu, T. Electrodialysis with bipolar membranes for sustainable development. Environ. Sci. Technol.
2006, 40, 5233–5243. [CrossRef] [PubMed]
7. Parasuraman, A.; Lim, T.; Menictas, C.; Skyllas-Kazacos, M. Review of material research and development
for vanadium redox flow battery applications. Electrochim. Acta 2013, 101, 27–40. [CrossRef]
8. Xia, J. Reverse Electrodialysis with Bipolar Membranes (REBP) as an Energy Storage System. Ph.D. Thesis,
Fakultät Verfahrenstechnik of Stuttgart University, Stuttgart, Germany, 10 August 2018.
9. Fumatech, Home Page. Available online: www.fumatech.com (accessed on 10 January 2020).
10. Yin, C.; Guo, S.; Fang, H.; Liu, J.; Li, Y.; Tang, H. Numerical and experimental studies of stack shunt current
for vanadium redox flow battery. Appl. Energy 2015, 151, 237–248. [CrossRef]
11. Fink, H.; Remy, M. Shunt currents in vanadium flow batteries: Measurement, modelling and implications for
efficiency. J. Power Sources 2015, 284, 547–553. [CrossRef]
12. Ye, Q.; Hu, J.; Cheng, P.; Ma, Z. Design trade-offs among shunt current, pumping loss and compactness in
the piping system of a multi-stack vanadium flow battery. J. Power Sources 2015, 296, 352–364. [CrossRef]
13. Kintrup, J.; Millaruelo, M.; Trieu, V.; Bulan, A.; Mojica, E.S. Gas Diffusion Electrodes for Efficient
Manufacturing of Chlorine and Other Chemicals. Electrochem. Soc. 2018, 26, 73–76. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

96
processes
Article
Facile Synthesis of Bio-Template Tubular MCo2O4
(M = Cr, Mn, Ni) Microstructure and Its Electrochemical
Performance in Aqueous Electrolyte
Deepa Guragain 1 , Camila Zequine 2 , Ram K Gupta 2 and Sanjay R Mishra 1, *
1 Department of Physics and Materials Science, The University of Memphis, Memphis, TN 38152, USA;
[email protected]
2 Department of Chemistry, Pittsburg State University, Pittsburg, KS 66762, USA;
[email protected] (C.Z.); [email protected] (R.K.G.)
* Correspondence: [email protected]

Received: 17 January 2020; Accepted: 10 March 2020; Published: 16 March 2020

Abstract: In this project, we present a comparative study of the electrochemical performance for tubular
MCo2 O4 (M = Cr, Mn, Ni) microstructures prepared using cotton fiber as a bio-template. Crystal
structure, surface properties, morphology, and electrochemical properties of MCo2 O4 are characterized
using X-ray diffraction (XRD), gas adsorption, scanning electron microscopy (SEM), Fourier transforms
infrared spectroscopy (FTIR), cyclic voltammetry (CV), and galvanostatic charge-discharge cycling
(GCD). The electrochemical performance of the electrode made up of tubular MCo2 O4 structures was
evaluated in aqueous 3M KOH electrolytes. The as-obtained templated MCo2 O4 microstructures
inherit the tubular morphology. The large-surface-area of tubular microstructures leads to a noticeable
pseudocapacitive property with the excellent electrochemical performance of NiCo2 O4 with specific
capacitance value exceeding 407.2 F/g at 2 mV/s scan rate. In addition, a Coulombic efficiency
~100%, and excellent cycling stability with 100% capacitance retention for MCo2 O4 was noted even
after 5000 cycles. These tubular MCo2 O4 microstructure display peak power density is exceeding
7000 W/Kg. The superior performance of the tubular MCo2 O4 microstructure electrode is attributed
to their high surface area, adequate pore volume distribution, and active carbon matrix, which allows
effective redox reaction and diffusion of hydrated ions.

Keywords: bio-template; MCo2 O4 (M = Cr, Mn, Ni); electrochemical; cyclic voltammetry;


specific capacitance

1. Introduction
Supercapacitors (SCs) are the energy storage device. SCs are in high demand because of their
greater power density compared to batteries and higher energy density compared to that of capacitors [1].
In the capacitor, there is no time lag during the charging process; hence it can give higher power
density, and in the battery, there is low self-discharge so that it can provide higher energy density [2].
Because of this, a supercapacitor is easy to charge within a short time and able to show significant
performance even after prolonged use. SCs are of three types: (i) electric double-layer capacitors
(EDLCs), (ii) pseudo capacitors (PCs), and (iii) hybrid capacitors. EDLCs are based on the principle
that physical adsorption takes place on the interface of a solid electrode, usually carbon-based material,
and liquid electrolytes [3,4]. In PCs, surface redox reaction takes place at the electrode-electrolyte
interface, which is responsible for storing electronic charges [5], where metal oxides and conducting
polymer-based materials are used as active electrode materials [6,7]. EDLCs have lower specific
capacitance and energy density as compared to PCs, hence, practically PCs are in higher demand.
A hybrid capacitor is a combination of both EDLCs and PCs; examples are carbon nanotubes, graphene,

Processes 2020, 8, 343; doi:10.3390/pr8030343 97 www.mdpi.com/journal/processes


Processes 2020, 8, 343

etc. [8,9]. They display hybrid charge–storage mechanisms and have the ability to deliver higher
capacity [10].
The electrolyte ion transport in supercapacitor devices occurs through an ion-transport layer
separated from the electrode. The charge storage mechanism follows at the electrode surface during
the charging-discharging process [11].
Transition metal oxides (TMOs) with novel nano-architectures and rich in redox reactions are
increasingly explored for their application in energy storage devices [12]. Among these, cobalt oxides
are highly attractive because of their higher theoretical value for specific capacitance, i.e., 3560 F/g [13].
The nanoarchitecture of these metal-oxides is controlled by the synthesis route, which often requires
complex technological strategies, including toxic organic reagents, which might make it difficult for
their mass industrial application. Thus, it is highly sought to explore cost-effective facile synthesis
strategies and environmentally benign techniques for preparing these electrodes. Furthermore, the ideal
electrode should have a high specific surface area for better specific capacitance, controlled porosity for
better rate capability as well as specific capacitance, and higher electronic conductivity to improve
rate capability and power density of supercapacitor. Nowadays, the bio-templating technique has
emerged as a convincing technique for the preparation of oxide supercapacitors [14–17]. Nature
offers rich and diverse bio-templates [18–20] like bamboo, lotus pollen grains [21], leaf [22], sorghum
straw [23], butterfly wing [24], jute fibers [25], and cotton [12]. Such bio-templates offer elaborate
interior and exterior surfaces, and geometries, which make these templates attractive materials to
produce multiscale hybrid and hierarchical morphologies.
Numbers of research are published on the study of transition metal oxide-based electrodes such as
MnCo2 O4 [5,26–28] and NiCo2 O4 [29–33], Co3 O4 [34], and NiFe2 O4 [35] for their supercapacitive
application. The benefit of transition metal oxides as electrode materials are innumerable, as their
multiple oxidation states facilitate multiple redox reactions during electrochemical reaction vis-a-vise
offers stable structure. The co-existence of two different cations provides abundant active sites to perform
fast reversible faradic redox reaction on the electrode interface; as a result, higher specific capacity,
and excellent rate capability are achieved [25,36]. Additionally, the types of bonds between transition
metal ions and ligands are dictated by electronegativity and ionization energies [37]; with the former,
the structure is dense, while with later the structure is more open. The valence state, ionic radius,
electronegativity [38], and the local environment of the cations are affected by the change in Gibbs free
energy and electrochemical potential of the electrode. An increase in the electrochemical potential of
cathodes is observed with the increase in the number of electrons in d orbitals of transition metal elements.
This implies a higher consumption of energy during electron transfer [39]. In mixed transition metal
oxides, there is a synergetic effect between metal cations; this produces higher electrical conductivity of
single metal oxide where there is low activation energy to transfer electron between metal cations and
gives excellent structural stability [40,41].
Here we present a comparative study to understand the electrochemical behavior of MCo2 O4
(M = Cr, Mn, and Ni) electrodes prepared via a facile bio-template method. The electronegativity
differences among M ions viz. Cr (1.66), Mn (1.55), Ni (1.99), and Co (1.88) could have a substantial effect
on the electrochemical performance of the said electrodes. The electronegativity difference determines
the structure, covalent vs. ionic, and the electric potential of the electrode for the charge transfer,
as discussed above. In the present study, MCo2 O4 electrode material is prepared via the bio-template
method, where the product assumes the morphology of the microstructure of bio-template and ends up
with a carbon matrix. The bio-template method adapted to produce active material inherently fixed in
the carbon matrix. The carbon matrix is known to enhance electrode electrochemical performance [42].
The template supported mineralization MCo2 O4 at room temperature (RT) produces 3D-hierarchical
and porous-MCo2 O4 superstructures with tubular-like morphologies. The doped Co3 O4 (MCo2 O4 ) is
explicitly explored in this study as dopant atoms or vacancies are known to affect the crystal field [43],
thus modifying the electronic structure and adjusting the electrochemical potential [44].

98
Processes 2020, 8, 343

2. Experimental

2.1. Synthesis
All the chemicals required for the synthsis, such as Cobalt nitrate hexahydrate; Co(NO3 )2 .6H2 O,
Chromium nitrate hexahydrate; Cr(NO3 )2 .6H2 O, Manganese nitrate hexahydrate; Mn(NO3 )2 .6H2 O,
and Nickel nitrate hexahydrate; Ni(NO3 )2 .6H2 O were purchased from Sigma-Aldrich, St. Louis,
Missouri, USA. The spinel MCo2 O4 (M = Cr, Mn, Ni) was synthesized by a facile bio-template method.
Quantities of 1.16 g of Co(NO3 )2 .6H2 O and 0.8 gm of Cr(NO3 )2 .6H2 O, 0.55 g of Co(NO3 )2 .6H2 O and
0.17 gm of Mn(NO3 )2 .6H2 O, and 0.86 g of Co(NO3 )2 .6H2 O and 0.43 gm of Ni(NO3 )2 .6H2 O were mixed
in 15 ml of distilled water separately, and the mixture was ultra-sonicated for 10 minutes to make
a homogenous solution. Then, 1.0 g of cotton was soaked in the mixture solutions for 5 minutes.
The resulting soaked cotton was filtered and dried at 150 ◦ C for 30 minutes. The dried cotton was later
calcined at 520 ◦ C for 3 hours in the air to obtain bio-templated CrCo2 O4, MnCo2 O4 , and NiCo2O4
tubular microstructure.

2.2. Characterization
The x-ray diffraction patterns were obtained via Bruker D8 Advance X-ray diffractometer
(Bruker Corporation, Madison, WI, USA) using Cu Kα radiation to check phase purity and determine
the crystalline parameters of as-prepared samples. A scanning electron microscope (Phenom) at
10 keV analyzed the morphology of samples. The Brunauer–Emmett–Teller (BET) method was used
to measure the specific surface area of the samples. The surface area measurement was carried out
by adsorption-desorption isotherms at 77 K, (Quantachrome, Boynton Beach, FL 33426, model No.
AS1MP) using nitrogen as adsorbing gas. Thermogravimetric analyses (TGA, Instrument Specialist,
Inc., Twin Lakes, WI, USA), were performed in 24 to 550 ◦ C temperature range. FTIR spectra were
collected via Theromo-Fisher Scientific FTIR spectrometer (Nicolet iS10, Thermo Fisher Scientific,
Waltham, MA, USA) between 450 and 1000 cm−1 .
Versastat 4–500 electrochemical workstation (Princeton Applied Research, USA) was used to
perform electrochemical measurements in a standard three-electrode configuration. To prepare an
electrode, slurry pastes of 80 wt % of the synthesized powder, 10 wt % of acetylene black, and 10 wt %
of polyvinylidene difluoride (PVDF) were mixed in the presence of N-methyl pyrrolidinone (NMP).
The thoroughly mixed paste was applied onto a nickel foam. Here, the active mass is 80% out of the
total pasted mass in the electrode. The prepared electrodes were dried under vacuum at 60 ◦ C for 10
hours. The loading mass of all samples was about 2–3 mg, measured by weighing the nickel foam
before and after deposition with an analytical balance (MS105DU, Mettler Toledo, 0.01 mg of resolution).
MCo2 O4 (M = Cr, Mn, Ni) coated nickel foam was used as a working electrode, a saturated calomel
electrode (SCE) as a reference electrode, and a platinum wire as a counter electrode. The electrochemical
performance of the electrodes was evaluated at RT in 3M KOH electrolyte via cyclic voltammetry and
galvanostatic charge-discharge techniques measurements.

3. Results and Discussion


Figure 1a shows the XRD patterns of the bio-templated CrCo2 O4 , MnCo2 O4 , and NiCo2 O4
microstructure. The XRD patterns match with the face-centered cubic phase of CrCo2 O4 , MnCo2 O4 ,
and NiCo2 O4 (International Centre for Diffraction Data (ICDD) #02-0770). The main peaks at 30.9◦ ,
36.4◦ , 44.3◦ , 58.6◦ , and 64.3◦ for CrCo2 O4 , 31.1◦ , 36.7◦ , 44.7◦ , 59.2◦ , and 65.9◦ for MnCo2 O4 , and 31.3◦ ,
36.8◦ , 44.8◦ , 59.4◦ , and 65.2◦ for NiCo2 O4 can be assigned to the (220), (311), (400), (511) and (440)
reflections of CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 respectively [45,46]. The pattern of NiCo2 O4 shows a
peak at 43.2◦ , which indicates the formation of NiO cubic phase as also confirmed by TOPAS fitting.
The lattice constants obtained using d-spacing for the sample are a = b = c = 0.816 nm, 0.808 nm, 0.807 nm,
and for CrCo2 O4, MnCo2 O4 , and NiCo2 O4 , respectively. The crystallite size of CrCo2 O4 , MnCo2 O4,
and NiCo2 O4 as calculated using Scherrer’s formula [47] is around 10.57 nm, 14.65 nm, and 19.97 nm for

99
Processes 2020, 8, 343

CrCo2 O4, MnCo2 O4 , and NiCo2 O4 , respectively (Table 1). FTIR spectrum, Figure 1b, further identifies
the structure of the bio-templated MCo2 O4 . The FTIR spectrum displays two distinct bands at 515.7 (ν1 )
and 637 (ν2 ) cm−1 , which arise from the stretching vibrations of the metal-oxygen bonds [48–50]. The ν1
band is characteristic of M-O (M = Cr, Mn, Ni) vibrations in octahedral coordination, and the ν2 band
is attributable to M-O (M - Co) bond vibration in tetrahedral coordination. These frequency bands are
the signature vibrational bands for the spinel lattice [51]. Hence FTIR spectrum at 519.1, 519.02, and
519.08 cm−1 indicate stretching vibration of Co3+ -O2− in the octahedral sites, and at 638.6, 639.9, and
641.3 cm−1 indicate vibration of Cr3+ -O− , Mn2+ -O2− , and Ni2+ -O− at tetrahedral sites for CrCo2 O4 ,
MnCo2 O4 , and NiCo2 O4 , respectively [52]. The presence of vibration bands confirms the development
of pure phase spinal CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 nanostructures.

Table 1. Crystallite size and physical properties of MCo2 O4 (M = Cr, Mn, Ni) determined using XRD,
the Barrette–Joyner–Halenda (BJH) method, and Brunauer–Emmett–Teller (BET) surface area analyzer.

BET Surface BJH Surface BJH Avg. Pore BJH Avg. Pore Crystallite
Sample
Area(m2 /g) Area (m2 /g) Radius (nm) Volume (cc/g) Size (nm)
CrCo2 O4 34.4 46.9 1.135 0.106 10.57
MnCo2 O4 32.2 47.7 0.839 0.071 14.65
NiCo2 O4 18.9 31.8 1.129 0.039 19.97

(a) (b)

(c) (d)

Figure 1. (a) x-ray diffraction pattern, (b) FTIR, (c) adsorption-desorption curve and inset pore volume
distribution, and (d) thermogravimetric curve of tubular MCo2 O4 (M = Cr, Mn, Ni) structures.

Figure 1c shows the BET specific surface area of tubular MCo2 O4 microstructures. The specific
surface area was determined from N2 adsorption-desorption isotherms obtained at 77 K between
relative pressure P/Po ~0.029 to 0.99, and the Barrette–Joyner–Halenda (BJH) method was used for
measuring corresponding pore sized distributions. The type IV isotherm hysteresis loops [53] suggest
the existence of mesopores in the samples. The BET specific surface area of biomorphic CrCo2 O4 ,
MnCo2 O4 , and NiCo2 O4 are 34.4 m2 /g, 32.2 m2 /g, and 18.9 m2 /g, respectively. Figure 1c inset shows
the pore size distribution. Inset curves indicate having a more favorable condition for the fast ion
transport phenomenon within the electrode surface [54–57], which is confirmed by the presence of a
significant number of pores distribution at around 0.4 nm to 4.3 nm with the highest pore volume.

100
Processes 2020, 8, 343

Additionally, the large BET surface area of tubular MCo2 O4 superstructures can provide plenty of
superficial electrochemical active sites to participate in the Faradaic redox reactions.
The thermogravimetric analysis was conducted on the infiltrated samples (cotton dipped in a
mixture of chemical solution and filtered it) to understand the temperature dependence mechanism
of the formation of biomorphic MCo2 O4 . Figure 1d shows the TGA plots of MCo2 O4 measured
in the temperature range of 24 to 550 ◦ C. The formation of MCo2 O4 from the nitrate salts results
in three steps. The weight loss at around 110 ◦ C for all three MCo2 O4 is due to water desorption,
the second weight loss up to 187 ◦ C for CrCo2 O4 , 241 ◦ C for MnCo2 O4 , and 160 ◦ C for NiCo2 O4
is due to burning of cotton and start of decomposition of Co(NO3 )2 ·6H2 O and Cr(NO3 )2 ·6H2 O,
Co(NO3 )2 ·6H2 O and Mn(NO3 )2 ·6H2 O, Co(NO3 )2 ·6H2 O and Ni(NO3 )2 ·6H2 O respectively, there was
no weight loss at beyond 315, 320, and 200 ◦ C which signifies the completion of the formation of
CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 . Upon immersing fiber into the precursor solution, the water and
Cr(NO3 )2 ·6H2 O, Mn(NO3 )2 ·6H2 O, Co(NO3 )2 ·6H2 O and Ni(NO3 )2 ·6H2 O molecules were absorbed
onto the hydroxyl-group-rich cotton fiber substrate. With the heat treatment above 520 ◦ C, nitrate salts
decomposed in the form CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 as follow [58],

2Co(NO3 )2 ·6H2 O + Cr(NO3 )2 ·6H2 O → CrCo2 O4 + 2O2 + 6NO2 + 18H2 O (1)

2Co(NO3 )2 ·6H2 O + Mn(NO3 )2 ·6H2 O → MnCo2 O4 + 2O2 + 6NO2 + 18H2 O (2)

2Co(NO3 )2 ·6H2 O + Ni(NO3 )2 ·6H2 O → NiCo2 O4 + 2O2 + 6NO2 + 18H2 O (3)

With the increase in calcination temperature, the removal of organic substance was achieved
where the remaining few portions of the organic substance change into carbon.
FE-SEM in Figure 2a–d displays tubular morphology of the cotton fibers, samples CrCo2 O4 ,
MnCo2 O4 , and NiCo2 O4 , respectively, which resembles a biomorphic structure. Figure 3a–c shows
SEM images obtained using elemental mapping at chromium, manganese, and nickel energy peaks
and shows that the tubular structure is well decorated with the CrCo2 O4 , MnCo2 O4 , and NiCo2 O4
nanoparticles. Table 2 gives the elemental composition and element distribution, which is obtained via
EDX (energy dispersive x-ray spectroscopy).

Figure 2. SEM images of (a) cotton fiber, (b), (c), and (d) bio-templated tubular MCo2 O4 (M = Cr, Mn,
Ni) structures.

101
Processes 2020, 8, 343

(a)

(b)

(c)

Figure 3. (a), (b), and (c) shows EDX mapping of tubular MCo2 O4 (M = Cr, Mn, Ni) structures, respectively.

Table 2. Elemental composition in wt % for MCo2 O4 (M = Cr, Mn, Ni) obtained using energy dispersive
X-Ray analysis (EDX). The elemental composition is approximately determined using EDX. Ideally, EDX
can prove which elements are abundant in the particles, but not obtain the exact chemical composition.

Co C O Cr Mn Ni
CrCo2 O4 19.5 49.4 23.2 7.8
MnCo2 O4 33.1 34.6 23.5 8.8
NiCo2 O4 27.2 42.4 19.3 11.1

The type of electrolytes and their molar concentration play a vital role in determining the
electrochemical behavior of oxide electrodes [59–61]. Therefore, many aqueous electrolytes such
as sulfates K2 SO4 , H2 SO4 , KNO3 , Na2 SO4 , hydroxyl KOH, NaOH, LiOH, and chlorides KCl,
NaCl have been explored to be used in supercapacitors [62–66]. The ultimate performance of
the electrode is based on the properties of the electrode material and the intercalation efficiency of
the cations [51]. Since KOH electrolyte provides lower electrochemical series resistance with better
conductivity as compared to other electrolytes [67], KOH is chosen as an electrolyte in this study for
the electrochemical measurement.
Cyclic voltammetry and charge-discharge curves were measured to investigate the electrochemical
behavior of MCo2 O4 nanoparticles. Figure 4 displays the CV curves for tubular MCo2 O4 electrodes
measured in the 3M KOH electrolyte. Figure 4a,c, and Figure 4e shows the CV curves measured in
the voltage window of 0.0 to 0.6 V and measured at different scan rate from 2 to 300 mV/s. A pair
of redox peaks associated with the redox reactions involved in the alkaline electrolyte during the
charging and discharging process was observed in all CV plots. The CV curve is asymmetric, which
indicates a quasi-reversible redox reaction [68], the anodic and cathodic peak separation are 0.121 V,
0.124 V, and 0.123 V at 2 mV/s and 0.310 V, 0.186 V, and 0.333 V at 300 mV/s for CrCo2 O4 , MnCo2 O4
and NiCo2 O4 respectively. The presence of anodic and cathodic peaks, indicating the usefulness of

102
Processes 2020, 8, 343

the materials as a pseudocapacitor. Typical pseudo-capacitance behavior of MCo2 O4 nanostructures


arises from the reversible surface or near-surface Faradic reactions for charge storage. The reversible
redox reaction involved in the charge-discharge process for MCo2 O4 can be described as follows by
Equations (5)–(7) [69–71].
MO + OH− ↔ MOOH + e− (4)

CrCo2 O4 + OH− + H2 O ↔ CrOOH + 2CoOOH + e− (5)


− −
MnCo2 O4 + OH + H2 O ↔ MnOOH + 2CoOOH + e (6)
− −
NiCo2 O4 + OH + H2 O ↔ NiOOH + 2CoOOH + e (7)

(a) (b)

(c) (d)

(e) (f)
Figure 4. (a,c,e) show cyclic voltammetry curves of tubular MCo2 O4 (M = Cr, Mn, Ni) electrode
obtained in the scan range of 5 mV/s to 300 mV/s measured in 3M KOH electrolyte. (b,d,f) show cyclic
stability curves measured up to 1000 cycles in 3M KOH electrolyte at scan rate of 40 mV/s.

Pseudocapacitive characteristics of electrodes are indicated by a non-rectangular form of CV


curves. Within the potential range from 0 to 0.6 V, a pair of reversible redox peaks can be observed.
With the increase in the scan rate, a small positive shift of the oxidation peak potential and a negative
shift of the reduction peak potential was observed, which can be primarily attributed to the influence
of the increasing electrochemical polarization as the scan rate scales up. Pairs of reversible redox curve
are indicative of pseudocapacitive behavior of the material with redox peaks attributed to M(II)/M(III)
redox process [72]. The redox potentials and shape of the CV curves are comparable to those reported

103
Processes 2020, 8, 343

for CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 electrodes [24,73–76], suggesting that the measured capacitance
mainly arises from the redox mechanism.
Figure 5a shows the specific capacitance, Csp , as a function of the voltage scan rate of the tubular
MCo2 O4 electrode. The specific capacitance, Csp , was calculated from the CV plots using the following
Equation (8) [77].
 V2
i ∗ V ∗ dV
Csp = V1 (8)
m ∗ v ∗ (V2 − V1)
where V1 and V2 stand for the working potential limits, i stands for the current, m stands for the mass
of the electroactive materials, and v is the scan rate in mV/s.

(a) (b)

(c)

Figure 5. (a) Specific capacitance vs. scan rate, (b) and peak current vs. (scan rate)1/2 , and (c) diffusion
and capacitive contribution to the specific capacitance.

It is evident for this figure that the electrode displays higher Csp up to 407.2 F/g for NiCo2 O4 in
3M KOH electrolyte at 2 mV/s, which is higher than the value that is observed for either CrCo2 O4
(Csp ~ 403.2 F/g) or MnCo2 O4 (Csp ~ 378.1 F/g) electrode, value are given in Table 3. The specific
capacitance for higher scan rates (>50 mV/s) remains practically constant because of limited ion
movement only at the surface of the electrode material. Hence EDLC becomes a dominant mechanism
at higher scan rates. At lower scan rates (<5 mV/s), the majority of active surface are utilized by the
ions for charge storage, and hence resulting in the higher specific capacitance. The CV cyclic stability
of the electrode was tested for 1000 cycles. Figure 4b,d, and Figure 4f show no significant differences in
the CV curves after the 100th, 500th, and 1000th cycle of repetition. The CV curves clearly show that
the current response is proportionally increased with the scan rate, indicating an excellent capacitive
behavior of the electrode materials. This can be ascribed to facile ion diffusion and large specific
surface area of the electrode materials. Furthermore, there is almost no relation between the shape of
CV curves and scan rates, which can be associated with the electron conduction and improved mass
transportation of electrode material [78].

104
Processes 2020, 8, 343

Table 3. Data of specific capacitance, energy density, and power density for MCo2 O4 (M = Cr, Mn, Ni)
obtained from cyclic voltammetry and charge-discharge curves.

CrCo2 O4 MnCo2 O4 NiCo2 O4


Specific Capacitance at 2mv/s 403.2 F/g 378.1 F/g 407.2 F/g
Specific Capacitance at 1A/g 231 F/g 161 F/g 190 F/g
Energy density 11.1 Wh/Kg 7.8 Wh/Kg 9.3 Wh/Kg
Power density 7287.34 W/Kg 7195.33 W/Kg 7186.12 W/Kg

The total stored charge has a contribution from three components; first is the Faradaic contribution
coming from the insertion process of electrolyte ions, second is the faradaic contribution from the
charge-transfer process with surface atoms, and third is pseudocapacitance and nonfaradic contribution
from the double layer effect [79]. Both pseudocapacitance and double-layer charging are substantial,
due to their higher surface area of nanoparticles. The capacitive effects are characterized by analyzing
the cyclic voltammetry data at various scan rates according to [80,81],

i = avb (9)

where i, v, a and b, are peak current (A), voltage scan rate (mV/s), and fitting parameters, respectively.
The charge storage mechanism is defined based on the value of the constant b, where b = 1 defines
capacitive or b = 0.5 defines diffusion-limited charge storage mechanism. Fitting the peak current,
i, vs. square root of the scan rate, SQRT (scan rate), v-1/2 , curves, Figure 5b, with Equation (6), gives b
values of ~ 0.646, 0.711, and 0.648 for CrCo2 O4 , MnCo2 O4 , and NiCo2 O4, respectively. This obtained b
value for our sample MCo2 O4 (M = Cr, Mn, Ni) indicates the diffusive nature of the charge storage
mechanism is prominent for NiCo2 O4 as compared to the other two.
Usually, the contribution to the current response at fixed potential comes from surface capacitive
effects and diffusion-controlled insertion processes [82,83]. These contributions to the specific
capacitance could be separated using the following Equation (10):

Csp = k1 + k2 v−1/2 (10)

For which k1 and k2 can be determined from the Csp vs. v−1/2 linear plot with slope k2 and intercept
k1 . k1 and k2 are fractions of diffusion and capacitive contribution to the net specific capacitance at a
given voltage rate. The Csp was plotted against the slow scan rate up to 20 mV/s, and a regression
fit was performed using Equation (10). The obtained k1 and k2 values were used to determine the
fractional contribution to the net specific capacitance. Figure 5c shows capacitive and diffusive
fractional contributions to net specific capacitance for a slow scan rate of up to 20 mV/s. By comparing
the lower green area with the total capacitance, we find that capacitive effects contribute by 48%, 54%,
and 38% of the total specific capacitance for CrCo2 O4 , MnCo2 O4 , and NiCo2 O4, respectively.
Figure 6a,c, and Figure 6e show the galvanostatic charge-discharge (GCD) plots measured in the
voltage window of 0.0 to 0.6 V at different current densities between 0.75 A/g to 30 A/g in 3M KOH.
From the observed non-linearity between the potential and time, it is confirmed that the capacitance of
the studied materials is not constant over the studied potential ranges. The specific capacitance of
electrodes was calculated using the following Equation (11):

I∗t
Csp = (11)
m ∗ ΔV
where Csp , I, ΔV, m, and t are the specific capacitance (F/g), charge-discharge current (A), the potential
range (V), and the mass of the electroactive materials, and the discharging time (s), respectively.
The GCD curves with a plateau, usually displayed by oxide electrodes, show pseudocapacitive
behavior of electrode with respect to their discharging time for all electrolytes. This typical GCD
behavior could arise from the electrochemical adsorption-desorption of OH- electrolyte and/or a
redox reaction at the interface of electrode/electrolyte [84,85]. It is observed that the discharging time

105
Processes 2020, 8, 343

in biomorphic MCo2 O4 is longer for CrCo2 O4 in the KOH electrolyte. The specific capacitances of
biomorphic CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 at 1 A/g are 231 F/g, 161 F/g, and 190 F/g in 3M KOH
electrolytes, respectively are shown in Table 3. Figure 7a shows the dependence of current density
on the specific capacitance of the electrode material. Usually, insufficient Faradic redox reaction is
achieved at the high discharge current densities. This leads to increased potential drop due to the
resistance of tubular MoCo2 O4 electrode resulting in an observed decrease in capacitance with the
increased discharge current density. This implies ion penetration is feasible at lower current densities
where ions have access to the inner structure, and thus all active area of the electrode. However,
at higher current densities, the effective use of the material is limited to only the outer surface of the
electrode. The specific capacitance of MCo2 O4 electrodes in this study is compared with the literature
values at current density 1 A/g, 2 A/g, and 5 A/g and are listed in Table 4. It is evident from Table 4 that
electrochemical performance of bio-templated MCo2 O4 comparable and, in some cases, outperformed
electrodes prepared via other techniques.

(a) (b)

(c) (d)

(e) (f)

Figure 6. (a,c,e) show charge-discharge (CD) curves of tubular MCo2 O4 (M = Cr, Mn, Ni) electrode
measured in the current density window of 1 to 30 A/g in 3M KOH electrolyte, where red color CD curve
is for 1A/g, blue color CD curve is for 1.5A/g, orange color CD curve is for 2A/g and continuously time
is decreasing with increasing current density. (b,d,f) show cyclic stability (black color) and coulombic
efficiency (blue color) tested at 10 A/g current density for 5000 cycles in 3M KOH electrolytes of MCo2 O4
(M = Cr, Mn, Ni).

106
Processes 2020, 8, 343

 
(a) (b)
Figure 7. (a) Comparison of specific capacitance as a function of current density and (b) Ragone plot of
power density vs. energy density.

Table 4. Comparison of electrochemical performance of MCo2 O4 (M = Cr, Mn, Ni) as available from
the literature.

Cyclic
Electrode Specific Energy Power
Electrolyte Performance Ref.
Material Capacitance Density Density
(retention)
MnCo2 O4 85 % after
108 F/g at 10 A/g 54 Wh/Kg 9851 W/Kg [5]
nanofiber 3000 cycles
Nanorods 92.7% after
1M KOH 349.8 F/g at 1 A/g 35.4 Wh/Kg 225 W/Kg [86]
MnCo2 O4 50 cycles
Nanoneedles 94.3% after
6M KOH 1535 F/g at 1 A/g 35.4 Wh/Kg 225 W/Kg [87]
MnCo2 O4 12000 cycles
Nanorods 90.2% after
2M KOH 845.6 F/g at 1 A/g 35.4 Wh/Kg 225 W/Kg [24]
MnCo2 O4 2000 cycles
Nanorods 88.76% after
1M KOH 308.3 F/g at 1 A/g 55.5 Wh/Kg 5400 W/Kg [88]
MnCo2 O4 2000 cycles
MnCo2 O4
nanowires - 2262 F/g at 1 A/g 85.7 Wh/Kg 800 W/Kg - [89]
@MnO2
Nanorods 84 % after
718 F/g at 0.5 A/g - - [90]
MnCo2 O4 1000 cycles
NiCo2 O4 77.6% after
2M KOH 565 F/g at 1 A/g - - [69]
nanorods 1000 cycles
RGO decorated
95% after
nanorods 6M KOH 1278F/g at 1 A/g - - [84]
1000 cycles
bundle NiCo2 O4
Nanorods
101.7% after
assemble 2M KOH 764 F/g at 2 A/g - - [91]
1500 cycles
NiCo2 O4
Nanorods 101.7% after
2M KOH 823 F/g at 0.823 A/g 28.51 Wh/Kg - [2]
NiCo2 O4 1500 cycles
GO/ Nanorods 94.3% after
1M KOH 709.7 F/g at 1 A/g 28 Wh/Kg 8000 W/Kg [92]
NiCo2 O4 5000 cycles
Nanorods 80% after
2M KOH 600 F/g at 5 A/g - - [70]
NiCo2 O4 1500 cycles
Nanorods
91% after
NiCo2 O4 1M H2 So4 901 F/g at 5 A/g - - [93]
3000 cycles
@PANI
92% after
5000 cycles
Templated 403.4, 378, and 407.2
7287.3, 91% after
CrCo2 O4, F/g at 2mV/s and 11.1, 7.8, and
3M 7195.3, and 5000 cycles
MnCo2 O4, and 231, 161, and 190 F/g 9.3 Wh/kg, [This work]
KOH 7186.1 W/Kg, and 100%
NiCo2 O4 at 1 A/g, respectively
respectively after
microstructure respectively
5000 cycles,
respectively.

107
Processes 2020, 8, 343

Estimation of the electrochemical utilization of the active materials (CrCo2 O4 , MnCo2 O4, and
NiCo2 O4 electrode), was evaluated from the fraction of cobalt sites, z. The fraction, z, can be evaluated
using Faraday’s law as following [94]:

z = Csp MW ΔV/F (12)

where Csp , MW, ΔV and F is the specific capacitance value, the molecular weight, the applied potential
window, and the Faraday’s constant, respectively. The z value of 1 indicates the complete involvement
of electroactive material. i.e., all active metal sites participating in the redox process. The molecular
weight of Co (84.03 g/mol) and Cr (51.996 g/mol) in CrCo2 O4 , Co (84.03 g/mol) and Mn (54.93 g/mol) in
MnCo2 O4 , and Co (84.03 g/mol) and Ni (58.69 g/mol) in NiCo2 O4 the specific capacitance at a current
density of 1 A/g (Csp~ 231 F/g, 161 F/g, and 190 F/g) for CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 , Figure 5b,
and a potential window of 0.6 V gives a z value of 0.195, 0.139, and 0.169 respectively. In other words,
~20%, 14%, and 17% of the total active material Cr, Mn, and Ni atoms participate in the redox reaction
for the charge storage. The observed low value of z suggests that the charge storage in tubular MCo2 O4
structure via a redox reaction process occurs mainly at the surface with little bulk interaction due to
diffusion of OH− ions into the material. It could be concluded that the charge storage due to the redox
process in MCo2 O4 mostly occurs only at the redox sites predominantly located on the surface of the
particles [71].
Cyclic stability tests were performed to evaluate the practical performance of electrodes as a
supercapacitor. The stability test of electrode materials was assessed via galvanostatic CD measurement
for 5000 cycles for a current density of 10 A g−1 in 3M KOH and is shown in Figure 6b,d, and Figure 6f.
The Coulombic efficiency (η) of the devices was calculated from its charging (Tc ) and the discharging
(Td ) times from GCD curves following the relation, η = Td /Tc × 100, and is plotted in Figure 6b,d,f
as a function of cyclic time. The initial η of the device was ~100%, which remained practically the
same even after 5000 cycles. For the practical applications, the study of cycling performance for
electroactive material is very significant parameter. The percentage retention in specific capacitance
was calculated using,

% retention in specific capacitance = (C# /C1 )×100 (13)

where C# and C1 are specific capacitance at various cycles and the 1st cycle, respectively. The specific
capacitance of the electrode is reduced by 7.68% in CrCo2 O4 , reduced by 9.12% in MnCo2 O4 , and
increased by 0.48% in NiCo2 O4 .
Figure 5h shows the Ragone plots of as-synthesized MCo2 O4 electrodes. The energy densities (E)
and power densities (P) of the electrochemical cells are calculated using the following equations [95]:

E = (1/2)CV 2 (14)

P = E/t (15)

where C, V, and t are the specific capacitance that depends on the mass of the electrodes, the operating
voltage of the cell, and discharge time in seconds, respectively. The essential point for high-performance
supercapacitors is to obtain a high energy density and meanwhile providing an outstanding power
density. It is observed from Figure 7b that the tubular CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 electrode
display superior performance over energy density up to 11.1 Wh/kg, 7.8 Wh/kg, and 9.3 Wh/kg with
a peak power density up to 7287.34 W/kg, 7195.33 W/kg, and 7186.12 W/kg, respectively are given
in Table 3. As supercapacitor is expected to provide higher power and energy density at the same
time, hence NiCo2 O4 displays an overall better energy density of 9.3 Wh/kg and a power density of
7186.12 W/kg as compared to MnCo2 O4 and CrCo2 O4 .

108
Processes 2020, 8, 343

Ideally, to develop a higher energy density battery with a given anode, a cathode with high
electrochemical interaction potential is desired. This is because the energy density of the device equals
the product of the working voltage, which is obtained from the electrochemical potentials different
between the cathode and anode and specific capacity of the electrode materials [63]. On the other hand,
the theoretical capacity of electrode materials depend on the number of reactive electrons (n) and molar
weight (M) of the materials and is expressed as Equation (16) [96],

Ct = n∗F/3.6∗M (16)

Here, F is the Faraday constant, n is the number of reactive electrons, and M is the molar weight
of materials. Theoretically, the equation predicts that an electrode material having a smaller molecular
weight can produce a higher capacity. The molecular weight of CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 is
~ 233.86, 236.80, and 240.55 g/mol, respectively. Thus, in line with the Equation (11), at low current
density, among three electrodes studied, the CrCo2 O4 displays higher specific capacitance. However,
overall superior performance in terms of energy and power density was displayed by the NiCo2 O4
electrode. NiCo2 O4 is known to possess rich electroactive sites, narrow pore size, and higher electrical
conductivity (at least two magnitudes higher) than that of Co3 O4 and NiO, which could be the reason
for the observed overall better performance of NiCo2 O4 [97,98].

4. Conclusions
In conclusion, biomorphic tubular CrCo2 O4 , MnCo2 O4 , and NiCo2 O4 nanostructures were
prepared using cotton by a cost-effective and straightforward bio-template method. The synthesized
tubular MCo2 O4 display excellent crystallinity, phase purity and display desirable electrochemical
properties, which indicate a good chance for the fabrication of high-performance supercapacitor
devices. Electrodes constructed using the tubular MCo2 O4 demonstrate high specific capacitance,
cyclic stability, power, and energy density when evaluated in 3M KOH electrolyte. The study suggests
that it is imperative to account for the nature of the electroactive sites and the conductivity of materials
when choosing materials from the series of transition metal oxide as the electrode for the supercapacitor
application. Furthermore, the superior electrochemical performance of tubular MCo2 O4 microstructure
owes to the presence of conducting a carbonaceous structure. The highly porous carbonaceous structure
can allow electrolyte access throughout the electrode structure. Thus, it produces a large surface area
for ion transfer between the electrolyte and the active materials, which leads to achieving ultrafast
storage and release of energy.

Author Contributions: Conceptualization, D.G.; methodology, D.G., and C.Z.; software, D.G. and C.Z.; validation,
S.R.M., R.K.G., and D.G.; formal analysis, D.G.; investigation, S.R.M. and R.K.G.; resources S.R.M.; data
curation, D.G., C.Z., and R.K.G.; writing—original draft preparation, D.G.; writing—review and editing, S.R.M.;
visualization, S.R.M. and D.G.; supervision, S.R.M. and R.K.G.; project administration, S.R.M.; funding acquisition,
S.R.M. All authors have read and agreed to the published version of the manuscript.
Funding: This is supported by the grants from FIT-DRONES and Biologistics at the University of Memphis,
Memphis; TN. Dr. Ram K. Gupta expresses his sincere acknowledgment of the Polymer Chemistry Initiative at
Pittsburg State University for providing financial and research support.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Suktha, P.; Chiochan, P.; Iamprasertkun, P.; Wutthiprom, J.; Phattharasupakun, N.; Suksomboon, M.;
Kaewsongpol, T.; Sirisinudomkit, P.; Pettong, T.; Sawangphruk, M. High-performance supercapacitor of
functionalized carbon fiber paper with high surface ionic and bulk electronic conductivity: Effect of organic
functional groups. Electrochim. Acta 2015, 176, 504–513. [CrossRef]
2. Sahoo, S.; Ratha, S.; Rout, C.S. Spinel NiCo2 O4 nanorods for supercapacitor applications. Am. J. Eng. Appl.
Sci. 2015, 8, 371–379. [CrossRef]

109
Processes 2020, 8, 343

3. El-Kady, M.F.; Strong, V.; Dubin, S.; Kaner, R.B. Laser scribing of high-performance and flexible
graphene-based electrochemical capacitors. Science 2012, 335, 1326–1330. [CrossRef]
4. Zhu, Y.; Murali, S.; Stoller, M.D.; Ganesh, K.J.; Cai, W.; Ferreira, P.J.; Pirkle, A.; Wallace, R.M.; Cychosz, K.A.;
Thommes, M.; et al. Carbon-based supercapacitors produced by activation of graphene. Science 2011, 332,
1537–1541. [CrossRef]
5. Pettong, T.; Iamprasertkun, P.; Krittayavathananon, A.; Sukha, P.; Sirisinudomkit, P.; Seubsai, A.;
Chareonpanich, M.; Kongkachuichay, P.; Limtrakul, J.; Sawangphruk, M. High-performance asymmetric
supercapacitors of MnCo2 O4 nanofibers and N-doped reduced graphene oxide aerogel. ACS Appl. Mater.
Interfaces 2016, 8, 34045–34053. [CrossRef] [PubMed]
6. Fan, L.Z.; Qiao, S.; Song, W.; Wu, M.; He, X.; Qu, X. Effects of the functional groups on the electrochemical
properties of ordered porous carbon for supercapacitors. Electrochim. Acta 2013, 105, 299–304. [CrossRef]
7. Chen, X.L.; Li, W.S.; Tan, C.L.; Li, W.; Wu, Y.Z. Improvement in electrochemical capacitance of carbon
materials by nitric acid treatment. J. Power Sources 2008, 184, 668–674. [CrossRef]
8. Yang, Q.; Lu, Z.; Li, T.; Sun, X.; Liu, J. Hierarchical construction of core–shell metal oxide nanoarrays with
ultrahigh areal capacitance. Nano Energy 2014, 7, 170–178. [CrossRef]
9. Vilatela, J.J.; Eder, D. Nanocarbon composites and hybrids in sustainability: A review. ChemSusChem 2012, 5,
456–478. [CrossRef]
10. Nguyen, T.; Montemor, M.D.F. Metal Oxide and Hydroxide–Based Aqueous Supercapacitors: From Charge
Storage Mechanisms and Functional Electrode Engineering to Need-Tailored Devices. Adv. Sci. 2019, 6,
1801797. [CrossRef]
11. Alqahtani, D.M.; Zequine, C.; Ranaweera, C.K.; Siam, K.; Kahol, P.K.; Poudel, T.P.; Mishra, S.R.; Gupta, R.K.
Effect of metal ion substitution on electrochemical properties of cobalt oxide. J. Alloy. Compd. 2019, 771,
951–959. [CrossRef]
12. Pendashteh, A.; Moosavifard, S.E.; Rahmanifar, M.S.; Wang, Y.; El-Kady, M.F.; Kaner, R.B.; Mousavi, M.F.
Highly ordered mesoporous CuCo2 O4 nanowires, a promising solution for high-performance supercapacitors.
Chem. Mater. 2015, 27, 3919–3926. [CrossRef]
13. Cheng, H.; Lu, Z.G.; Deng, J.Q.; Chung, C.Y.; Zhang, K.; Li, Y.Y. A facile method to improve the high rate
capability of Co3 O4 nanowire array electrodes. Nano Res. 2010, 3, 895–901. [CrossRef]
14. Yan, D.; Zhang, H.; Chen, L.; Zhu, G.; Li, S.; Xu, H.; Yu, A. Biomorphic synthesis of mesoporous Co3 O4
microtubules and their pseudocapacitive performance. ACS Appl. Mater. Interfaces 2014, 6, 15632–15637.
[CrossRef] [PubMed]
15. Sieber, H. Biomimetic synthesis of ceramics and ceramic composites. Mater. Sci. Eng. A 2005, 412, 43–47.
[CrossRef]
16. Habibi, Y.; Lucia, L.A.; Rojas, O.J. Cellulose nanocrystals: Chemistry, self-assembly, and applications. Chem.
Rev. 2010, 110, 3479–3500. [CrossRef]
17. Liu, Y.; Lv, B.; Li, P.; Chen, Y.; Gao, B.; Lin, B. Biotemplate-assisted hydrothermal synthesis of tubular porous
Co3 O4 with excellent charge-discharge cycle stability for supercapacitive electrodes. Mater. Lett. 2018, 210,
231–234. [CrossRef]
18. Fan, T.X.; Chow, S.K.; Zhang, D. Biomorphic mineralization: From biology to materials. Prog. Mater. Sci.
2009, 54, 542–659. [CrossRef]
19. Sotiropoulou, S.; Sierra-Sastre, Y.; Mark, S.S.; Batt, C.A. Biotemplated nanostructured materials. Chem. Mater.
2008, 20, 821–834. [CrossRef]
20. Zhou, H.; Fan, T.; Zhang, D. Biotemplated materials for sustainable energy and environment: Current status
and challenges. ChemSusChem 2011, 4, 1344–1387. [CrossRef]
21. Shim, H.W.; Lim, A.H.; Kim, J.C.; Jang, E.; Seo, S.D.; Lee, G.H.; Kim, T.D.; Kim, D.W. Scalable one-pot
bacteria-templating synthesis route toward hierarchical, porous-Co3 O4 superstructures for supercapacitor
electrodes. Sci. Rep. 2013, 3, 2325. [CrossRef] [PubMed]
22. Han, L.; Yang, D.P.; Liu, A. Leaf-templated synthesis of 3D hierarchical porous cobalt oxide nanostructure
as direct electrochemical biosensing interface with enhanced electrocatalysis. Biosens. Bioelectron. 2015, 63,
145–152. [CrossRef]
23. Song, P.; Zhang, H.; Han, D.; Li, J.; Yang, Z.; Wang, Q. Preparation of biomorphic porous LaFeO3 by sorghum
straw biotemplate method and its acetone sensing properties. Sens. Actuators B Chem. 2014, 196, 140–146.
[CrossRef]

110
Processes 2020, 8, 343

24. Weatherspoon, M.R.; Cai, Y.; Crne, M.; Srinivasarao, M.; Sandhage, K.H. 3D Rutile Titania-Based Structures
with Morpho Butterfly Wing Scale Morphologies. Angew. Chem. Int. Ed. 2008, 47, 7921–7923. [CrossRef]
[PubMed]
25. Yan, D.; Li, S.; Zhu, G.; Wang, Z.; Xu, H.; Yu, A. Synthesis and pseudocapacitive behaviors of biomorphic
mesoporous tubular MnO2 templated from cotton. Mater. Lett. 2013, 95, 164–167. [CrossRef]
26. Mondal, A.K.; Su, D.; Chen, S.; Ung, A.; Kim, H.S.; Wang, G. Mesoporous MnCo2 O4 with a flake-like
structure as advanced electrode materials for lithium-ion batteries and supercapacitors. Chem. A Eur. J. 2015,
21, 1526–1532. [CrossRef]
27. Xu, J.; Sun, Y.; Lu, M.; Wang, L.; Zhang, J.; Tao, E.; Qian, J.; Liu, X. Fabrication of the porous MnCo2 O4
nanorod arrays on Ni foam as an advanced electrode for asymmetric supercapacitors. Acta Mater. 2018, 152,
162–174. [CrossRef]
28. Dong, Y.; Wang, Y.; Xu, Y.; Chen, C.; Wang, Y.; Jiao, L.; Yuan, H. Facile synthesis of hierarchical nanocage
MnCo2 O4 for high performance supercapacitor. Electrochim. Acta 2017, 225, 39–46. [CrossRef]
29. Yang, J.; Cho, M.; Lee, Y. Synthesis of hierarchical NiCo2 O4 hollow nanorods via sacrificial-template accelerate
hydrolysis for electrochemical glucose oxidation. Biosens. Bioelectron. 2016, 75, 15–22. [CrossRef]
30. Zhao, Z.; Geng, F.; Bai, J.; Cheng, H.M. Facile and controlled synthesis of 3D nanorods-based urchinlike
and nanosheets-based flowerlike cobalt basic salt nanostructures. J. Phys. Chem. C 2007, 111, 3848–3852.
[CrossRef]
31. Jadhav, A.R.; Bandal, H.A.; Kim, H. NiCo2 O4 hollow sphere as an efficient catalyst for hydrogen generation
by NaBH4 hydrolysis. Mater. Lett. 2017, 198, 50–53. [CrossRef]
32. Huang, W.; Lin, T.; Cao, Y.; Lai, X.; Peng, J.; Tu, J. Hierarchical NiCo2 O4 hollow sphere as a peroxidase
mimetic for colorimetric detection of H2 O2 and glucose. Sensors 2017, 17, 217. [CrossRef]
33. Pu, J.; Wang, J.; Jin, X.; Cui, F.; Sheng, E.; Wang, Z. Porous hexagonal NiCo2 O4 nanoplates as electrode
materials for supercapacitors. Electrochim. Acta 2013, 106, 226–234. [CrossRef]
34. Park, J.; Shen, X.; Wang, G. Solvothermal synthesis and gas-sensing performance of Co3 O4 hollow nanospheres.
Sens. Actuators B Chem. 2009, 136, 494–498. [CrossRef]
35. Lasheras, X.; Insausti, M.; Gil de Muro, I.; Garaio, E.; Plazaola, F.; Moros, M.; De Matteis, L.; M de la Fuente, J.;
Lezama, L. Chemical synthesis and magnetic properties of monodisperse nickel ferrite nanoparticles for
biomedical applications. J. Phys. Chem. C 2016, 120, 3492–3500. [CrossRef]
36. Kuang, L.; Ji, F.; Pan, X.; Wang, D.; Chen, X.; Jiang, D.; Zhang, Y.; Ding, B. Mesoporous MnCo2 O4 nanoneedle
arrays electrode for high-performance asymmetric supercapacitor application. Chem. Eng. J. 2017, 315,
491–499. [CrossRef]
37. Melot, B.C.; Tarascon, J.M. Design and preparation of materials for advanced electrochemical storage.
Acc. Chem. Res. 2013, 46, 1226–1238. [CrossRef]
38. Chen, K.; Xue, D. Materials chemistry toward electrochemical energy storage. J. Mater. Chem. A 2016, 4,
7522–7537. [CrossRef]
39. Principles and Applications of Lithium Secondary Batteries; Park, J.K. (Ed.) John Wiley & Sons: Hoboken, NJ,
USA, 2012.
40. Zhang, Y.; Ma, M.; Yang, J.; Su, H.; Huang, W.; Dong, X. Selective synthesis of hierarchical mesoporous spinel
NiCo2 O4 for high-performance supercapacitors. Nanoscale 2014, 6, 4303–4308. [CrossRef]
41. Yuan, C.; Wu, H.B.; Xie, Y.; Lou, X.W. Mixed transition-metal oxides: Design, synthesis, and energy-related
applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [CrossRef]
42. Zhang, L.L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38,
2520–2531. [CrossRef]
43. Zheng, Y.Z.; Ding, H.; Uchaker, E.; Tao, X.; Chen, J.F.; Zhang, Q.; Cao, G. Nickel-mediated polyol synthesis of
hierarchical V2 O5 hollow microspheres with enhanced lithium storage properties. J. Mater. Chem. A 2015, 3,
1979–1985. [CrossRef]
44. Lithium Batteries: Science and Technology; Nazri, G.A.; Pistoia, G. (Eds.) Springer Science & Business Media:
Berlin/Heidelberg, Germany, 2008.
45. Li, J.; Wang, J.; Liang, X.; Zhang, Z.; Liu, H.; Qian, Y.; Xiong, S. Hollow MnCo2 O4 submicrospheres with
multilevel interiors: From mesoporous spheres to yolk-in-double-shell structures. ACS Appl. Mater. Interfaces
2013, 6, 24–30. [CrossRef] [PubMed]

111
Processes 2020, 8, 343

46. Bhojane, P.; Sen, S.; Shirage, P.M. Enhanced electrochemical performance of mesoporous NiCo2 O4 as an
excellent supercapacitive alternative energy storage material. Appl. Surf. Sci. 2016, 377, 376–384. [CrossRef]
47. Jenkins, R.; Snyder, R.L. Introduction to X-ray Powder Diffractometry; (No. 543.427 JEN); Wiley: Hoboken, NJ,
USA, 1996.
48. Lin, H.K.; Chiu, H.C.; Tsai, H.C.; Chien, S.H.; Wang, C.B. Synthesis, characterization and catalytic oxidation
of carbon monoxide over cobalt oxide. Catal. Lett. 2003, 88, 169–174. [CrossRef]
49. Spencer, C.D.; Schroeer, D. Mössbauer study of several cobalt spinels using Co57 and Fe57 . Phys. Rev. B 1974,
9, 3658. [CrossRef]
50. Kurtulus, F.; Guler, H. A Simple Microwave-Assisted Route to Prepare Black Cobalt, Co3 O4 . Inorg. Mater.
2005, 41, 483–485. [CrossRef]
51. St, G.C.; Stoyanova, M.; Georgieva, M.; Mehandjiev, D. Preparation and characterization of a higher cobalt
oxide. Mater. Chem. Phys. 1999, 60, 39–43.
52. Khairy, M.; Mousa, M. Synthesis of Ternary and Quaternary metal oxides based on Ni, Mn, Cu, and Co for
high-performance Supercapacitor. J. Ovonic Res. 2019, 15, 181–198.
53. Štěpánek, F.; Marek, M.; Adler, P.M. Modeling capillary condensation hysteresis cycles in reconstructed
porous media. Aiche J. 1999, 45, 1901–1912. [CrossRef]
54. Wang, R.; Li, Q.; Cheng, L.; Li, H.; Wang, B.; Zhao, X.S.; Guo, P. Electrochemical properties of manganese
ferrite-based supercapacitors in aqueous electrolyte: The effect of ionic radius. Colloid. Surfaces A Physicochem.
Eng. Aspects 2014, 457, 94–99. [CrossRef]
55. Hou, L.; Yuan, C.; Yang, L.; Shen, L.; Zhang, F.; Zhang, X. Urchin-like Co3 O4 microspherical hierarchical
superstructures constructed by one-dimension nanowires toward electrochemical capacitors. Rsc Adv. 2011,
1, 1521–1526. [CrossRef]
56. Yuan, C.; Yang, L.; Hou, L.; Shen, L.; Zhang, X.; Lou, X.W.D. Growth of ultrathin mesoporous Co3 O4
nanosheet arrays on Ni foam for high-performance electrochemical capacitors. Energy Environ. Sci. 2012, 5,
7883–7887. [CrossRef]
57. Adhikari, H.; Ghimire, M.; Ranaweera, C.K.; Bhoyate, S.; Gupta, R.K.; Alam, J.; Mishra, S.R. Synthesis and
electrochemical performance of hydrothermally synthesized Co3 O4 nanostructured particles in presence of
urea. J. Alloy. Compd. 2017, 708, 628–638. [CrossRef]
58. Sun, D.; He, L.; Chen, R.; Liu, Y.; Lv, B.; Lin, S.; Lin, B. Biomorphic composites composed of octahedral
Co3O4 nanocrystals and mesoporous carbon microtubes templated from cotton for excellent supercapacitor
electrodes. Appl. Surf. Sci. 2019, 465, 232–240. [CrossRef]
59. Brousse, T.; Bélanger, D. A Hybrid Fe3 O4 MnO2 Capacitor in Mild Aqueous Electrolyte. Electrochem.
Solid-State Lett. 2003, 6, A244–A248. [CrossRef]
60. Wang, S.Y.; Ho, K.C.; Kuo, S.L.; Wu, N.L. Investigation on capacitance mechanisms of Fe3 O4 electrochemical
capacitors. J. Electrochem. Soc. 2006, 153, A75–A80. [CrossRef]
61. Tiruye, G.A.; Munoz-Torrero, D.; Palma, J.; Anderson, M.; Marcilla, R. All-solid state supercapacitors
operating at 3.5 V by using ionic liquid based polymer electrolytes. J. Power Sources 2015, 279, 472–480.
[CrossRef]
62. Gao, F.; Shao, G.; Qu, J.; Lv, S.; Li, Y.; Wu, M. Tailoring of porous and nitrogen-rich carbons derived from
hydrochar for high-performance supercapacitor electrodes. Electrochim. Acta 2015, 155, 201–208. [CrossRef]
63. Selvam, M.; Srither, S.R.; Saminathan, K.; Rajendran, V. Chemically and electrochemically prepared
graphene/MnO2 nanocomposite electrodes for zinc primary cells: A comparative study. Ionics 2015, 21,
791–799. [CrossRef]
64. Tang, Y.; Liu, Y.; Yu, S.; Gao, F.; Zhao, Y. Comparative study on three commercial carbons for supercapacitor
applications. Russ. J. Electrochem. 2015, 51, 77–85. [CrossRef]
65. Sankar, K.V.; Selvan, R.K. Improved electrochemical performances of reduced graphene oxide based
supercapacitor using redox additive electrolyte. Carbon 2015, 90, 260–273. [CrossRef]
66. Sahu, V.; Shekhar, S.; Sharma, R.K.; Singh, G. Ultrahigh performance supercapacitor from lacey reduced
graphene oxide nanoribbons. ACS Appl. Mater. Interfaces 2015, 7, 3110–3116. [CrossRef] [PubMed]
67. Fic, K.; Platek, A.; Piwek, J.; Frackowiak, E. Sustainable materials for electrochemical capacitors. Mater. Today
2018, 21, 437–454. [CrossRef]
68. Meher, S.K.; Justin, P.; Rao, G.R. Nanoscale morphology dependent pseudocapacitance of NiO: Influence of
intercalating anions during synthesis. Nanoscale 2011, 3, 683–692. [CrossRef]

112
Processes 2020, 8, 343

69. Xiao, J.; Yang, S. Sequential crystallization of sea urchin-like bimetallic (Ni, Co) carbonate hydroxide and
its morphology conserved conversion to porous NiCo2 O4 spinel for pseudocapacitors. Rsc Adv. 2011, 1,
588–595. [CrossRef]
70. Liu, X.Y.; Zhang, Y.Q.; Xia, X.H.; Shi, S.J.; Lu, Y.; Wang, X.L.; Gu, C.D.; Tu, J.P. Self-assembled porous NiCo2 O4
hetero-structure array for electrochemical capacitor. J. Power Sources 2013, 239, 157–163. [CrossRef]
71. Wang, X.; Han, X.; Lim, M.; Singh, N.; Gan, C.L.; Jan, M.; Lee, P.S. Nickel cobalt oxide-single wall carbon
nanotube composite material for superior cycling stability and high-performance supercapacitor application.
J. Phys. Chem. C 2012, 116, 12448–12454. [CrossRef]
72. Liu, X.; Long, Q.; Jiang, C.; Zhan, B.; Li, C.; Liu, S.; Zhao, Q.; Huang, W.; Dong, X. Facile and green synthesis
of mesoporous Co3 O4 nanocubes and their applications for supercapacitors. Nanoscale 2013, 5, 6525–6529.
[CrossRef]
73. Zhu, Y.; Pu, X.; Song, W.; Wu, Z.; Zhou, Z.; He, X.; Lu, F.; Jing, M.; Tang, B.; Ji, X. High capacity NiCo2 O4
nanorods as electrode materials for supercapacitor. J. Alloy. Compd. 2014, 617, 988–993. [CrossRef]
74. Jokar, E.; Shahrokhian, S. Synthesis and characterization of NiCo2 O4 nanorods for preparation of
supercapacitor electrodes. J. Solid State Electrochem. 2015, 19, 269–274. [CrossRef]
75. Veeramani, V.; Madhu, R.; Chen, S.M.; Sivakumar, M.; Hung, C.T.; Miyamoto, N.; Liu, S.B. NiCo2 O4 -decorated
porous carbon nanosheets for high-performance supercapacitors. Electrochim. Acta 2017, 247, 288–295.
[CrossRef]
76. Zhai, Y.; Mao, H.; Liu, P.; Ren, X.; Xu, L.; Qian, Y. Facile fabrication of hierarchical porous rose-like NiCo2 O4
nanoflake/MnCo2 O4 nanoparticle composites with enhanced electrochemical performance for energy storage.
J. Mater. Chem. A 2015, 3, 16142–16149. [CrossRef]
77. Ghosh, D.; Giri, S.; Das, C.K. Hydrothermal synthesis of platelet β Co(OH)2 and Co3 O4 : Smart electrode
material for energy storage application. Environ. Prog. Sustain. Energy 2014, 33, 1059–1064. [CrossRef]
78. Yi, H.; Wang, H.; Jing, Y.; Peng, T.; Wang, X. Asymmetric supercapacitors based on carbon nanotubes@ NiO
ultrathin nanosheets core-shell composites and MOF-derived porous carbon polyhedrons with super-long
cycle life. J. Power Sources 2015, 285, 281–290. [CrossRef]
79. Shen, B.; Guo, R.; Lang, J.; Liu, L.; Liu, L.; Yan, X. A high-temperature flexible supercapacitor based on
pseudocapacitive behavior of FeOOH in an ionic liquid electrolyte. J. Mater. Chem. A 2016, 4, 8316–8327.
[CrossRef]
80. Augustyn, V.; Come, J.; Lowe, M.A.; Kim, J.W.; Taberna, P.L.; Tolbert, S.H.; Abruña, H.D.; Simon, P.; Dunn, B.
High-rate electrochemical energy storage through Li+ intercalation pseudocapacitance. Nat. Mater. 2013, 12,
518. [CrossRef]
81. Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.E. Li+ ion
insertion in TiO2 (anatase). 1. Chronoamperometry on CVD films and nanoporous films. J. Phys. Chem. B
1997, 101, 7710–7716. [CrossRef]
82. Liu, T.C.; Pell, W.G.; Conway, B.E.; Roberson, S.L. Behavior of molybdenum nitrides as materials for
electrochemical capacitors comparison with ruthenium oxide. J. Electrochem. Soc. 1998, 145, 1882–1888.
[CrossRef]
83. Wang, J.; Polleux, J.; Lim, J.; Dunn, B. Pseudocapacitive contributions to electrochemical energy storage in
TiO2 (anatase) nanoparticles. J. Phys. Chem. C 2007, 111, 14925–14931. [CrossRef]
84. Zhao, D.D.; Bao, S.J.; Zhou, W.J.; Li, H.L. Preparation of hexagonal nanoporous nickel hydroxide film and its
application for electrochemical capacitor. Electrochem. Commun. 2007, 9, 869–874. [CrossRef]
85. Sugimoto, W.; Iwata, H.; Yasunaga, Y.; Murakami, Y.; Takasu, Y. Preparation of ruthenic acid nanosheets
and utilization of its interlayer surface for electrochemical energy storage. Angew. Chem. Int. Ed. 2003, 42,
4092–4096. [CrossRef] [PubMed]
86. Li, L.; Zhang, Y.Q.; Liu, X.Y.; Shi, S.J.; Zhao, X.Y.; Zhang, H.; Ge, X.; Cai, G.F.; Gu, C.D.; Wang, X.L.; et al.
One-dimension MnCo2 O4 nanowire arrays for electrochemical energy storage. Electrochim. Acta 2014, 116,
467–474. [CrossRef]
87. Hui, K.N.; San Hui, K.; Tang, Z.; Jadhav, V.V.; Xia, Q.X. Hierarchical chestnut-like MnCo2 O4 nanoneedles
grown on nickel foam as binder-free electrode for high energy density asymmetric supercapacitors. J. Power
Sources 2016, 330, 195–203. [CrossRef]
88. Shi, X.; Liu, Z.; Zheng, Y.; Zhou, G. Controllable synthesis and electrochemical properties of MnCo2 O4
nanorods and microcubes. Colloids Surf. A: Physicochem. Eng. Asp. 2017, 522, 525–535. [CrossRef]

113
Processes 2020, 8, 343

89. Liu, S.; Hui, K.S.; Hui, K.N. 1 D hierarchical MnCo2 O4 nanowire@ MnO2 sheet core–shell arrays on graphite
paper as superior electrodes for asymmetric supercapacitors. ChemNanoMat 2015, 1, 593–602. [CrossRef]
90. Venkatachalam, V.; Alsalme, A.; Alghamdi, A.; Jayavel, R. High performance electrochemical capacitor based
on MnCo2 O4 nanostructured electrode. J. Electroanal. Chem. 2015, 756, 94–100. [CrossRef]
91. Zhu, Y.; Ji, X.; Yin, R.; Hu, Z.; Qiu, X.; Wu, Z.; Liu, Y. Nanorod-assembled NiCo2 O4 hollow microspheres
assisted by an ionic liquid as advanced electrode materials for supercapacitors. Rsc Adv. 2017, 7, 11123–11128.
[CrossRef]
92. Mao, J.W.; He, C.H.; Qi, J.Q.; Zhang, A.B.; Sui, Y.W.; He, Y.Z.; Meng, Q.K.; Wei, F.X. An Asymmetric
Supercapacitor with Mesoporous NiCo2 O4 Nanorod/Graphene Composite and N-Doped Graphene
Electrodes. J. Electron. Mater. 2018, 47, 512–520. [CrossRef]
93. Jabeen, N.; Xia, Q.; Yang, M.; Xia, H. Unique core–shell nanorod arrays with polyaniline deposited into
mesoporous NiCo2 O4 support for high-performance supercapacitor electrodes. Acs Appl. Mater. Interfaces
2016, 8, 6093–6100. [CrossRef]
94. Srinivasan, V.; Weidner, J.W. Studies on the capacitance of nickel oxide films: Effect of heating temperature
and electrolyte concentration. J. Electrochem. Soc. 2000, 147, 880–885. [CrossRef]
95. Lin, C.; Ritter, J.A.; Popov, B.N. Characterization of sol-gel-derived cobalt oxide xerogels as electrochemical
capacitors. J. Electrochem. Soc. 1998, 145, 4097–4103. [CrossRef]
96. Zhi, M.; Xiang, C.; Li, J.; Li, M.; Wu, N. Nanostructured carbon–metal oxide composite electrodes for
supercapacitors: A review. Nanoscale 2013, 5, 72–88. [CrossRef] [PubMed]
97. Wu, Z.; Zhu, Y.; Ji, X. NiCo2 O4 -based materials for electrochemical supercapacitors. J. Mater. Chem. A 2014,
2, 14759–14772. [CrossRef]
98. Bitla, Y.; Chin, Y.Y.; Lin, J.C.; Van, C.N.; Liu, R.; Zhu, Y.; Liu, H.J.; Zhan, Q.; Lin, H.J.; Chen, C.T.; et al. Origin
of metallic behavior in NiCo 2 O 4 ferrimagnet. Sci. Rep. 2015, 5, 15201. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

114
processes
Article
Investigation and Improvement of Scalable Oxygen
Reducing Cathodes for Microbial Fuel Cells by
Spray Coating
Thorben Muddemann 1, *,† , Dennis Haupt 2,† , Bolong Jiang 1 , Michael Sievers 2 and Ulrich Kunz 1
1 Institute of Chemical and Electrochemical Process Engineering, Clausthal University of Technology,
38678 Clausthal-Zellerfeld, Germany; [email protected] (B.J.); [email protected] (U.K.)
2 CUTEC Research Center for Environmental Technologies, 38678 Clausthal-Zellerfeld, Germany;
[email protected] (D.H.); [email protected] (M.S.)
* Correspondence: [email protected]
† These authors contributed equally to this work.

Received: 5 November 2019; Accepted: 17 December 2019; Published: 19 December 2019

Abstract: This contribution describes the effect of the quality of the catalyst coating of cathodes for
wastewater treatment by microbial fuel cells (MFC). The increase in coating quality led to a strong
increase in MFC performance in terms of peak power density and long-term stability. This more
uniform coating was realized by an airbrush coating method for applying a self-developed polymeric
solution containing different catalysts (MnO2 , MoS2 , Co3 O4 ). In addition to the possible automation of
the presented coating, this method did not require a calcination step. A cathode coated with catalysts,
for instance, MnO2 /MoS2 (weight ratio 2:1), by airbrush method reached a peak and long-term power
density of 320 and 200–240 mW/m2 , respectively, in a two-chamber MFC. The long-term performance
was approximately three times higher than a cathode with the same catalyst system but coated with
the former paintbrush method on a smaller cathode surface area. This extraordinary increase in MFC
performance confirmed the high impact of catalyst coating quality, which could be stronger than
variations in catalyst concentration and composition, as well as in cathode surface area.

Keywords: microbial fuel cell; wastewater treatment; oxygen reduction reaction; municipal
wastewater; MnO2 ; MoS2 ; Co3 O4 ; spray method

1. Introduction
The permanent rise in standard of living correlates with an increasing energy demand [1].
Particularly in view of global warming and the intended abandonment of fossil and nuclear fuels,
processes must be developed that provide environmentally friendly energy [2]. Another challenge
is water scarcity, causing 1.2 billion people to suffer from a lack of water and leaving another 1.6
billion people without any access to hygienically safe drinking water [3]. Microbial fuel cells (MFC)
are a promising technology that can contribute to a decentralized solution for these future challenges,
enabling expansion and assurance of water availability and energy supply [4,5].
Microbial fuel cells (MFCs) combine electrical energy generation with simultaneous wastewater
treatment by microorganisms [5]. Anodic exoelectrogenic bacteria oxidize organic compounds from
wastewater [6,7]. In combination with a coupled reduction reaction, a positive potential difference,
and therefore a net power, is achieved [8]. Oxygen is usually used as final electron acceptor in the
coupled oxygen reduction reaction (ORR) at the cathode, as oxygen is available in ambient air [8]. The
schematic structure of the investigated MFC is shown in Figure 1.

Processes 2020, 8, 11; doi:10.3390/pr8010011 115 www.mdpi.com/journal/processes


Processes 2020, 8, 11

Figure 1. Schematic drawing of the investigated microbial fuel cell (MFC): a planar anode overgrown
with bacteria (dark gray), an oxygen-reducing cathode (light gray), and cation exchange membrane
(white). The desired reactions and subsequent products are sketched. An electrical load (El. Load) is
placed in the outer electrical circuit to use the generated power.

The net power generation of MFCs is rather low, especially due to the performance-limiting
cathode [9–12]. To overcome this bottleneck, large electrode surfaces [13], improved cathode
manufacturing methods, and especially scalable methods are needed [5]. In the laboratory, cathodes
for MFCs are often prepared by brushing [9,14–18], dipping [19], rolling [20], pressing [21], doctor
blading [12], spin coating [22], or electro deposition [19,23]. Most of these methods are either not
applicable, or less so, for economical technical scale electrodes because they are time-consuming,
complex, or not adaptable. Several patents describe manufacturing methods for large-scale cathodes for
chemical fuel cells [24–26] or chlor-alkali electrolysis [27,28], but these methods are not yet adapted for
MFCs. For that reason, a cathode-coating method is presented in this study, which is also a promising
approach for automation, and for the use of large-scale cathodes for MFCs for the first time.
This coating is realized by an airbrush method for applying a self-developed polymeric solution
containing catalysts. The coating process was investigated with different catalysts, their different
mixing ratios, and different metal meshes.
Next to carbon, MnO2 , MoS2 (with varying mixing ratios), and Co3 O4 (with TiO2 as support
material) were investigated as additional catalysts for ORR. As a benchmark for ORR [8], one electrode
was coated with Pt as the catalyst. Unfortunately, Pt is very expensive and susceptible to catalyst
poisoning by S2 − or NH4 + [29], which are commonly present in municipal wastewater, and as a
consequence Pt is not suitable for wastewater treatment applications.
Several publications investigated MnO2 as an ORR catalyst for MFCs, including the influence of
its different modifications (alpha, beta, gamma), different mixing ratios, and varying manufacturing
methods. Modifications of MnO2 were studied by Zhang et al. in an MFC (glucose as feed). The
maximum power densities differ from 125 mW/m2 (alpha) to 172 mW/m2 (beta) and 88 mW/m2
(gamma) [30]. This approach was followed by Roche et al. with similar performances up to 161 mW/m2

116
Processes 2020, 8, 11

(synthetic wastewater) [31]. Newer studies focus on more complex MnO2 -based cathodes. Li et al.
studied a manganese oxide catalyst with a cryptomelane-type octahedral molecular sieve structure.
This catalyst was additionally doped with Co, Cu, and Ce. The highest power density was measured for
the Co-doped system with 180 mW/m2 (domestic wastewater, acetate) [16], followed by an optimized
system with a continuous flow MFC. The maximum power density was 201 mW/m2 (wastewater
inoculum; acetate as feed) [17]. Recent studies partially increased the performance of MnO2 even further.
Zhou et al. investigated the behavior of the combination of polyaniline and MnO2 nanocomposites
and reached power densities of 248 mW/m2 (anaerobic digester sludge inoculum) [32]. Furthermore,
nano-scaled systems based on low-cost MnO2 nanowires (on carbon support) achieved a maximum
of 180 mW/m2 (combination of domestic and artificial wastewater) [33]. Farahani et al. investigated
another MnOx -based cathode and found that nitrogen-doped carbon with MnOx revealed a peak power
density of 467 mW/m2 (acetate as feed) [34]. However, the described cathode production process is
complex and is difficult to scale. A further study with modified MnO2 and induced carbon nanotubes
(CNT) revealed a drastically lower power density of approximately 12 mW/m2 (calculated with the
given volumetric power density of 216 mW/m3 ) [35].
Jiang et al. investigated the performance of MnO2 and its combination with MoS2 (real wastewater).
The best combination of MnO2 and MoS2 revealed peak power densities of up to 165 mW/m2 [9,14].
Another investigated composition within this study was MoS2 , which is sparsely described in
literature. The combination of MoS2 /C is considerably less investigated than MnOx catalysts. Hao et
al. evaluated N-doped MoS2 /C catalysts and reached high power densities (815 mW/m2 ) in a small
reactor (28 mL) with artificial wastewater (at 30 ◦ C) due to the beneficial catalyst and the optimized
process conditions [20].
In comparison, Co3 O4 /C cathodes are more often studied. Ge et al. revealed power densities
of up to 1500 mW/m2 in an MFC (artificial wastewater). This was confirmed by Xia et al. with
power densities of up to 1540 mW/m2 . The performance of real wastewater with high chemical
oxygen demand (COD; 56,500 mg/L) was investigated by Kumar et al. They reached power densities
up to 503 mW/m2 [36]. A following study with flower-like Co3 O4 and lower COD loads (digester
sludge and acetate) showed lower maximum power densities of 248 mW/m2 [15]. Our working group
already demonstrated the long-term stability of MnO2 [14] , MoS2 [9], and Co3 O4 [37] electrodes. It
was shown that MnO2 , MoS2 , and Co3 O4 have decent ORR properties and fulfill the requirements of
MFC cathodes—they are inexpensive, long-term stable, environmentally friendly [38,39], reduce the
overvoltage, and are immune to catalyst poising. Unfortunately, these cathodes were produced by a
time-consuming manual paintbrush method. Therefore, a novel spray coating method newly adapted
to MFC applications by a dedicated suspension is presented, which is also adaptable to automation
processes and scalable electrodes.

2. Materials and Methods

2.1. Experimantal Setup


To compare and characterize all coated electrodes, a laboratory MFC (self-made) was installed.
The test facility consists of eight identical test cells, and an overview is given in Figure 2. Purchased
graphite compound anodes were used, made of 86% graphite and 14% olefinic polymer binder
(Eisenhuth GmbH, Osterode am Harz, Germany). This material allowed the microorganisms to grow
on the surface and was also scalable. A low-cost membrane type FKS-PET 130 (Fumatech GmbH,
Bietigheim-Bissingen, Germany) was used to separate the anaerobic anode chamber and the aerobic
cathode chamber. For more information regarding the cell and anode design, refer to [14].

117
Processes 2020, 8, 11

Figure 2. Presentation of the used laboratory system, whereby all measurements were done in the
white- marked MFC for comparability.

Each MFC was fed with wastewater via a pump (Pumpen und Anlagentechnik, Lutherstadt
Wittenberg, Germany, model: 18ZP-VA 0,68-D30-118 FU) and all cathodes were connected to one water
basin aerated with ambient air. For all characterizations, the same test cell (Figure 2, circle) was used to
increase the measurements’ comparability.
During the operation, the temperature, pH values, electrical potentials, cell voltage, and current
were monitored through a LabView system (National instruments, Austin, TX, USA). All experiments
were carried out at a temperature of approximately 16.5 ◦ C (room temperature in the laboratory).
The pH values were kept at approximately 8 (anolyte) and approximately 10 (catholyte) by adding a
sodium carbonate solution (20 wt %) (Carl Roth GmbH & Co. KG, Karlsruhe, Germany). A constant
current source was used to enable a defined microbial growth at continuously controlled maximum
power point conditions. For detailed information about the constant current source, refer to [9,14].
The inoculation process took 2 weeks and was performed with real wastewater taken after the
primary clarifier of the sewage treatment plant (STP) Goslar (Eurawasser Betriebsführungsgesellschaft
mbH, Goslar, Germany). After filling the anode tank with about 10 L of wastewater and turning on the
pumps, the cells were kept in open circuit for 5 days. Afterwards, each cell was connected to its own
constant current source for individual cell operation and regulation.
To evaluate the electrodes under real conditions, the effluent of the primary clarifier from the
same STP Goslar with an initial COD of 150 to 200 mg/L was treated. Despite weekly wastewater
exchanges, 10 mL of a solution of 200 g/L glucose (Carl Roth GmbH & Co. KG, Karlsruhe, Germany).
and 200 g/L NaAc (Carl Roth GmbH & Co. KG, Karlsruhe, Germany) was added daily to keep a
constant COD concentration level of approximately 200 mg/L, and to ensure the comparability of
results at different dates.

2.2. Catalysts and Carrier Materials for Cathode


Printex 6L carbon (Orion Engineered Carbons S.A., Houston, TX, USA) was used for the electrode
coating. MnO2 by EMD Millipore Corporation (article number: 805958) (Burlington, MA, USA) and
MoS2 by Metallpulver24 Corp. (article number: 22020) (Sakt Augustin, Germany) were used as
catalysts. Butanone served as solvent (Sigma-Aldrich, St. Louis, MO, USA, 443468). Due to the high
prices of Co3 O4 , carrier supported catalysts (Co3 O4 @ TiO2 ) were evaluated. For these Co3 O4 catalysts,
Co(NO3 )2 was obtained by Sigma-Aldrich (article number: 239267) (St. Louis, MO, USA) and anatase
TiO2 nano particles by Cofermin Chemical GmbH (Essen, Germany). Anatase TiO2 in a whiskers
morphology was produced according to reference [40]. Fuel cell grade Pt black by De Nora North
America ETEK Division (S990670; Painesville, OH, USA) was used for the Pt coating for benchmarking.
Three different types of metal mesh were used as carrier material for the coating. The physical
details of the investigated meshes are given in Table 1. Hereinafter, coarse stainless-steel mesh is

118
Processes 2020, 8, 11

abbreviated as “cSS”, fine stainless steel mesh as “fSS”, and nickel net with a very fine mesh width
as “vfNi”.

Table 1. Physical data of the investigated wire meshes for electrode coating.

Surface Area
Mesh Type Mesh Size (mm2 ) Wire Gauge (mm) Manufacturer
(mm2 cm−2 )
Coarse stainless steel
1.8 × 1.8 0.32 94.8 Spörl KG, Germany
(cSS)
Fine stainless steel
1.0 × 1.0 0.22 116.3 Spörl KG, Germany
(fSS)
Very fine nickel Harver & Böcker,
0.106 × 0.118 0.063 627.2
(vfNi) Germany

2.3. Electrode Spray Coating Process


The coating process for the cathodes is shown in Figure 3. At first, a coating solution has to be
prepared, consisting of a basic polymer solution, catalyst, and additional butanone. The basic polymer
binder solution was prepared by dissolving three table tennis balls (ink-free with no commercial logos)
in 150 mL butanone to obtain a celluloid solution. Due to the strict standards, table tennis balls form
high-quality educts at low prices and fit perfectly to the low-cost approach for a low-cost flexible
binder. Then, the basic polymer solution (20 mL) was merged with additional butanone (100 mL). To
obtain a homogeneous dispersion, the solution was stirred, and the catalysts were added stepwise (4 g
carbon + 0.4 g additional catalyst: MnO2 and/or MoS2 , Co3 O4 ). Then, the mixture was homogenized
for 15 min with a dispersing tool (Heidolph, Germany, model SilentCrusher at 13,500 rpm).

Figure 3. Flow diagram of the process steps for electrode preparation using the spray method.

A spray table with heating option and the degreased (degreaser: isopropanol) metal mesh were
heated and controlled at a very stable surface temperature of 110 ◦ C. The dispersion was manually
applied with an airbrush spray gun (Harder & Steenbeck, Norderstedt, Germany, model: Evolution
Solo) in cross-coat. Inert nitrogen (Linde Gases Division, Pullach, Germany) was used as carrier gas for
spraying (3.5 N grade and a pressure of 2 bar). The applicated mass was controlled by weighting the
cathode before, during, and after the coating process. Furthermore, the optimum catalyst loading was
evaluated and finally fixed to 0.4 mg/cm2 by preliminary tests. The coating process was completed
when the desired catalyst loading was reached. No calcination step was required after coating; therefore,
the coating process was timesaving and avoided further changes to the catalyst.

2.4. Measuring Procedure and Analytical Methods


A Luggin capillary is usually used to investigate the electrode potential compared with a reference
electrode. In order to not only measure the potential at one geometric point, a membrane extension
operated as a salt bridge between the cathode and the reference electrode. Therefore, the potentials of
the electrodes were measured in an integrative way over the entire active area [41]. The membrane
extension was kept moist by an applied carbon fleece in order to stabilize the ionic conductivity and
the measurement (Figure 4). A reversible hydrogen electrode (RHE, Gaskatel GmbH, Kassel, Germany)

119
Processes 2020, 8, 11

served as a reference electrode, which was operated in a buffer solution at pH 7 (Libutec GmbH,
Langenfeld, Germany; type: buffer solution pH 7).

Figure 4. Sketch of the measurement setup and the salt bridge-like membrane extension for potential
determination. This modified version is based on [41].

The cathodes were characterized by measuring the electrode potentials and the cell voltages as
a function of the current. Furthermore, the power density characteristic curves were calculated and
normalized to the geometric area. The power was adjusted between 0 mA to short circuit current in
1 mA steps. After each step, the stationary state was awaited (average duration: 5 min).
To analyze the chemical composition, an X-ray powder diffraction (XRD) analysis was conducted
with a D/max-2200PC-X-ray (Rigaku Corp., The Woodlands, TX, USA) diffractometer (40 kV, 20 mA)
using CuKα radiation (0.15404 nm) and a scan range between 10◦ to 80◦ , with 10◦ /min.
Physical properties were analyzed by the Brunauer-Emmet-Teller (BET) method with a NOVA2000e
device (Micromeritics GmbH, Unterschleissheim, Germany) with a prearranged outgassing step at
200 ◦ C at a vacuum pressure of 6 mmHg of the samples. Selected cathodes were examined with a
scanning electron microscope (SEM) at the Institute of Mechanical Process Engineering at Clausthal
University of Technology (Carl Zeiss Microscopy GmbH, Jena, Germany, model: DSM 982 Gemini.M).

3. Results and Discussion

3.1. Characterization of Catalysts and Supports


MnO2 and MoS2 have already been evaluated using XRD and BET; reference is made to [9]. It
should be noted that gamma MnO2 was intentionally used, which is significantly cheaper than other
modifications. Crystalline phases of the investigated TiO2 supports, and its combination with Co3 O4 ,
are shown in Figure 5. The whiskers type of TiO2 was abbreviated as TiO2 -W, whereas TiO2 -A is the
abbreviation for the anatase modification. The XRD pattern of the supports showed peaks at 2θ =
25.2◦ , 37.8◦ , 48.0◦ , and 55.0◦ , confirming both modifications were anatase phase TiO2 . The analysis of
the catalyst-loaded TiO2 supports revealed a successful loading, as the peaks of TiO2 were still present,
next to the peaks of Co3 O4 at 2θ = 36.9◦ and 65.2◦ .

120
Processes 2020, 8, 11

Figure 5. X-ray powder diffraction (XRD) pattern of TiO2 supports (a) and in combination with Co3 O4
catalyst (b).

The BET surface areas of the TiO2 -supported and prepared Co3 O4 /TiO2 catalysts are presented in
Table 2. The comparison between the investigated TiO2 modifications showed that the TiO2 -W had a
higher specific surface area (85.9 to 70 m2 /g), and larger pore volume and size. After catalyst loading,
the surface area decreased, but the loaded TiO2 -W support still had a greater surface area.

Table 2. Catalyst content and physical properties of TiO2 carrier and its combination with Co3 O4 .
TiO2 -A: anatase modification of TiO2 , TiO2 -W: whiskers type of TiO2 .

Sample SBET (m2 g−1 ) VP (cm3 g−1 ) d (nm) Surface Content (weight %)
TiO2 -A 70.0 0.241 13.5 -
TiO2 -W 85.9 0.400 18.3 -
Co3 O4 /TiO2 -A 60.9 0.231 14.1 13.0
Co3 O4 /TiO2 -W 77.8 0.323 15.9 4.2

Figure 6 illustrates the Co 2p3/2 spectra of the prepared catalysts. All patterns revealed two peaks
at 786.2 eV and 779.4 eV, whereas the small peak can be attributed to Co2+ species, and the bigger peak
at 779.4 eV to Co3+ . Using this reasoning, Co3 O4 was the predominant phase for all samples.

Figure 6. Co 2p3/2 spectra for the supported catalysts. The binding energy for Co3 O4 (779.4 eV) and
Co2 O3 (786.2 eV) is shown.

121
Processes 2020, 8, 11

3.2. Performance of Coated Electrodes

3.2.1. Comparison of MnO2 , MoS2 , and Co3 O4 -Based Cathodes


The presented spraying method was applied to a variety of catalysts (MnO2 , MoS2 , Co3 O4 , and Pt)
and different metal meshes. A summary of the fabricated and tested combinations is given in Table 3.

Table 3. Overview of the produced electrodes by the novel spray method on different meshes.

Coarse Stainless Fine Stainless Very Fine Ni


Added Catalyst
Steel Mesh Steel Mesh Mesh
Pt x
MnO2 x x x
MoS2 x x x
MnO2 + MoS2 with 1:1 x x x
MnO2 + MoS2 with 1:2 x x x
MnO2 + MoS2 with 2:1 x x x
Co3 O4 at TiO2 -A x
Co3 O4 at TiO2 -W x

In order to achieve a concise comparison of the investigated cathodes, all electrode configurations
were characterized as described. Figure 7 shows an example of a power density curve, including the
anode and cathode potential curve, as well as the current–voltage curve.


9ROWDJH
$QRGHSRWHQWLDO 
 &DWKRGHSRWHQWLDO
3RZHUGHQVLW\
 
3RZHUGHQVLW\>P:P@

9ROWDJH>P9@






 




        
&XUUHQW>P$@
Figure 7. Characteristic curves of an microbial fuel cell (MFC) with a spray-coated cathode based on
MnO2 -MoS2 (1:1) on fSS wire mesh. This example also underlines the cathodic bottleneck of the system,
determined by the stronger decreasing cathodic potential in comparison to the anodic potential.

The maximum performance of each electrode is given in Figure 8. For the sake of clarity, electrodes
with different meshes, but identical catalysts, are presented in clusters. A cathode based on the
reference catalyst Pt (fuel cell grade Pt at cSS substrate) was produced and used for comparison in
Figure 8 (left).

122
Processes 2020, 8, 11

Figure 8. Comparison of maximum power densities of different cathodes. All electrodes were
manufactured with a catalyst load of 0.4 mg/cm2 .

The comparison clearly indicates that all catalysts benefit from an increasing surface area of the
metal meshes. The performance increases from coarse (cSS) to fine stainless-steel mesh (fSS) and is
outperformed by the cathodes on very fine nickel mesh (vfNi). However, the 5.4-times increase of
surface area of vfNi mesh cathode compared to fSS mesh cathodes only led to a slight increase in MFC
performance for different catalyst systems, including the best one (MnO2 -MoS2 2:1). For the MnO2
catalyst, a significant increase of approximately 80 mW/m2 (30%) was found, but still at a lower level
than the best catalyst system. These results indicate an existing limit for the impact of electrode surface
area on the improvement of MFC performance.
Electrodes with CO3 O4 (on TiO2 support) and MoS2 showed the lowest power density (only
60–88 mW/m2 ). Both investigated Co3 O4 on TiO2 systems (in combination with carbon and polymer)
showed similar results, but the Co3 O4 /TiO2 -W cathode was slightly better (74.4 mW/m2 ) in comparison
to Co3 O4 /TiO2 -A (63.6 mW/m2 ), probably caused by the increased surface area.
The increasing surface area caused by the finer mesh width also had a positive effect on MoS2 -based
cathodes, but the performance was still low (60.2, 65.1, and 88.5 mW/m2 ). The oxygen reduction
capability of MnO2 was considerably higher. The increase from cSS to fSS enhanced the maximum
power density from 145.4 to 153.9 mW/m2 . A further improvement was given by the usage of the vfNi
mesh, showing power densities up to 240.9 mW/m2 . Sulphur compounds in wastewater are known
as catalyst poisons, and it was recently shown that a catalyst combination with MoS2 increases the
long-term stability of MFC cathode catalysts [9]. Therefore, different mixing ratios were investigated.

123
Processes 2020, 8, 11

The (weight %) mixing ratio of 1:1 (MnO2 -MoS2 ) led to a performance deterioration compared
to the pure MnO2 system, and the electrodes showed similar performances to the MoS2 cathodes on
all substrates. The mixing ratios of 1:2 and 2:1 performed differently, but they showed an improving
effect. In particular, the electrodes with a mixing ratio of 2:1 showed power densities of 198.3 mW/m2
(MnO2 -MoS2 2:1, cSS), 312.3 mW/m2 (MnO2 -MoS2 2:1, fSS), and 320.6 mW/m2 (MnO2 -MoS2 2:1, vfNi).
An even better performance was reached with the fSS and vfNi meshes than for the Pt-coated cathode.
Potential/current curves of selected cathodes are shown in Figure 9. The diagram clearly indicates the
most stable cathode potential curve for MnO2 /MoS2 (2:1). Furthermore, the cathode potential stabilized
in terms of current stability from the cSS mesh (grey curve) over the fSS mesh (blue curve) to the vfNi
mesh (green curve), with simultaneously increasing cathode surface area. All cathodes showed a
potential drop, especially at higher currents. This was probably caused by mass transport limitations.


0Q20R6F66
 0Q20R6I66
0Q20R6YI1L
 0Q2F66
&DWKRGHSRWHQWLDO>P9@

0Q2I66
 0Q2YI1L










      
&XUUHQW>P$@
Figure 9. Cathode potentials of selected spray-coated cathodes on different substrates.

Due to the high costs for Co3 O4 , carrier-supported catalysts (Co3 O4 at TiO2 ) were evaluated.
However, these showed low ORR activity and the achieved currents were below published levels.
Obviously, pure Co3 O4 is expected to show significantly better performance, but this is not suitable for
wastewater treatment by MFC due to the low cost-efficiency. In conclusion, the MnO2 -based cathodes
showed significantly better performances. Especially a combination with MoS2 could surpass the
performances related to scalable systems described in the literature to date.

3.2.2. Comparison to Paintbrush-Manufactured Cathodes


Novel spray coated cathodes were compared to cathodes previously produced by the paintbrush
method [9] by SEM. All cathodes had a catalyst ratio of 2:1 (MnO2 /MoS2 ). According to Figure 10, the
paintbrush process produced a tight fit between the interstitial spaces, whereby some were closed,
whereas other interstitial spaces remained empty. The coating was not uniform and showed a tendency
for detaching. The application consisted mainly of elongated particles attached to each other in a
porous structure, which were fixed in position by form closure.

124
Processes 2020, 8, 11

Figure 10. Scanning electron microscope (SEM) image of a cathode produced by paintbrush method
with a catalyst ratio of 2:1 (MnO2 /MoS2 ) in different scales.

In contrast, the cathodes produced by the spray coating method showed a more regular coating
on the surface of the wire mesh. The finer the wire mesh, the more uniform the distribution. The SEM
image of coated vfNi substrate is shown as an example in Figure 11.

Figure 11. SEM image of a spray coated cathode with a catalyst ratio of 2:1 (MnO2 /MoS2 ) in
different scales.

125
Processes 2020, 8, 11

Thus, it was shown that the presented spray method enabled uniform and reproducible catalytic
coatings. The developed spray coating method seems to be a promising process for the production of
high-performance electrodes and also for upscaling. The method also enabled a suitable comparison
between different catalysts through homogenous coatings. No additional calcination steps were
necessary, which could modify the catalyst structure.

3.2.3. Long-Term Performance of Selected Electrodes


The long-term performance of the best performing spray-coated cathode (MnO2 -MoS2 2:1 vfNi) in
comparison to a paintbrush-coated cathode (according to [9]) is given in Figure 12. It was noted that the
paintbrush method was applied on a cSS net, as the paintbrush method would cause clogging on finer
mesh width. Although a drop in power density was noticeable at the beginning of the measurement, a
relatively constant power density of 150 mW/m2 was reached from the fourth day onwards. From
day 6, the power density rose, reaching 200–240 mW/m2 from day 10 onwards. The long-term
performance of paintbrush-coated electrode was about 60 to 100 mW/m2 . The comparison of the
long-term performance between both electrodes by different coating methods revealed a performance
improvement by factor 2 to 3, or in terms of power density, by 100 to 180 mW/m2 for the spray-coated
cathode. The comparison between the improvement of long-term power density and the improvement
of maximum power density showed that the improvement in long-term performance was higher than
that in maximum performance. Therefore, the spray-coating method was identified as enabling the
production of high-performing and long-term stable cathodes for MFC application.

7LPH>G@
       

 VSUD\FRDWHG0Q20R6ZLWKRQYI1L

 SDLQWEUXVKFRDWHG0Q20R6ZLWKRQF66


3RZHUGHQVLW\>P:P@
















       
7LPH>K@
Figure 12. Long-term performance of the spray-coated MnO2 -MoS2 (2:1) vfNi cathode.

4. Conclusions
To scale-up MFCs to the dimensions of technical-scale wastewater treatment plants, it is necessary
to manufacture scalable electrodes. Therefore, the composition of a universal sprayable suspension
was developed within this study, which was used for a simplified airbrush spray-coating method.
The performance was evaluated regarding different catalysts (Co3 O4 , MnO2 , MoS2 , and selected

126
Processes 2020, 8, 11

combinations) on different carrier meshes and materials. The spray-coating method facilitated a
homogeneous coating and stood out with an increased cathode performance. This could be enhanced
further by suitable substrates with a finer mesh width and a higher surface area. The most promising
cathode catalyst composition of the tested systems was the combination of MnO2 and MoS2 mixture
(2:1) with a maximum and long-term power density of 320 and 200–240 mW/m2 , respectively. The
coated electrodes also demonstrated long-term stability. This contribution confirmed the effect of the
quality of the catalyst coating of cathodes for wastewater treatment by microbial fuel cells (MFC).
A promising and more rapid manufacturing method for better catalyst comparisons and large-scale
applications was identified. This process could also be the basis for an automated coating. Moreover,
this work demonstrated that not only the catalytic components and their composition influenced
electrocatalytic properties. The method of preparation and the carrier mesh also exerted a large effect
on the electrode performance.

Author Contributions: Conceptualization, methodology, validation, data curation, visualization, writing—


original draft preparation, T.M. and D.H.; investigation, T.M., D.H., and B.J.; writing—review and editing, all
authors; supervision, project administration, funding acquisition, M.S. and U.K. All authors have read and agreed
to the published version of the manuscript.
Funding: This research was funded by the Federal Ministry of Education and Research (Bundesministerium für
Bildung und Forschung), BMBF, Germany, grant number WTER0219813.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Muddemann, T.; Haupt, D.; Sievers, M.; Kunz, U. Electrochemical Reactors for Wastewater Treatment.
ChemBioEng Rev. 2019, 6, 142–156. [CrossRef]
2. United Nations. The Millennium Development Goals Report; United Nations: New York, NY, USA, 2006.
3. UN Water and FAO. Coping with Water Scarcity—Challenge of the Twenty-First Century; UN Water and FAO:
Rome, Italy, 2007.
4. Logan, B.E. Microbial Fuel Cells; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008; ISBN 978-0-470-23948-3.
5. Hernández-Fernández, F.J.; Pérez de los Ríos, A.; Salar-García, M.J.; Ortiz-Martínez, V.M.; Lozano-Blanco, L.J.;
Godínez, C.; Tomás-Alonso, F.; Quesada-Medina, J. Recent progress and perspectives in microbial fuel cells
for bioenergy generation and wastewater treatment. Fuel Process. Technol. 2015, 138, 284–297. [CrossRef]
6. Zhang, X.; He, W.; Ren, L.; Stager, J.; Evans, P.J.; Logan, B.E. COD removal characteristics in air-cathode
microbial fuel cells. Bioresour. Technol. 2015, 176, 23–31. [CrossRef] [PubMed]
7. Malvankar, N.S.; Lovley, D.R. Microbial nanowires: A new paradigm for biological electron transfer and
bioelectronics. ChemSusChem 2012, 5, 1039–1046. [CrossRef] [PubMed]
8. Yuan, H.; Hou, Y.; Abu-Reesh, I.M.; Chen, J.; He, Z. Oxygen reduction reaction catalysts used in microbial
fuel cells for energy-efficient wastewater treatment: A review. Mater. Horiz. 2016, 3, 382–401. [CrossRef]
9. Jiang, B.; Muddemann, T.; Kunz, U.; Silva e Silva, L.G.; Bormann, H.; Niedermeiser, M.; Haupt, D.; Schläfer, O.;
Sievers, M. Graphite/MnO2 and MoS2 Composites Used as Catalysts in the Oxygen Reduction Cathode of
Microbial Fuel Cells. J. Electrochem. Soc. 2017, 164, E519–E524. [CrossRef]
10. Rahimnejad, M.; Adhami, A.; Darvari, S.; Zirepour, A.; Oh, S.-E. Microbial fuel cell as new technology for
bioelectricity generation: A review. Alex. Eng. J. 2015, 54, 745–756. [CrossRef]
11. Rismani-Yazdi, H.; Carver, S.M.; Christy, A.D.; Tuovinen, O.H. Cathodic limitations in microbial fuel cells:
An overview. J. Power Sources 2008, 180, 683–694. [CrossRef]
12. Santoro, C.; Arbizzani, C.; Erable, B.; Ieropoulos, I. Microbial fuel cells: From fundamentals to applications.
A review. J. Power Sources 2017, 356, 225–244. [CrossRef]
13. Muddemann, T.; Haupt, D.R.; Gomes Silva e Silva, L.; Jiang, B.; Kunz, U.; Bormann, H.; Niedermeiser, M.;
Schlaefer, O.; Sievers, M. Integration of Upscaled Microbial Fuel Cells in Real Municipal Sewage Plants. ECS
Trans. 2017, 77, 1053–1077. [CrossRef]
14. Jiang, B.; Muddemann, T.; Kunz, U.; Bormann, H.; Niedermeiser, M.; Haupt, D.; Schläfer, O.; Sievers, M.
Evaluation of Microbial Fuel Cells with Graphite Plus MnO2 and MoS2 Paints as Oxygen Reduction Cathode
Catalyst. J. Electrochem. Soc. 2016, 164, H3083–H3090. [CrossRef]

127
Processes 2020, 8, 11

15. Kumar, R.; Singh, L.; Zularisam, A.W. Enhanced oxygen reduction reaction in air-cathode microbial fuel
cells using flower-like Co3 O4 as an efficient cathode catalyst. Int. J. Hydrog. Energy 2017, 42, 19287–19295.
[CrossRef]
16. Li, X.; Hu, B.; Suib, S.; Lei, Y.; Li, B. Manganese dioxide as a new cathode catalyst in microbial fuel cells. J.
Power Sources 2010, 195, 2586–2591. [CrossRef]
17. Li, X.; Hu, B.; Suib, S.; Lei, Y.; Li, B. Electricity generation in continuous flow microbial fuel cells (MFCs) with
manganese dioxide (MnO2 ) cathodes. Biochem. Eng. J. 2011, 54, 10–15. [CrossRef]
18. Nandy, A.; Sharma, M.; Venkatesan, S.; Taylor, N.; Gieg, L.; Thangadurai, V. Comparative Evaluation of
Coated and Non-Coated Carbon Electrodes in a Microbial Fuel Cell for Treatment of Municipal Sludge.
Energies 2019, 12, 1034. [CrossRef]
19. Zhang, Y.; Liu, L.; van der Bruggen, B.; Yang, F. Nanocarbon based composite electrodes and their application
in microbial fuel cells. J. Mater. Chem. A 2017, 5, 12673–12698. [CrossRef]
20. Hao, L.; Yu, J.; Xu, X.; Yang, L.; Xing, Z.; Dai, Y.; Sun, Y.; Zou, J. Nitrogen-doped MoS2 /carbon as highly
oxygen-permeable and stable catalysts for oxygen reduction reaction in microbial fuel cells. J. Power Sources
2017, 339, 68–79. [CrossRef]
21. Kodali, M.; Santoro, C.; Serov, A.; Kabir, S.; Artyushkova, K.; Matanovic, I.; Atanassov, P. Air Breathing
Cathodes for Microbial Fuel Cell using Mn-, Fe-, Co- and Ni-containing Platinum Group Metal-free Catalysts.
Electrochim. Acta 2017, 231, 115–124. [CrossRef]
22. Tsai, H.-Y.; Hsu, W.-H.; Liao, Y.-J. Effect of Electrode Coating with Graphene Suspension on Power Generation
of Microbial Fuel Cells. Coatings 2018, 8, 243. [CrossRef]
23. Satar, I.; Daud, W.R.W.; Kim, B.H.; Somalu, M.R.; Ghasemi, M.; Bakar, M.H.A.; Jafary, T.; Timmiati, S.N.
Performance of titanium–nickel (Ti/Ni) and graphite felt-nickel (GF/Ni) electrodeposited by Ni as alternative
cathodes for microbial fuel cells. J. Taiwan Inst. Chem. Eng. 2018, 89, 67–76. [CrossRef]
24. Hideki, M. Manufacturing Equipment and Manufacturing Method of Membrane Electrode Assembly.
US20100051181A1, 4 March 2010.
25. Kwon, N.H.; Hwang, I.C.; Ahn, B.K.; Lim, T.W. Method and Apparatus for Preparing Catalyst Slurry for
Fuel Cells. US20100086450A1, 8 April 2010.
26. Tochigi, T.T.; Tochigi, S.T.; Tochigi, K.Y. Verfahren zur Herstellung einer Elektrodenschicht für eine
Brennstoffzelle. DE1020006046373A1, 5 April 2007.
27. Bulan, A. Sauerstoffverzehrelektrode und Verfahren zu ihrer Herstellung. DE102010024053A1, 22 December 2011.
28. Kintrup, J.; Bulan, A.; Hammarberg, E.; Sepeur, S.; Frenzer, G.; Gross, F. Beschichtungsmischung Herstellung
Covestro. US2017067172A1, 9 March 2017.
29. Yang, W.; Li, J.; Lan, L.; Fu, Q.; Zhang, L.; Zhu, X.; Liao, Q. Poison tolerance of non-precious catalyst towards
oxygen reduction reaction. Int. J. Hydrog. Energy 2018, 43, 8474–8479. [CrossRef]
30. Zhang, L.; Liu, C.; Zhuang, L.; Li, W.; Zhou, S.; Zhang, J. Manganese dioxide as an alternative cathodic
catalyst to platinum in microbial fuel cells. Biosens. Bioelectron. 2009, 24, 2825–2829. [CrossRef] [PubMed]
31. Roche, I.; Katuri, K.; Scott, K. A microbial fuel cell using manganese oxide oxygen reduction catalysts. J.
Appl. Electrochem. 2010, 40, 13–21. [CrossRef]
32. Zhou, X.; Xu, Y.; Mei, X.; Du, N.; Jv, R.; Hu, Z.; Chen, S. Polyaniline/β-MnO2 nanocomposites as cathode
electrocatalyst for oxygen reduction reaction in microbial fuel cells. Chemosphere 2018, 198, 482–491. [CrossRef]
33. Majidi, M.R.; Shahbazi Farahani, F.; Hosseini, M.; Ahadzadeh, I. Low-cost nanowired α-MnO2 /C as an ORR
catalyst in air-cathode microbial fuel cell. Bioelectrochemistry 2018, 125, 38–45. [CrossRef]
34. Shahbazi Farahani, F.; Mecheri, B.; Reza Majidi, M.; Costa de Oliveira, M.A.; D’Epifanio, A.; Zurlo, F.;
Placidi, E.; Arciprete, F.; Licoccia, S. MnOx -based electrocatalysts for enhanced oxygen reduction in microbial
fuel cell air cathodes. J. Power Sources 2018, 390, 45–53. [CrossRef]
35. Woon, C.W.; Ong, H.R.; Chong, K.F.; Chan, K.M.; Khan, M.M.R. MnO2 /CNT as ORR Electrocatalyst in
Air-Cathode Microbial Fuel Cells. Procedia Chem. 2015, 16, 640–647. [CrossRef]
36. Kumar, R.; Singh, L.; Zularisam, A.W.; Hai, F.I. Potential of porous Co3 O4 nanorods as cathode catalyst for
oxygen reduction reaction in microbial fuel cells. Bioresour. Technol. 2016, 220, 537–542. [CrossRef]
37. Jiang, B. Optimization of Cathode Performance of Microbial Fuel Cells for Wastewater Treatment and
Electrochemical Power Evaluation. Ph.D. Thesis, Clausthal University of Technology, Clausthal-Zellerfeld,
Germany, 2018.

128
Processes 2020, 8, 11

38. Rossi, R.; Jones, D.; Myung, J.; Zikmund, E.; Yang, W.; Gallego, Y.A.; Pant, D.; Evans, P.J.; Page, M.A.;
Cropek, D.M.; et al. Evaluating a multi-panel air cathode through electrochemical and biotic tests. Water Res.
2018, 148, 51–59. [CrossRef]
39. Brown, R.K.; Harnisch, F.; Wirth, S.; Wahlandt, H.; Dockhorn, T.; Dichtl, N.; Schröder, U. Evaluating the
effects of scaling up on the performance of bioelectrochemical systems using a technical scale microbial
electrolysis cell. Bioresour. Technol. 2014, 163, 206–213. [CrossRef]
40. Chen, S.; Zhu, Y.; Li, W.; Liu, W.; Li, L.; Yang, Z.; Liu, C.; Yao, W.; Lu, X.; Feng, X. Synthesis, Features, and
Applications of Mesoporous Titania with TiO2 (B). Chin. J. Catal. 2010, 31, 605–614. [CrossRef]
41. He, W.; van Nguyen, T. Edge Effects on Reference Electrode Measurements in PEM Fuel Cells. J. Electrochem.
Soc. 2004, 151, A185. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

129
processes
Article
Preparation and Characterization of Porous
Ti/SnO2–Sb2O3/PbO2 Electrodes for the Removal of
Chloride Ions in Water
Kangdong Xu 1 , Jianghua Peng 2 , Pan Chen 1 , Wankai Gu 1 , Yunbai Luo 1 and Ping Yu 1, *
1 College of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China;
[email protected] (K.X.); [email protected] (P.C.); [email protected] (W.G.);
[email protected] (Y.L.)
2 MOE Key Laboratory of Thermo-Fluid Science and Engineering, Xi’an Jiaotong University, Xi’an 710049,
China; [email protected]
* Correspondence: [email protected]; Tel.: +86-027-6875-2511

Received: 14 September 2019; Accepted: 14 October 2019; Published: 18 October 2019

Abstract: Porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes for electrocatalytic oxidation of chloride ions
were studied by exploring the effects of different operating conditions, including pore size, initial
concentration, current density, initial pH, electrode plate spacing, and the number of cycles. In
addition, a physicochemical characterization and an electrochemical characterization of the porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes were performed. The results showed that Ti/SnO2 –Sb2 O3 /PbO2
electrodes with 150 μm pore size had the best removal effect on chloride ions with removal ratios
amounting up to 98.5% when the initial concentration was 10 g L−1 , the current density 125 mA cm−2 ,
the initial pH = 9, and the electrode plate spacing 0.5 cm. The results, moreover, showed that the
oxygen evolution potential of 150 μm porous Ti/SnO2 -Sb2 O3 /PbO2 electrodes was highest, which
minimized side reactions involving oxygen formation and which increased the removal effect of
chloride ions.

Keywords: electrocatalytic oxidation; chloride ions removal ratio; the porous electrode;
influencing factors

1. Introduction
In recent years, China has raised the level of environmental protection, which requires industrial
wastewaters not to be discharged from various enterprises. Therefore, in order to ensure the normal
operation of its own production, each enterprise must realize the recycling of water. At present, the
commonly used method is to properly dispose of the drainage and then replenish it into the industrial
water system. However, various ions are continuously enriched in the water when the water is reused,
which causes various adverse effects on the operation of equipment. The amount of scale cations (such
as calcium ions and magnesium ions, etc.) in water can be reduced by changing the pH of the water
body and by inducing flocculation sedimentation [1], but there is no effective method of reducing
anions (such as chloride ions) which have corrosive effects on equipment in water [2]. Therefore, there
is a need to find a way to quickly and easily reduce chloride ions in water.
At present, the methods for removing chloride ions include biological [3,4], reverse osmosis [5,6],
distillation, multi-effect evaporation [7], electrodialysis [8–10], and ions exchange methods. The
biological method involves high operating costs, while the removal effect does not work well. The
reverse osmosis method is burdened by high consumption of acid and alkali as well as energy
consumption. The distillation method and the multi-effect evaporation methods involve high energy
consumption. The electrodialysis method involves high operation voltages, and membranes are easily

Processes 2019, 7, 762; doi:10.3390/pr7100762 131 www.mdpi.com/journal/processes


Processes 2019, 7, 762

contaminated. The ion exchange method is expensive and tends to cause secondary pollution during
the elution process. The electrocatalytic oxidation technology is a new method for removing chloride
ions, which has the advantages of high efficiency in removing chloride ions, simple process, simple
operation, and low operating cost, and there are few related studies.
In this study, porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were studied and it was found that
these have a significant effect on electrocatalytic oxidation of removing chloride ions. Self-made
porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes with different pore sizes were used as anodes to construct
an electrolyzer. The high potential of the anode and unique catalytic oxidation characteristics
of the surface coating were used for removing chloride ions. Three different kinds of porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes were prepared and characterized systematically with regard to
morphology, crystal structure, and various electrochemical performances. NaCl solutions were used
for simulating chlorine-containing wastewaters for studying the effects of initial concentration, pore
size, current density, initial pH, electrode plate spacing, and the number of cycles. The results showed
that the porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were efficiently reducing the number of chloride ions
in water, thus providing a practical method of removing chloride ions from water.

2. Experimental

2.1. Materials and Reagents


The substrates used in this study were commercial samples of porous Ti (20 mm × 10 mm × 1 mm),
which had average pore sizes of 50 μm, 100 μm, and 150 μm. In regard to removing chloride ions,
porous Ti substrates with larger pore sizes do not meet the requirements for removing chloride ions
in terms of hardness, our studies were restricted to the three pore sizes mentioned above. In this
paper, all chemicals were analytical grade without any other impurities. SnCl4 ·5H2 O, SbCl3 , HCl,
Pb(NO3 )2 , Cu(NO3 )2 ·3H2 O, NaF, HNO3 , NaOH, H2 C2 O4 , NaCl, and other chemicals were obtained
from Shanghai Wo Kai Biotechnology Co., Ltd. (Shanghai, China). All the solutions used in these
experiments were prepared with deionized water. For the simulation of chloride ion contaminated
waste waters, NaCl solutions were used.

2.2. Electrode Preparation


First, at a temperature of 70 ◦ C, porous Ti substrates (50 μm, 100 μm, and 150 μm) were heated in
sodium hydroxide (20% m%) for 1 h to remove all traces of oil on the surface and were then washed
in deionized water. Thereafter, the porous Ti substrates were etched in oxalic acid (15% m%) at a
temperature of 85 ◦ C for 1 h to obtain a uniformly rough surface. Finally, the samples were washed in
deionized water and stored in deionized water [11,12].
Second, the coating solution, consisting of 1.20 g SnCl4 ·5H2 O and 0.20 g SbCl3 was dissolved
in 40 mL ethanol, and 1 mL of concentrated hydrochloric acid was added. The treated porous Ti
substrates were immersed in the coating solution for 5 min, and then dried at about 120 ◦ C for 15 min,
and thereafter calcined at 500 ◦ C for 20 min in a muffle furnace. All above processes were repeated ten
times and the electrodes were annealed at 500 ◦ C for 60 min in the last process [13,14]. Finally, the
porous Ti/SnO2 –Sb2 O3 electrode was prepared.
Third, the deposition solution consisted of 40 g Pb(NO3 )2, 15 g Cu(NO3 )2 ·3H2 O, 0.5 g NaF, and
0.1 M HNO3 . The porous Ti/SnO2 –Sb2 O3 electrodes were used as anodes and a pure titanium plate as a
cathode. The current density was 5 mA cm−2 and the PbO2 was deposited on the porous Ti/SnO2 –Sb2 O3
electrode at 65 ◦ C for 0.5 h under stirring conditions. Finally, the prepared porous Ti/SnO2 –Sb2 O3 /PbO2
electrodes were washed in deionized water [15–17].

2.3. Electrode Characterization


The surface morphologies of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were characterized by Field
Emission Scanning Electron Microscopy (Zeiss SIGMA, Carl Zeiss Corporation, Jena, Germany). The

132
Processes 2019, 7, 762

composition and the chemical state of the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode was determined
by an X-ray photoelectron spectrometer (ESCALAB250Xi, Thermo Fisher Scientific, Waltham, MA,
USA). The crystal structure of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were determined by an X-ray
Diffractometer (XRD–6100, Shimadzu Corporation, Kyoto, Japan) with Cu–Kα (λ = 0.154 nm) incident
radiation at a scanning rate of 2 min−1 in 2θ mode from 20 to 90.
The electrochemical characterization including linear sweep voltammetry (LSV) curves and
cyclic voltammetry of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were carried out using a CHI760E
electrochemical workstation (CH Instruments, Chenhua Co., Shanghai, China) with a conventional
three-electrode cell. The working electrode was the PbO2 electrode; the reference electrode was a
saturated calomel electrode (SCE) and a platinum electrode served as a counter electrode.

2.4. Electrocatalytic Oxidation of Chloride Ions


The porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes were used as anodes and a Ti electrode as the cathode.
The produced chlorine was absorbed in an absorption tank. After the reaction, AgNO3 solution
was used as a titrant and K2 CrO4 solution as a color developer. In order to obtain optimal removal
conditions, the effects of the following factors were considered for electrocatalytic oxidation of chloride
ions, including initial concentration (from 5 g L−1 to 25 g L−1 ), pore size (from 50 μm to 150 μm),
current densities (from 50 mA cm−2 to 125 mA cm−2 ), initial pH (from 3 to 11), and electrode plate
spacing (from 0.5 cm to 1 cm). The removal ratio of chloride ions was calculated as follows:

B1 − B2
Removal ratios of chloride ions = × 100%
B1

B1 is the concentration of the original chloride ions in the NaCl solution, and B2 the concentration
of the remaining chloride ions in the NaCl solution after the reaction.

3. Results and Discussion

3.1. Physicochemical Characterization

3.1.1. Scanning Electron Microscopy (SEM) Characterization


In Figure 1, the scanning electron microscopy (SEM) images of the three different pore sizes
(50 μm, 100 μm, and 150 μm) are shown. As shown in Figure 1a,d,g, the surfaces of the porous
Ti substrates are very rough. Figure 1b,e,h show that the porous Ti/SnO2 –Sb2 O3 electrodes still
have a lot of irregular pores, which indicates that the porous Ti/SnO2 –Sb2 O3 electrodes have large
specific surface areas. By zooming 500 and 4000 times, it is seen that the SnO2 –Sb2 O3 intermediate
layers are uniformly distributed and crack-free, which is beneficial for prolonging the service life
of the electrode. Figure 1c,f,i show that the porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes still have many
irregular pores. Although some of the pores became during electrodeposition, large specific surface
areas were still retained compared to the planar electrodes. By zooming 500 and 4000 times, it is
revealed that the PbO2 grains are also very compact and evenly distributed. In short, the porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes have large surface areas, which can provide many active sites for
electrochemical oxidation [18].

133
Processes 2019, 7, 762

Figure 1. Scanning electron microscopy (SEM) images: (a) 50 μm porous Ti substrate; (b) 50 μm
porous Ti/SnO2 –Sb2 O3 electrode; (c) 50 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode; (d) 100 μm porous
Ti substrate; (e) 100 μm porous Ti/SnO2 –Sb2 O3 electrode; (f) 100 μm porous Ti/SnO2 –Sb2 O3 /PbO2
electrode; (g) 150 μm porous Ti substrate; (h) 150 μm porous Ti/SnO2 –Sb2 O3 electrode; (i) 150 μm
porous Ti/SnO2 –Sb2 O3 /PbO2 electrode.

3.1.2. X-ray Photoelectron Spectrometer (XPS) Characterization


In order to investigate the chemical state of each element in the porous Ti/SnO2 –Sb2 O3 /PbO2
electrodes, the electrodes were analyzed by XPS. Since the XPS spectra of the three different-pore-size
electrodes are the same, only the 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode is analyzed here.
Figure 2a is the full spectrum of a porous Ti/SnO2 –Sb2 O3 /PbO2 electrode. It can be seen that there are
mainly Ti, Sn, Sb, O, Pb, and C peaks in the whole electrode. Figure 2b shows the Sn 3d spectrum with
two characteristic peaks at 487.1 eV and 495.5 eV. Figure 2c shows the Sb 3d and O 1s spectra with
characteristic peaks at 540.3 eV and 531.6 eV. Figure 2d shows the Pb 4f spectrum with characteristic
peaks at 138.8 eV and 143.7 eV. The composition and chemical state of the porous Ti/SnO2 –Sb2 O3 /PbO2
electrode could be determined by the XPS characterization results, and it was inferred that porous
Ti/SnO2 –Sb2 O3 /PbO2 electrode was successfully prepared.

134
Processes 2019, 7, 762

Figure 2. (a) XPS spectrum of the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode; (b) XPS spectrum of Sn 3d;
(c) XPS spectrum of O 1s and Sb 3d; (d) XPS spectrum of Pb 4f.

3.1.3. X-ray Diffraction (XRD) Characterization


To further verify the results, the XRD patterns of the electrode coatings prepared on the three
different-pore-size Ti substrates are shown in Figure 3. Since the XRD images of the three pore sizes
electrodes are the same, only the 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes are analyzed here.
Figure 3 shows the β–PbO2 diffraction peaks, which are (110), (101), (200), (211), (220), (310), (301),
(321), (312), and (411). At the same time, Figure 3 shows the weak α–PbO2 diffraction peaks, which is
(041). In addition, there are no diffraction peaks of Ti, SnO2 , and Sb2 O3 , which proves that the PbO2
coating had completely covered the porous Ti/SnO2 –Sb2 O3 electrodes.

135
Processes 2019, 7, 762

Figure 3. The X-ray diffraction (XRD) pattern of 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode.

3.2. Electrochemical Characterization

3.2.1. Linear Sweep Voltammetry (LSV) Curves


Figure 4 shows the linear sweep voltammetry (LSV) curves of the three different kinds of porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes as obtained in a 0.5 mol L−1 H2 SO4 solution at a scan rate of 5 mV s−1 .
The result shows that the oxygen evolution potential is 2.02 V, 1.99 V, and 1.96 V at the 150 μm,
100 μm, and 50 μm porous electrodes, respectively. High oxygen evolution potentials indicate that side
reactions involving the formation of oxidized species are not very likely to occur [19]. 150 μm porous
Ti/SnO2 –Sb2 O3 /PbO2 electrodes therefore have the best removal effect on chloride ions.

Figure 4. Linear sweep voltammetry (LSV) curves of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes in
0.5 mol L−1 H2 SO4 solution at a scan rate of 5 mV s−1 .

136
Processes 2019, 7, 762

3.2.2. Cyclic Voltammetry


Figure 5 shows the cyclic voltammetry of the three different-pore-size Ti/SnO2 –Sb2 O3 /PbO2
electrodes in 0.5 mol L−1 H2 SO4 solution at a scan rate of 5 mV s−1 . The results show that the redox
peak of the 50 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode is highest, while it is lowest at the 150 μm
porous Ti/SnO2 –Sb2 O3 /PbO2 electrode. The low oxygen evolution potential of the 50 μm porous
Ti/SnO2 –Sb2 O3 /PbO2 electrode indicates that it has a high areal density of oxygen evolution sites. The
high oxygen evolution potential of the 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode, on the other
hand, indicates a low areal density of oxygen evolution sites, which is beneficial for the electrocatalytic
oxidation of chloride ions [20].

Figure 5. Cyclic voltammetry curve of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes in 0.5 mol L−1 H2 SO4
solution at a scan rate of 5 mV s−1 .

3.2.3. Linear Sweep Voltammetry (LSV) Curves in Different pH Solutions


Figure 6 shows the LSV curves of a 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode in different
pH solutions at a scan rate of 5 mV s−1 . Solutions with pH values between pH = 3 and pH = 5 were
prepared from buffer solution of citric acid and sodium citrate, and solutions with pH = 7 from buffer
solution of potassium monohydrogen phosphate and potassium dihydrogen phosphate. Solutions with
pH values between pH = 9 and pH = 11, finally, were composed from sodium carbonate and sodium
bicarbonate buffer solution. It can be seen from the Figure 6 that under the same voltage conditions,
the current density is smallest at pH = 9, which indicates that oxygen evolution is least likely to occur
under these conditions and that therefore 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes will exhibit
an optimum chloride ion removal effect at pH = 9.

137
Processes 2019, 7, 762

Figure 6. Linear sweep voltammetry (LSV) curves in different pH solutions of the 150 μm porous
Ti/SnO2 –Sb2 O3 /PbO2 electrode in 0.5 mol L−1 H2 SO4 solution at a scan rate of 5 mV s−1 .

3.3. Electrocatalytic Oxidation of Chloride Ions

3.3.1. Effect of Pore Size


With the electrolysis time fixed at 4 h, the electrode plate spacing at 1.0 cm, the NaCl concentration
at 10 g L−1 , and the current density at 100 mA cm−2 , the chloride ion removal ratios of the electrodes
were determined to explore the optimal electrode pore size among three different kinds of porous
electrodes in NaCl solution. As shown in Figure 7a, the optimal pore size of porous Ti/SnO2 -Sb2 O3 /PbO2
electrode is 150 μm after 4 h, and the removal ratio of chloride ions is 92.8%, which confirms that the
150 μm pore size electrode performs best among the above three different kinds of porous electrodes.
In addition, the effect of removing chloride ions from the 150 μm porous Ti/SnO2 -Sb2 O3 electrode and
porous Ti/SnO2 –Sb2 O3 /PbO2 electrode is also compared. It can be seen from Figure 7b that the porous
electrode with PbO2 coating has a much better effect on removing chloride ions than the electrode
without PbO2 coating. The reason for this improvement is that the stability and the oxygen evolution
potential of the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode are further improved by the addition of the
PbO2 coating, which is beneficial for achieving an increased removal ratio of chloride ions.

Figure 7. (a) Effect of pore size on chloride ion removal efficiency; (b) chloride ion removal efficiency
as observed with 150 μm porous Ti/SnO2 -Sb2 O3 and 150 μm porous Ti/SnO2 -Sb2 O3 /PbO2 electrodes.

138
Processes 2019, 7, 762

3.3.2. Effect of Initial NaCl Concentration


Related studies have shown that the initial ion concentration can affect the chloride ion removal
ratio and cell voltage. As shown in Figure 8, with the electrolysis time fixed at 4 h, the plate spacing at
1.0 cm, and the current density at 100 mA cm−2 , the removal ratio of chloride ions and the cell voltage
both decrease as the NaCl concentration is increased. The reason for this effect is that the conductivity is
increased and the electrical resistance decreased as the NaCl concentration is increased [1]. In addition,
it could be seen from the figure that the 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrode has a superior
removal effect on chloride ions at low concentrations. Therefore, considering the combined effects
of the removal ratio chloride ions and the cell voltage, 10 g L−1 NaCl solution were selected as the
following study.

Figure 8. (a) Chloride ions removal efficiency under various initial NaCl concentration; (b) change in
cell voltage under various initial NaCl concentrations.

3.3.3. Effect of Current Density


With the electrolysis time at 4 h, the optimal reaction current density was determined for 150 μm
porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes. As shown in Figure 9, with a 150 μm electrode, a NaCl
concentration 10 g L−1 , and the electrode plate spacing 1.0 cm, the removal ratio of chloride ions almost
reached an upper limit at 125 mA cm−2 within 4 h. Current densities higher than 125 mA cm−2 , tended
to damage the electrodes, resulting in reduced electrode lifetimes. So, it could be concluded that
125 mA cm−2 is the optimal current density.

Figure 9. Chloride ion removal efficiency under various current densities.

139
Processes 2019, 7, 762

3.3.4. Effect of Initial pH


As can be seen in Figure 10, with the 150 μm electrode, the NaCl concentration 10 g L−1 , the
electrode plate spacing 1.0 cm, and the current density 125 mA cm−2 , the removal ratio of chloride
ions is shown under the pH from 3 to 11. The solution of pH = 3 and pH = 5 was prepared by adding
diluted concentrated sulfuric acid, and the solution of pH = 9 and pH = 11 was prepared by adding a
diluted sodium hydroxide solution. It can be seen from the figure when the initial pH is the weak
acidity, weak alkali, or neutral conditions, and the removal ratio of chloride ions is very effective, the
removal ratio could reach 98.5% at pH = 9. In addition, the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode has
a good removal effect over the whole range of pH values.

Figure 10. Chloride ion removal efficiency as a function of initial pH value.

3.3.5. Effect of Electrode Plate Spacing


Through the previous work, it was found that a change in the spacing between cathode and anode
plates would cause a change in the cell voltage, which will impact the energy consumption of the
whole process. As can be seen from Figure 11, the cell voltage at a current density of 125 mA cm−2 is
about 11.7 V, when the spacing between both the plates is 1.0 cm and when the 150 μm electrode is
operated in a NaCl solution with a concentration of 10 g L−1 and at an initial pH = 9. Reduction of the
plate spacing to 0.75 cm lowers the cell voltage to about 10.4 V; a further reduction to 9.4 V occurs
when the plate spacing is reduced to 0.5 cm. Therefore, it can be seen that the voltage decreases as the
spacing between both plates is reduced. A reduction in the spacing of the anode and cathode plates
therefore can lower the cell voltage and thereby the power consumption.

140
Processes 2019, 7, 762

Figure 11. Change in cell voltage with electrode plate spacing.

3.3.6. Number of Cycles


With the 150 μm electrode, a NaCl concentration of 10 g L−1 , an initial pH = 9, a current density
125 mA cm−2 , and the spacing between plates 0.5 cm, Figure 12a shows the change in removal ratio of
chloride ions of the porous electrode after repeating the removal process under optimal conditions for
ten times. It can be seen that the removal ratio of chloride ions decreases after 10 times. However,
the removal ratio of chloride ions of the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode is still more than 90%,
which indicates that the porous Ti/SnO2 –Sb2 O3 /PbO2 electrode has great advantages in stability. In
addition, Figure 12b,c show the SEM and XRD patterns of the electrode after recycling, Figure 12c
shows that the XRD spectrum of the electrode has not changed with the previous electrode after
recycling, and Figure 12b shows that the PbO2 deposition coating of the electrode has suffered partial
damage after recycling. By zooming 4000 times, it is found that the PbO2 deposition coating has
developed many cracks, and thus it may be suspected that the reduction in removal rate might be
related to these cracks.

Figure 12. (a) Change in chloride removal rate of a porous Ti/SnO2 –Sb2 O3 /PbO2 electrode during
ten successive removal cycles; (b) Scanning electron microscopy (SEM) pattern of the porous
Ti/SnO2 –Sb2 O3 /PbO2 electrode after recycling; (c) X-ray diffraction (XRD) pattern of the porous
Ti/SnO2 –Sb2 O3 /PbO2 electrode after recycling.

141
Processes 2019, 7, 762

3.3.7. Mechanism of Removing Chloride Ions


During the experiment, the sodium chloride solution was poured into the electrolytic cell. The
chlorine evolution reaction and other side reactions mainly occurred at the anode. The hydrogen
evolution reaction and other side reactions mainly occurred at the cathode. The chloride ions formed
chlorine gas at the anode and the produced chlorine gas was absorbed in the absorption tank, which in
turn resulted in the removal of chloride ions.
Anode reaction:
2Cl− − 2e− → Cl2 ↑ (1)

4OH− − 4e− → 2H2 O + O2 ↑ (2)

Cathode reaction:
2H2 O + 2e− → 2OH− + H2 ↑ (3)

ClO− + 2H2 O + 2e− → 2OH− + Cl− (4)

Side reaction in solution:

Cl2 + 2OH− → ClO− + Cl− + H2 O (5)

Cl2 + H2 O → HClO + H+ + Cl− (6)

HClO + H2 O → H+ + ClO− (7)

It can be seen from the above reaction formulae that the side reaction (2) competes with the
desired chloride removal reaction (1) for holes from the anode. Due to the high oxygen evolution
potential of porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes, reaction (2), however, is inhibited. With reaction
(2) being inhibited, the generation of OH− ions at the cathode via reaction (3) will also be inhibited.
Under alkaline conditions, side reactions (5)–(7) will tend to produce ClO− ions, which, in turn,
will become converted into Cl− ions via reaction (4) and which will finally become removed by
the main reaction (1). As previous studies have shown that the concentrations of hypochlorite and
perchlorate ions which are generated in the electrolysis of sodium chloride solutions are actually quite
small [21], side reactions (5)–(7) are not expected to make any major contribution to the overall chlorine
removal process [22,23]. Overall, our considerations therefore reveal that porous Ti/SnO2 –Sb2 O3 /PbO2
electrodes can efficiently remove chloride ions from aqueous solutions, thus opening very broad
application prospects.

4. Conclusions
In this paper, 150 μm porous Ti/SnO2 –Sb2 O3 /PbO2 electrodes have been demonstrated to exhibit
a good removal effect on chloride ions from NaCl solutions via the process of electrocatalytic oxidation.
Factors that were found to influence the process of electrocatalytic oxidation were initial concentration,
pore size, current density, initial pH, electrode plate spacing, and the number of removal cycles.
Removal ratios of chloride ions up to 98.5% were observed under the following conditions: initial
concentration 10 g L−1 , current density 125 mA cm−2 , electrode plate pore size 150 μm, initial pH = 9, and
electrode plate spacing 0.5 cm. Under these conditions, the porous electrodes exhibited good stability.
Physicochemical and electrochemical characterization results showed that porous Ti/SnO2 –Sb2 O3 /PbO2
electrodes had a high oxygen evolution potential. Overall, it appears that porous Ti/SnO2 –Sb2 O3 /PbO2
electrodes can play an important role in the electrocatalytic oxidation of chloride ions in water, thus
opening prospects for a wide range of applications.

142
Processes 2019, 7, 762

Author Contributions: Conceptualization, K.X.; Data curation, P.C.; Formal analysis, W.G.; Project administration,
P.Y.; Resources, Y.L.; Writing—review & editing, J.P.
Funding: The authors declare no funding.
Acknowledgments: The study was supported by the college of chemistry and molecular sciences at
Wuhan University.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Cui, L.; Li, G.P.; Li, Y.Z.; Yang, B.; Zhang, L.Q.; Dong, Y.; Ma, C.Y. Electrolysis-electrodialysis process for
removing chloride ion in wet flue gas desulfurization wastewater (DW): Influencing factors and energy
consumption analysis. Chem. Eng. Res. Des. 2017, 123, 240–247. [CrossRef]
2. Abdel-Wahab, A.; Batchelor, B. Chloride removal from recycled cooling water using ultra-high lime with
aluminum process. Water Environ. Res. 2002, 74, 256–263. [CrossRef] [PubMed]
3. Woolard, C.R.; Irvine, R.L. Biological treatment of hypersaline wastewater by a biofilm of halophific bacteria.
Water Environ. Res. 1994, 66, 230–235. [CrossRef]
4. Vyrides, I.; Stuckey, D.C. Saline sewage treatment using a submerged anaerobic membrane reactor (SAMBR):
Effects of activated carbon addition and biogas-sparging time. Water Res. 2009, 43, 933–942. [CrossRef]
[PubMed]
5. Walha, K.; Amar, R.B.; Firdaous, L.; Quéméneur, F.; Jaouen, P. Brackish groundwater treatment by
nanofiltration, reverse osmosis and electrodialysis in tunisia: Performance and cost comparison. Desalination
2007, 207, 95–106. [CrossRef]
6. Mansoor, K. New nanopore zeolite membranes for water treatment. Desalination 2010, 251, 176–180.
7. Shaw, W.A. Fundamentals of zero liquid discharge system design. Power 2011, 155, 56–63.
8. Lv, L.; Sun, P.D.; Gu, Z.Y.; Du, H.G.; Pang, X.J.; Tao, X.H.; Xu, R.F.; Xu, L.L. Removal of chloride ion from
aqueous solution by ZnAl-NO3 layered double hydroxides as anion-exchanger. J. Hazard. Mater. 2009, 161,
1444–1449. [CrossRef]
9. Li, H.S.; Chen, Y.H.; Long, J.Y.; Jiang, D.Q.; Liu, J.; Li, S.J.; Qi, J.Y.; Zhang, P.; Wang, J.; Gong, J.; et al.
Simultaneous removal of thallium and chloride from a highly saline industrial wastewater using modified
anion exchange resins. J. Hazard. Mater. 2017, 333, 179–185. [CrossRef]
10. Jiang, S.X.; Li, Y.N.; Bradley, P.L. A review of reverse osmosis membrane fouling and control strategies.
Sci. Total Environ. 2017, 595, 567–583. [CrossRef]
11. Li, D.; Tang, J.Y.; Zhou, X.Z.; Li, J.S.; Sun, X.Y.; Shen, J.Y.; Wang, L.J.; Han, W.Q. Electrochemical degradation
of pyridine by Ti/SnO2 -Sb tubular porous electrode. Chemosphere 2016, 149, 49–56. [CrossRef] [PubMed]
12. Yang, K.; Lin, H.; Liang, S.T.; Xie, R.Z.; Lv, S.H.; Niu, J.F.; Chen, J.; Hu, Y.Y. A reactive electrochemical filter
system with an excellent penetration flux porous Ti/SnO2 -Sb filter for efficient contaminant removal from
water. RSC Adv. 2018, 8, 13933–13944. [CrossRef]
13. An, H.; Li, Q.; Tao, D.J.; Cui, H.; Xu, X.T.; Ding, L.; Sun, L.; Zhai, J.P. The synthesis and characterization of
Ti/SnO2 -Sb2 O3 /PbO2 electrodes: The influence of morphology caused by different electrochemical deposition
time. Appl. Surf. Sci. 2011, 258, 218–224. [CrossRef]
14. Ding, H.Y.; Feng, Y.J.; Liu, J.F. Preparation and properties of Ti/SnO2 -Sb2 O5 electrodes by electrodeposition.
Mater. Lett. 2007, 61, 4920–4923. [CrossRef]
15. Zhao, W.; Xing, J.T.; Chen, D.H.; Bai, Z.L.; Xia, Y.S. Study on the performance of an improved
Ti/SnO2 -Sb2 O3 /PbO2 based on porous titanium substrate compared with planar titanium substrate. RSC Adv.
2015, 5, 26530–26539. [CrossRef]
16. Zhao, W.; Xing, J.T.; Chen, D.H.; Jin, D.Y.; Shen, J. Electrochemical degradation of Musk ketone in aqueous
solutions using a novel porous Ti/SnO2 -Sb2 O3 /PbO2 electrodes. J. Electroanal. Chem. 2016, 775, 179–199.
[CrossRef]
17. Xing, J.T.; Chen, D.H.; Zhao, W.; Peng, X.L.; Bai, Z.L.; Zhang, W.W.; Zhao, X.X. Preparation and characterization
of a novel porous Ti/SnO2 -Sb2 O3 -CNT/PbO2 electrode for anodic oxidation of phenol wastewater. RSC Adv.
2015, 5, 53504–53513. [CrossRef]
18. Sun, J.R.; Lu, H.Y.; Lin, H.B.; Huang, W.M.; Li, H.D.; Lu, J.; Cui, T. Boron doped diamond electrodes based on
porous Ti substrate. Mater. Lett. 2012, 83, 112–114. [CrossRef]

143
Processes 2019, 7, 762

19. Ishibashi, K.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Detection of active oxidative species in TiO2
photocatalysis using the fluorescence technique. Electrochem. Commun. 2000, 2, 207–210. [CrossRef]
20. Zhang, W.L.; Lin, H.B.; Kong, H.S.; Lu, H.Y.; Yang, Z.; Liu, T.T. High energy density PbO2 /activated carbon
asymmetric electrochemical capacitor based on lead dioxide electrode with three-dimensional porous
titanium substrate. Int. J. Hydrogen Energy 2014, 39, 17153–17161. [CrossRef]
21. Neodo, S.; Rosestolato, D.; Ferro, S.; De Battisti, A. On the electrolysis of dilute chloride solutions: Influence
of the electrode material on Faradaic efficiency for active chlorine, chlorate and perchlorate. Electrochim. Acta
2012, 80, 282–291. [CrossRef]
22. Duirk, S.E.; Desetto, L.M.; Davis, G.M. Transformation of Organophosphorus Pesticides in the Presence of
Aqueous Chlorine: Kinetics, Pathways, and Structure−Activity Relationships. Environ. Sci. Technol. 2009, 43,
2335–2340. [CrossRef] [PubMed]
23. Cherney, D.P.; Duirk, S.E.; Tarr, J.C.; Collette, T.W. Monitoring the speciation of aqueous free chlorine from
pH 1 to 12 with Raman spectroscopy to determine the identity of the potent low-pH oxidant. Appl. Spectrosc.
2006, 60, 764–772. [CrossRef] [PubMed]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

144
processes
Article
A Novel Porous Ni, Ce-Doped PbO2 Electrode for
Efficient Treatment of Chloride Ion in Wastewater
Sheng Liu, Lin Gui, Ruichao Peng and Ping Yu *
College of Chemistry & Molecular Science, Wuhan University, Wuhan 430072, China; [email protected] (S.L.);
[email protected] (L.G.); [email protected] (R.P.)
* Correspondence: [email protected]; Tel.: +86-027-6875-2511

Received: 29 March 2020; Accepted: 9 April 2020; Published: 16 April 2020

Abstract: The porous Ti/Sb-SnO2 /Ni-Ce-PbO2 electrode was prepared by using a porous Ti plate as a
substrate, an Sb-doped SnO2 as an intermediate, and a PbO2 doped with Ni and Ce as an active layer.
The surface morphology and crystal structure of the electrode were characterized by scanning
electron microscope(SEM), energy dispersive spectrometer(EDS), and X-Ray diffraction(XRD).
The electrochemical performance of the electrodes was tested by linear sweep voltammetry (LSV),
cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and electrode life test.
The results show that the novel porous Ni-Ce-PbO2 electrodes with larger active surface area have
better electrochemical activity and longer electrode life than porous undoped PbO2 electrodes and flat
Ni-Ce-PbO2 electrodes. In this work, the removal of Cl− in simulated wastewater on three electrodes
was also studied. The results show that the removal effect of the porous Ni-Ce-PbO2 electrode is
obviously better than the other two electrodes, and the removal rate is 87.4%, while the removal
rates of the other two electrodes were 72.90% and 80.20%, respectively. In addition, the mechanism
of electrochemical dechlorinating was also studied. With the progress of electrolysis, we find that
the increase of OH- inhibits the degradation of Cl− , however, the porous Ni-Ce-PbO2 electrode can
effectively improve the removal of Cl− .

Keywords: porous Ni-Ce-PbO2 ; co-doping; active surface area; removal rate

1. Introduction
The widespread use of chlorinated compounds such as HCl, NaCl, and MgCl2 in the industrial
field has increased the content of chloride ion in wastewater [1,2]. If it is discharged into the water
body beyond control, the water environment will be seriously damaged. The accumulation of chloride
ions will make the soil salinized and alkalized, and excessive intake of chlorine by the human body
will cause organ damage. Chloride ions are corrosive to pipelines, boilers, etc., and can erode buildings
and reduce durability of the concrete structure. For example, a large amount of Cl− in desulfurization
wastewater discharged from thermal power plants can corrode pipelines and equipments [3].
At present, the most widely used method for treating Cl− in wastewater is chemical precipitation [4].
However, the concentration of Cl− in the treated wastewater is still high. The treatment of chloride ion
in wastewater by chemical precipitation will be limited in the future [5]. How to treat desulfurization
wastewater in depth, meet the discharge requirements, and reduce the impact on the environment has
always been a difficult problem in the field of wastewater treatment at home and abroad. Therefore,
advanced oxidation processes (AOPs) [6,7] such as catalytic ozonation, Fenton oxidation, supercritical
water oxidation, electrochemical oxidation, photocatalytic oxidation, etc., [8] have been studied.
Electrochemical oxidation is one of the most eye-catching AOPs. It has the advantages of good
treatment effect, small floor area, high degradation efficiency, short residence time and no secondary
pollution, and has broad application prospects in the advanced treatment of salt compounds and

Processes 2020, 8, 466; doi:10.3390/pr8040466 145 www.mdpi.com/journal/processes


Processes 2020, 8, 466

organic compounds [9]. In the recent years, electrochemical oxidation technology has been increasingly
studied for the treatment of chloride ion in wastewater [10].
The degradation efficiency and degradation products of electrochemical oxidation process
change with the anode material [11], which means that anode material is one of the main factors in
electrochemical oxidation process [12,13]. Because of the critical role of anode materials, scholars
have been studying new anodes for many years [14]. The earliest graphite and carbon electrodes
have the disadvantages of low current efficiency and poor mechanical strength [15]. Subsequently,
metal anodes were developed. At present, metal oxide electrode is widely used and its preparation
process is mature, such as dimensionally stable anode (DSA) (e.g., RuO2 , IrO2 , PbO2 , SnO2 ) [11]. With
the deepening of research, 3D porous structure compounds are widely used in the preparation of
anode materials because of their large specific surface area, such as carbon nanotubes (CNT), porous
graphene (GE), etc., [16,17], which can significantly change the structure of the coating when doped in
the metal oxide coating [11]. Scholars have found that doping metal elements in metal oxide coatings
can significantly improve the electrode activity and extend service life [18], such as Ce, Bi, Fe, Co,
etc., [11]. In addition, it has been found that many rare earth oxides doped into PbO2 can significantly
improve the electrochemical performance of the electrode [19]. Kong et al. reported that doping Er2 O3 ,
Gd2 O3 , La2 O3 , and Ce2 O3 on PbO2 electrode could promote the degradation of 4-chlorophenol [20].
Jin et al. [20] found that doping Ce can form smaller crystal size on the surface of the electrode, resulting
in the increase of specific surface area and catalytically active sites. Xia et al. [21] reported that proper
doping of Ni can make the grains dense, which not only facilitates electron movement, and improves
electrochemical performance, but also helps to extend electrode life.
In this work, we intended to use a porous Ti plate as the substrate, because the porous structure has
large specific surface area. We can improve the electrochemical activity and stability of the electrode by
doping Ce and Ni in the active layer PbO2 electrode [22]. The doping of Ce and Ni reduces the crystal
size of the active layer [11,21], but increases the surface area of the active layer and the electrochemical
activity of the electrode, in addition, the service life of the electrode is extended to some extent. Finally,
the electrochemical activity and stability of porous Ni-Ce-PbO2 electrodes were investigated and
compared with the conventional porous undoped PbO2 electrode and the doped flat Ni-Ce-PbO2
electrode. In this experiment, simulated wastewater with high chlorine content was selected as the
target pollutant. The electrochemical performance of the electrode was investigated by comparing the
effects of three kinds of electrodes on the treatment of chloride ions in simulated wastewater.

2. Experimental

2.1. Materials
Porous titanium plate with a purity of 99.9% was purchased from Baoji Jinkai Technology Co.,
Ltd., Jinan, China. Ni(NO3 )2 6H2 O was purchased from Xiqiao Chemical Co. Ltd., Foshan, China. All
other chemicals were purchased from Sino pharm Chemical Reagent Co. Ltd., Shangai, China. All
chemicals were of analytical grade and used as received. In this work, deionized water was used in
all solutions.

2.2. Electrode Preparation

2.2.1. Titanium Surface Preparation


The porous titanium plates (20 mm × 10 mm × 2.8 mm) bought by Baoji Jinkai Technology Co.,
Ltd., Jinan, China, had a purity of 99.9% and an average pore diameter of 50 μm to pretreat the substrate.
The porous titanium plate was ultrasonically cleaned in acetone for 15 min, washed in 20% NaOH
solution at 90 ◦ C for 1 h, and etched in a 15 wt % oxalic acid solution at 90 ◦ C for 1 h until a gray matte
titanium matrix was formed, and finally saved in ethanol.
The surface treatment of the flat titanium plate (20 mm × 10 mm × 2.8 mm) is similar to porous
titanium plate.

146
Processes 2020, 8, 466

2.2.2. Coating SnO2 –Sb2 O3


The electrode intermediate layer was prepared by thermal decomposition method [11]. Total
of 1.2 g SnCl4 , 0.2 g Sb2 O3 , and 10 mL concentrated hydrochloric acid were dissolved in 25 mL of
isopropanol to obtain a precursor-coating solution. The solution was colorless, transparent, slightly
sticky. Then the pretreated porous titanium plate was immersed in the precursor solution. After
soaking, it was taken out and dried in an oven at 110 ◦ C, and then calcined in a muffle furnace at 500 ◦ C
for 15 min. After cooling the plate, the drying and calcining process was repeated several times, and
the high-temperature baking time in the last time was extended to 1 h to obtain a porous Ti/Sb-SnO2
electrode. The main purpose of this layer is to improve the conductivity of the electrode and prevent
the titanium matrix from being oxidized to form TiO2 .
The precursor solution can be directly brushed on the surface of the flat titanium plate with a
brush. Other preparation processes of the intermediate layer of the flat electrode are similar to that of
the porous electrode.

2.2.3. Electrochemical Deposition Ni–Ce–PbO2


Pb(NO3 )2 , Ni(NO3 )2 , and Ce(NO3 )2 were dissolved in 250 mL water at a ratio of 100:1:1 of Pb, Ni,
and Ce. Then, 0.04 M NaF and 4 mL/L of PTFE were added, before adding 0.1 mol/L HNO3 to adjust
the pH to 1, to form an electrodeposition solution. The control temperature was 65 ◦ C, the current
density was 20 mA/cm2 , the electrodeposition time was 1 h, so as to deposit a surface layer of lead
dioxide with Ce, Ni co-doped on the surface of the intermediate layer tin antimony oxide, that is,
porous Ti/Sb-SnO2 /Ni-Ce-PbO2 electrode.
In addition to the raw materials (Ni (NO3 )2 and Ce (NO3 )2 ), the preparation process of porous
PbO2 electrode is similar to that of porous Ni-Ce PbO2 electrode

2.3. Electrode Characterization


The surface morphology was observed using a scanning electron microscope (Quanta 200 of
FEI, Hillsboro, OR, USA). X-ray diffraction (XRD) patterns of samples were obtained with an X-ray
diffractometer (PANalytical, Almelo, The Netherlands). Cyclic voltammetry (CV), linear sweep
voltammetry (LSV), and electrochemical impedance spectroscopy (EIS) were performed at room
temperature using a computer-controlled electrochemical workstation (CHI 660E, CH Instruments,
Shanghai, China) with a conventional three-electrode system. The prepared PbO2 -based electrode
(20 mm × 20 mm) was used as the working electrode, a saturated Ag/AgCl electrode was employed as
the reference electrode, and a stainless steel sheet was applied as the counter electrode. All potentials
were referred to the SCE. The stability tests (up to 20 h) were performed by the accelerated life test with
a current density of 1 A·cm−2 and a temperature of 60 ◦ C in 2 M H2 SO4 solution for porous undoped
PbO2 electrodes, porous Ni-Ce–PbO2 electrodes, and Flat Ni-Ce–PbO2 electrodes. These tests were
performed in a three-electrode system.

2.4. Electrochemical Oxidation


A simulated chlorine-containing wastewater with a chloride ion concentration of 4 g/L and a
volume of 200 mL was selected as the experimental wastewater; the temperature was 298 K in all
experiments. The electrochemical oxidation experiment was conducted by a batch method, and the
device was mainly composed of a DC power source, a collector types magnetic stirrer, and a glass
reactor. The anode (PbO2 -based electrodes) and the cathode (flat titanium plate) were placed parallel
to each other and perpendicular to the solution level with a distance of 1 cm. The volume of simulated
wastewater in all experiments was 200 mL and the Cl− concentration was 4 g/L. The temperature
of all experiments was maintained at 298 K. All pH values in the experiment were determined by a
pH meter and all Cl− concentration in the experiment were determined by titration with a standard
AgNO3 solution.

147
Processes 2020, 8, 466

3. Results and Discussion

3.1. Surface Morphological and Crystallographic Analysis


The SEM images of the plate-modified Ni-Ce-PbO2 electrode, the porous undoped PbO2 electrode,
and the porous Ni-Ce-PbO2 electrode are shown in Figure 1. It can be seen that the porous undoped
PbO2 electrode and the porous Ni-Ce-PbO2 electrode still have many small holes in the electrode
surface while compared with porous Ti plate (Figure 1a), indicating that the PbO2 coating did not block
the porous structure. From Figure 1b,f it can be seen that the porous Ni-Ce-PbO2 electrodes apparently
have larger pores and specific surface area compared to the flat Ni-Ce-PbO2 electrodes.

(a) (b)

50 um

(c) (d)

(e) (f)

Figure 1. SEM of different electrodes, (a) (porous titanium plate, 80×); (b) (flat Ni-Ce-PbO2 electrode,
1200×) and (c) (porous undoped PbO2 electrode, 80×); (d) (porous undoped PbO2 electrode, 1200×)
and (e) (porous Ni-Ce-PbO2 electrode, 80×); (f) (porous Ni-Ce-PbO2 electrode, 1200×).

148
Processes 2020, 8, 466

The porous PbO2 electrode doped with Ni and Ce (Figure 1e,f) can make the holes smaller or bigger
on the surface of the electrodes, and the specific surface area was larger than porous undoped PbO2
electrodes (Figure 1c,d). Besides, the EDS spectrum of different PbO2 electrodes is shown in Figure 2a,
which confirms that there are O, Pb, Ce, Ni elements in the porous Ni-Ce-PbO2 electrode, while there
are only O and Pb elements in the porous undoped PbO2 electrodes. Thus, it can be concluded that the
Ni and Ce was successfully doped into the PbO2 films and doping of Ni and Ce can reduce the grain
size of the PbO2 coating.

(a) (b)
Figure 2. (a) EDS of a (porous PbO2 electrode) and b (porous Ni-Ce-PbO2 electrode); (b) XRD diffraction
patterns of different Electrodes A (Porous Ti/Sb-SnO2 /PbO2 electrode), B (Porous Ti/Sb-SnO2 /Ni-Ce-PbO2
electrode), and C (Standard XRD pattern of PbO2 ).

From Figure 2b, it can be seen that the diffraction peaks of undoped porous show that dioxide
electrodes appear at 25.4 degrees, 31.9 degrees, 36.1 degrees, 48.9 degrees, 58.8 degrees, 62.4 degrees,
and 66.8 degrees, which are consistent with the JCPDS card (41–1492) in pattern, indicating that the
main component of undoped PbO2 surface layer is β-PbO2 and the coating of porous Ni-Ce-PbO2
electrode consists of a mixture of crystalline phases of α and β-PbO2 . The content of α-PbO2 is higher
than that of undoped PbO2 because the doping of Ce changes the preferred crystalline orientation of
the electrode surface and forms smaller grains. In addition, compared with the undoped PbO2 , the
intensity of the diffraction peaks of Ni-Ce-PbO2 decreases or even disappears because of the doping of
Ni and Ce. The reason may be that doping of Ni and Ce changes the nucleation and growth of crystals
in the coating, making the electrode have smaller crystal size than the undoped PbO2 electrode, which
can also be seen from the SEM image. The average crystallite size calculated from the width of [101]
diffraction peaks by Scherrer’s formula is 17.64 nm (Ce–Ni) and 28.76 nm (undoped). According to
literature [20], the diffraction peak width is inversely proportional to the crystallite size. The result
indicates that the deposited Ni-Ce-PbO2 has smaller crystallite size than other electrode. Smaller crystal
size may help to form larger specific surface area which may lead to better electrochemical performance.

3.2. Electrochemical Performance Test


As shown in Figure 3a, no redox peak signal was observed on any of the electrodes in the blank
Na2 SO4 solution, indicating that PbO2 is an electrochemically inert material in the blank Na2 SO4
solution. After the addition of Cl− (Figure 3b), a distinct oxidation peak was obtained. There is no
doubt that the oxidation peak is attributed to the oxidation of Cl− on the surface of the anode. However,
no corresponding reduction peak was observed in the reverse scanning from 3 V to 0 V, indicating that
the oxidation of Cl− is a completely irreversible electrode reaction process. The oxidation peak potential
of the porous Ni-Ce-PbO2 electrode (2.01 V vs. SCE) was lower than that of the other two electrodes

149
Processes 2020, 8, 466

(2.24 and 2.48 V), but the oxidation peaks current (0.031 A) was significantly higher than the other
two electrodes (0.028 A and 0.024 A), which showed that the porous Ni-Ce-PbO2 electrode has higher
electro-catalytic activity for Cl− , and the improvement of its electro-catalytic activity is not only related
to the increase in electrode surface area caused by its porous structure and doping with Ni and Ce, but
also due to changes in the PbO2 band structure. The electrode doping Ni and Ce not only increases the
donor area of PbO2 , but also increases the donor level of PbO2 , making it easier for electrons to jump
from the donor level to the conduction band [21]. Therefore, the conductivity of PbO2 is improved.

(a) (b)

(c) (d)

(e) (f)

Figure 3. (a) Cyclic voltammetry (CV) of different PbO2 electrodes measured in 0.1mol·L−1 Na2 SO4
solution, (b) 0.1 mol·L−1 Na2 SO4 solution (pH = 6.5) containing 4 g L−1 Cl− , scan rate: 50 mV s−1 ,
T = 298 K, (c) LSV curves of different electrodes measured in 0.5 M Na2 SO4 , scan rate: 10 mV s−1 ,
T = 298 K, (d) EIS of different electrodes. Conditions: T = 298 K; [H2 SO4 ] = 1 M. (e) Electrode stability
tests: electrode potential vs. time for the electrolysis using different electrodes. (f) The mass losses of
electrodes after accelerated life tests for 20 h.

150
Processes 2020, 8, 466

The oxygen evolution overpotential (OEP) of different electrodes could be measured by LSV.
According to the polarization curve in Figure 3c, the OEP of the porous Ni-Ce-PbO2 electrode was
the highest with 2.09 V (vs. SCE) compared with the flat Ni-Ce-PbO2 electrode of 1.81 V (vs. SCE)
and the porous PbO2 electrode of 1.91 V (vs. SCE), respectively. The electrodes with high OEP values
can produce more hydroxyl radicals [23]. In addition, it can be seen from Figure 3c that the porous
Ni-Ce-PbO2 electrodes had higher oxygen evolution current than the other two electrodes. The higher
oxygen releases current also verified that the active surface area of the porous Ni-Ce-PbO2 electrode is
larger than that of the other PbO2 -based electrodes.
In Figure 3d, there is an obvious semicircle in the electrochemical impedance spectroscopy (EIS)
of the three electrodes. The radius of the semicircle usually reflects the magnitude of the transfer
resistance Rct of the chlorine evolution or oxygen evolution reaction, which is available from the EIS
spectrum. The Rct sizes of Ni-Ce-PbO2 electrodes, porous PbO2 electrodes, and porous Ni-Ce-PbO2
electrodes were 59.4, 21.2, and 12.2, respectively. It is indicated that the porous Ni-Ce-PbO2 electrode
has the best conductivity, the most active surface sites, and the highest chlorine evolution activity.
The service life is a critical factor in the practical application of the electrode. Under the condition
of anode current of 1A pa−2 and temperature of 60 ◦ C, the accelerated life test of different PbO2
electrodes was carried out in 2M H2 SO4 solution for 20 h. According to Figure 3e,f, the Ni-Ce-PbO2
electrode exhibited better electrochemical stability and less mass loss than the pure PbO2 electrode.
It is well-known that the main reasons of mass loss of electrode are the separation and dissolution of
PbO2 film, and the mass loss is proportional to the service life of the electrode. Therefore, it can be
concluded that the stability of Ni-Ce-PbO2 electrode is much better than that of pure PbO2 electrode by
the modification of Ni and Ce. The mass loss of the porous Ni-Ce-PbO2 electrode is slightly larger
than that of the flat Ni-Ce-PbO2 , mainly because of the loose porosity of the structure. In fact, in
the experiments, we found that film peeled off on the pure PbO2 electrode, but not on the other two
modified PbO2 electrodes.

3.3. Electrochemical Oxidation of Cl− in Simulated Wastewater


The electrochemical oxidation of simulated wastewater was compared with three kinds of
electrodes. From Figure 4a, after electrolysis for 100 min, the removal rate of Cl− on the porous
Ni-Ce-PbO2 electrode was as high as 87.4%. At the same time, the plate Ni-Ce-PbO2 electrode and the
porous PbO2 electrode were used as anodes, and the highest removal rates of Cl− were 72.90% and
80.20%, respectively. Obviously, the above results show that the electrochemical oxidation ability of the
porous Ni-Ce-PbO2 electrode was much higher than the other two electrodes. This may be due to the
fact that the porous structure increases the surface area of the electrode and the doping of Ni and Ce
further increases the surface area of the electrode, which facilitates the adsorption of Cl− on the surface
of the anode and promotes the mass transfer and exchange of the reactants. Thus, the electrochemical
oxidation ability of the electrode is improved. As is shown in Figure 4b that in the early stage of
electrolysis, the removal rate of Cl− was higher than that in the later stage because of the enrichment
of Cl− in wastewater. The results show that higher current density result in higher Cl− removal rate,
and the difference in Cl− removal rate was not obvious at relatively higher current densities. With the
progress of electrolysis, the decrease of Cl− concentration inhibited the oxidation of Cl− and reached
equilibrium at a certain time, in which Cl− cannot be completely removed. However, the porous
Ni-Ce-PbO2 electrode with large specific surface area can effectively improve the removal rate of Cl− .
During the process of degradation of Cl− , the chlorine evolution reaction and oxygen evolution side
reaction in the anode proceed simultaneously, the content of Cl− is higher and the content of OH− is
lower in the initial stage of reaction process (Figure 4c). The reaction formulas are as follows:

Anode: 2Cl− → Cl2 + 2e− (1)

Cl2 + 2H2 O → HClO+H3 O+ + Cl− (2)

151
Processes 2020, 8, 466

HClO + H2 O → H3 O+ + ClO− (3)


+ −
Cathode: H2 O → H + OH (4)
+ −
2H +2e → H2 (5)

(a) (b)

(c) (d)

Figure 4. (a) Variation of Cl−removal percentage with time during electrochemical oxidation on
different electrodes, conditions: current density = 50 mA/cm2 ; T = 298 K; [Cl− ] = 4 g/L, (b) at different
current densities, conditions: porous Ni-Ce-PbO2 electrode; T = 298 K; [Cl− ] = 4 g/L and electrochemical
reaction process of electrode dechlorinating (c) at 30 min and (d) at 100 min.

According to the investigation [24], under the experimental conditions, the anode mainly generates
Cl2 (Equation (1)). The cathode produces a large amount of OH− , which exists in the solution to raise
the pH of the solution quickly, except a small portion is transferred to the anode for consumption.
The increase of OH− concentration promotes the disproportionation of chlorine and forms a large
number of ClO− (Equations (2) and (3)). In addition, Figure 4a also shows the fact that the Cl− removal
rate increases significantly with the increase of the initial concentration. But after the electrolysis
time reaches 90 min, a large amount of Cl− is removed and the OH− concentration in the solution is
gradually increased. This can also be seen from Figure 4c,d, the reaction formulas are as follows:

H2 O → H+ + OH−
Anode: 2OH− − 4e− → O2 + 2H+ (6)
Cathode: 4H+ + 4e− →2H2

A large amount of OH groups accumulates on the surface of the anode, which hinders the contact
between a small amount of Cl− in the solution and the anode, so that the oxygen evolution reaction
on the surface of the anode gradually becomes the main reaction. The consumption of OH− on the
anode results in a small part of residual Cl− in the solution which cannot be removed, and another

152
Processes 2020, 8, 466

small part of Cl− is converted into ClO− and HClO [24] (Equations (1)–(3)), which greatly limits the
removal of Cl− . Also as a DSA electrode, the porous Ni-Ce-PbO2 electrode degraded 87.4% of Cl− at
90 min because of its large active surface area, while the degradation rate of the plated Ni-Ce-PbO2
electrode and the porous PbO2 electrode were only 72.90% and 80.20%, respectively. Therefore, the
porous Ni-Ce-PbO2 electrode breaks through the limitation of the conventional electrode in removing
Cl− . In addition, we also detected the concentration of Pb2+ and Ni2+ in the solution after the 5th
oxidation, and found that the concentration of Pb2+ was only about 0.0085 mg/L, and Ni2+ was even
less, far lower than the standard of the World Health Organization. It shows that it is environment
friendly and will not cause secondary pollution.

4. Conclusions
In this paper, porous Ni-Ce-doped PbO2 electrodes were successfully prepared on a porous
titanium substrate by thermal deposition and electrodeposition. The surface morphology, crystal
structure, electrochemical activity, and electrode stability of the flat Ni-Ce-PbO2 , porous undoped
PbO2 , and porous Ni-Ce-PbO2 were tested, and the electro-catalytic properties of these three electrodes
in simulated wastewater were compared.
The porous Ni-Ce-PbO2 electrodes possessed porous structure and smaller grain size than the
other two electrodes. The doping of Ni and Ce can change the nucleation and growth of crystals on
the surface of the electrode, making the electrode have smaller particles, larger electrochemical active
surface area, and better electrode life.
At a current density of 50 mA, the porous Ni-Ce-PbO2 electrode was used to treat Cl− in the
simulated wastewater. The removal rate was as high as 87.4%, while the highest removal rates of the
porous pure PbO2 electrode and the flat Ni-Ce-PbO2 electrode were 72.90% and 80.20%, respectively.
In addition, in the later stage of Cl− oxidation, because of the increase of pH of the electrolyte,
oxygen evolution reaction mainly occurs in the anode, which results in a part of Cl− that cannot
be removed from the solution and the other part that dissolves in the solution in the form of ClO−
and HClO; removal rate of Cl− is restricted. The novel porous Ni-Ce-PbO2 electrodes can effectively
improve the removal of Cl− because of its greater electrochemically active surface area.

Author Contributions: S.L. and P.Y. conceived and designed the experiments; S.L., L.G., and R.P. performed the
experiments; S.L. analyzed the data and wrote the paper; S.L. and P.Y. revised the manuscript. All authors have
read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Guàrdia, M.D.; Guerrero, L.; Gelabert, J.; Gou, P.; Arnau, J. Sensory characterisation and consumer
acceptability of small calibre fermented sausages with 50% substitution of NaCl by mixtures of KCl and
potassium lactate. Meat Sci. 2008, 80, 1225–1230. [CrossRef] [PubMed]
2. Xia, J.; Liu, Q.F.; Mao, J.H.; Qian, Z.H.; Jin, S.J.; Hu, J.Y.; Jin, W.L. Effect of environmental temperature
on efficiency of electrochemical chloride removal from concrete. Constr. Build. Mater. 2018, 193, 189–195.
[CrossRef]
3. Yang, B.; Chen, Z.; Zhang, M.; Zhang, H.; Zhang, X.; Pan, G.; Zou, J.; Xiong, Z. Effects of elevated atmospheric
CO2 concentration and temperature on the soil profile methane distribution and diffusion in rice-wheat
rotation system. J. Environ. Sci. (China) 2015, 32, 62–71. [CrossRef] [PubMed]
4. Gupta, V.K.; Ali, I.; Saleh, T.A.; Nayak, A.; Agarwal, S. Chemical treatment technologies for waste-water
recycling—An overview. RSC Adv. 2012, 2, 6380–6388. [CrossRef]
5. Hu, S.; Ding, S.F.; Fan, Z.S. Zero release technology of desulfurization waste water in coal—Fired power
plant. Clean Coal Technol. 2015, 21, 129–133.
6. Liu, H.; Liu, Y.; Zhang, C.; Shen, R. Electrocatalytic oxidation of nitrophenols in aqueous solution using
modified PbO2 electrodes. J. Appl. Electrochem. 2008, 38, 101–108. [CrossRef]

153
Processes 2020, 8, 466

7. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015,
1, 167–176. [CrossRef]
8. Shmychkova, O.; Luk’yanenko, T.; Velichenko, A.; Meda, L.; Amadelli, R. Bi-doped PbO2 anodes:
Electrodeposition and physico-chemical properties. Electrochim. Acta 2013, 111, 332–338. [CrossRef]
9. Xu, M.; Wang, Z.; Wang, F.; Hong, P.; Wang, C.; Ouyang, X.; Zhu, C.; Wei, Y.; Hun, Y.; Fang, W. Fabrication of
cerium doped Ti/nanoTiO2 /PbO2 electrode with improved electrocatalytic activity and its application in
organic degradation. Electrochim. Acta 2016, 201, 240–250. [CrossRef]
10. Shuangchen, M.; Jin, C.; Gongda, C.; Weijing, Y.; Sijie, Z. Research on desulfurization wastewater evaporation:
Present and future perspectives. Renew. Sustain. Energy Rev. 2016, 58, 1143–1151. [CrossRef]
11. Duan, X.; Zhao, Y.; Liu, W.; Chang, L.; Li, X. Electrochemical degradation of p-nitrophenol on carbon
nanotube and Ce-modified-PbO2 electrode. J. Taiwan Inst. Chem. Eng. 2014, 45, 2975–2985. [CrossRef]
12. Dai, Q.; Xia, Y.; Chen, J. Mechanism of enhanced electrochemical degradation of highly concentrated aspirin
wastewater using a rare earth La-Y co-doped PbO2 electrode. Electrochim. Acta 2016, 188, 871–881. [CrossRef]
13. Xia, Y.; Dai, Q.; Chen, J. Electrochemical degradation of aspirin using a Ni doped PbO2 electrode. J. Electroanal.
Chem. 2015, 744, 117–125. [CrossRef]
14. Yao, Y.; Teng, G.; Yang, Y.; Huang, C.; Liu, B.; Guo, L. Electrochemical oxidation of acetamiprid using
Yb-doped PbO2 electrodes: Electrode characterization, influencing factors and degradation pathways.
Sep. Purif. Technol. 2019, 211, 456–466. [CrossRef]
15. Elaissaoui, I.; Akrout, H.; Grassini, S.; Fulginiti, D.; Bousselmi, L. Effect of coating method on the structure
and properties of a novel PbO2 anode for electrochemical oxidation of Amaranth dye. Chemosphere 2019, 217,
26–34. [CrossRef] [PubMed]
16. Xu, F.; Chang, L.; Duan, X.; Bai, W.; Sui, X.; Zhao, X. A novel layer-by-layer CNT/PbO2 anode for high-efficiency
removal of PCP-Na through combining adsorption/electrosorption and electrocatalysis. Electrochim. Acta
2019, 300, 53–66. [CrossRef]
17. Zhou, X.; Liu, S.; Yu, H.; Xu, A.; Li, J.; Sun, X.; Shen, J.; Han, W.; Wang, L. Electrochemical oxidation of
pyrrole, pyrazole and tetrazole using a TiO2 nanotubes based SnO2 -Sb/3D highly ordered macro-porous
PbO2 electrode. J. Electroanal. Chem. 2018, 826, 181–190. [CrossRef]
18. Santos, J.E.L.; de Moura, D.C.; da Silva, D.R. Application of TiO2 -nanotubes/PbO2 as an anode for the
electrochemical elimination of Acid Red 1 dye. J. Solid State Electrochem. 2018, 23, 351–360. [CrossRef]
19. Du, H.; Duan, G.; Wang, N.; Liu, J.; Tang, Y.; Pang, R.; Chen, Y.; Wan, P. Fabrication of Ga2 O3 –PbO2 electrode
and its performance in electrochemical advanced oxidation processes. J. Solid State Electrochem. 2018, 22,
3799–3806. [CrossRef]
20. Jin, Y.; Wang, F.; Xu, M.; Hun, Y.; Fang, W.; Wei, Y.; Zhu, C.G. Preparation and characterization of Ce and PVP
co-doped PbO2 electrode for waste water treatment. J. Taiwan Inst. Chem. Eng. 2015, 51, 135–142. [CrossRef]
21. Wang, Z.; Xu, M.; Wang, F.; Liang, X.; Wei, Y.; Hu, Y.; Zhu, C.G.; Fang, W. Preparation and characterization
of a novel Ce doped PbO2 electrode based on NiO modified Ti/TiO2 NTs substrate for the electrocatalytic
degradation of phenol wastewater. Electrochim. Acta 2017, 247, 535–547. [CrossRef]
22. Yao, Y.; Huang, C.; Yang, Y.; Li, M.; Ren, B. Electrochemical removal of thiamethoxam using three-dimensional
porous PbO2 -CeO2 composite electrode: Electrode characterization, operational parameters optimization
and degradation pathways. Chem. Eng. J. 2018, 350, 960–970. [CrossRef]
23. Xie, R.; Meng, X.; Sun, P.; Niu, J.; Jiang, W.; Bottomley, L.; Li, D.; Chen, Y.; Crittenden, J. Electrochemical
oxidation of ofloxacin using a TiO2 -based SnO2 -Sb/polytetrafluoroethylene resin-PbO2 electrode: Reaction
kinetics and mass transfer impact. Appl. Catal. B Environ. 2017, 203, 515–525. [CrossRef]
24. Neodo, S.; Rosestolato, D.; Ferro, S.; De Battisti, A. On the electrolysis of dilute chloride solutions: Influence
of the electrode material on Faradaic efficiency for active chlorine, chlorate and perchlorate. Electrochim. Acta
2012, 80, 282–291. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

154
processes
Article
Experimental Study of Micro Electrochemical
Discharge Machining of Ultra-Clear Glass with
a Rotating Helical Tool
Yong Liu 1, *, Chao Zhang 2 , Songsong Li 1 , Chunsheng Guo 1,3 and Zhiyuan Wei 1
1 Associated Engineering Research Center of Mechanics & Mechatronic Equipment, Shandong University,
Weihai 264209, China; [email protected] (S.L.); [email protected] (Z.W.)
2 Department of Mechanical Engineering, Weihai Vocational Secondary School, Weihai 264213, China;
[email protected]
3 Shenzhen Research Institute of Shandong University, Virtual University Park, Nanshan, Shenzhen 518057,
China; [email protected]
* Correspondence: [email protected]; Tel.: +86-1356-312-3255

Received: 12 March 2019; Accepted: 29 March 2019; Published: 4 April 2019

Abstract: Electrochemical discharge machining (ECDM) is one effective way to fabricate


non-conductive materials, such as quartz glass and ceramics. In this paper, the mathematical model
for the machining process of ECDM was established. Then, sets of experiments were carried out
to investigate the machining localization of ECDM with a rotating helical tool on ultra-clear glass.
This paper discusses the effects of machining parameters including pulse voltage, duty factor, pulse
frequency and feed rate on the side gap under different machining methods including electrochemical
discharge drilling, electrochemical discharge milling and wire ECDM with a rotary helical tool. Finally,
using the optimized parameters, ECDM with a rotary helical tool was a prospective method for
machining micro holes, micro channels, micro slits, three-dimensional structures and complex closed
structures with above ten micrometers side gaps on ultra-clear glass.

Keywords: electrochemical discharge machining; rotating helical tool; side gap; micro structures;
closed structure; ultra-clear glass

1. Introduction
In recent years, ECDM has gained attention. Micro electromechanical systems (MEMS),
including micro reactors and micro medical devices, often consist of the micro structures of
nonconductive materials, such as glass, ceramics and silicon nitride. Therefore, the traditional
machining method is difficult to use to fabricate micro structures composed of brittle and hard
nonconductive materials. However, ECDM can machine micro structures on hard nonconductive
materials. ECDM is a hybrid machining method including electrochemical machining and electric
discharge machining [1]. When discharge takes place between the tool electrode and the surrounding
electrolyte, local high temperatures and chemical reactions remove the workpiece material. Micro
structures have been widely applied to micro accelerometers, micro pumps, micro containers and
biological medical instruments, which could be machined by electrochemical machining (ECM), electro
discharge machining (EDM) or ECDM [2–5]. Glass has superior properties, including transparency,
high oxidation resistance, wear resistance, biological compatibility, and low electrical conductivity
properties. ECDM, with a different machining method, can fabricate complex micro structures on
glass, such as micro holes, micro channels, micro slits and complicated three-dimensional features.
ECDM was first put forward by Kurafuji in 1968. Because of machining brittle and hard
nonconductive material at the micro level, this machining method was investigated further. Nasim,

Processes 2019, 7, 195; doi:10.3390/pr7040195 155 www.mdpi.com/journal/processes


Processes 2019, 7, 195

Mohammad researched the generation of single gas bubbles at the tool electrode surface. Finally,
he found that the wettability of the tool electrode and the surface tension between the bubble and
electrolyte affected the gas film thickness [6]. Zhang and Huang explored critical voltage under
different machining conditions by using ECDM on glass to investigate the time it took to form the gas
film, via the mean current of discharge. They concluded that better machining precision and surface
quality could be obtained by selecting optimized parameters [7]. Sathisha proposed the empirical
model for the process of machining grooves with multiple regression analysis [8]. Jawalkar fabricated
micro channels by electrochemical discharge milling. The experiment results showed that voltage
plays a leading role in the parameters of both material removal rate and tool wear [9]. Cao found
a new method indicating that the grinding process under polycrystalline diamond tools reduced
the surface roughness of ECDM structures from a few tens of a micron to 0.05 μm Ra [10]. Elhami
utilized special equipment to generate only a single discharge in ultrasonic-assisted electrochemical
discharge machining (UAECDM) and studied two important characteristics: material removal and
tool wear [11]. Many scholars conducted further studies of UAECDM [12–14]. Han and Min proposed
a method of using the side insulation tool and low concentration electrolytes to reduce undesirable over
cutting [15]. Furutani concluded that the width, depth and surface roughness of grooves machined by
electrochemical discharge milling increased with higher voltage [16]. Kun investigated the precision
and stability of quartz fabricated by ECDM and explored optimal machining parameters including the
size of the electrode and the machining speed [17].
Wire electrochemical discharge machining (WECDM) was proposed by Tsuchiya [18]. Jain utilized
traveling wire as a tool in WECDM, and studied the effects of voltage, the concentration of the
electrolyte on material removal rate and tool wear [19]. Panda and Yadava established a 3D finite
element transient thermal model and predicted the temperature field and MRR in traveling wire
electrochemical spark machining (TW-ECSM) [20]. Kuo found a new wire ECDM approach to machine
quartz glass. In their experiments, electrolytes were supplied by titrated flow and the machining
quality and efficiency were improved [21]. Wang studied the surface integrity of alumina machined
by WECDM [22]. A host of literature proved that electrolyte circulation plays an important role in
machining performance. Many approaches to enhancing the electrolyte circulation in ECDM and
wire ECDM have been proposed [23–25]. Fang used rotary helical electrodes in wire ECDM, which
accelerated the cycle of the electrolyte [26]. Wang and Zhang researched the flow field of ECDM with
a rotating helical tool. In their experiments, the gas–liquid phase distribution and the velocity vectors
of the electrolyte in the machining gap were investigated [27].
In this paper, a rotating helical electrode was used in different ECDM processes, including
electrochemical discharge drilling, electrochemical discharge milling and wire ECDM. The rotary
helical electrode produced an axial velocity and axial force, dragging the electrolyte from the bottom
of the workpiece. Therefore, the machining accuracy of ECDM is fine. The machining parameters,
including voltage, frequency, duty factors, and feed rate, were considered in the experiments and
their effect on the side gap was investigated. The optimized parameters were utilized to successfully
machine micro holes, micro channels, micro slits and complicated three-dimensional features with ten
several-micron side gaps.

2. Experimental Set-Up and Model for Machining Process

2.1. Experimental Set-Up


Most past efforts have been spent on studying the mechanisms of ECDM. The ECDM process
can be depicted as in Figure 1. In all of the following experiments, the tool electrode with Φ105 μm
was a rotating helical tungsten carbide (WC) electrode while the auxiliary anode is a graphite plate
and the electrolyte was 3 mol/L KOH (Shuangshuang chemical industry, Yantai, China). This process
can be divided into five steps. In the first step the pulse power was imposed on the tool electrode
and the auxiliary anode, which were immersed in the electrolyte. Because of electrolysis, hydrogen

156
Processes 2019, 7, 195

and oxygen gas bubbles were generated around the tool electrode and auxiliary anode, respectively.
The second step involved the hydrogen gas bubbles accumulating rapidly and embracing the tool
electrode. The third step was when the formation rate of the hydrogen gas bubbles was equal to
the rate of that escaping from the electrode. The gas film around the tool electrode was formed and
completely separated the tool electrode from the surrounding electrolyte. In the fourth step there was
a narrow gap between the tool electrode and the electrolyte according to the third step. When the
applied voltage rose to a critical value, there was a spark in the gas film. As is known, a large amount
of heat generated by discharge will instantaneously melt the surface material of the workpiece when
the tool electrode is close to the workpiece. In addition, some material is removed due to evaporation
and localized high temperature, leading KOH electrolytes to corrode the workpiece. The fifth step
began when a gas film was staved when the tool electrode contacted with the electrolyte again. Then,
the process switched back to the first step, beginning the cycle anew.

Figure 1. A schematic view of electrochemical discharge machining (ECDM).

The architecture of this experimental system for ECDM is illustrated in Figure 2. The experimental
system contains four subsystems: the power supply system, machine tool system, microelectrode
system, and processing control and monitoring system. The power supply system was plays
a significant part in ECDM, which provides a series of variable ranges including pulse voltage,
duty factor, and pulse frequency. The machine tool system is mainly comprised of an optical precision
platform, the L shaped marble frame, feed device, high speed motorized spindle, lifting platform,
fixture, and other components. The optical precision platform ensured high accuracy for micro
ECDM. To guarantee the verticality of the machine tool, the L shaped marble frame possessing
vibration isolation performance was used. The feed device, controlled by the MP-C154 motion control
card, accurately controlled the feeding of the electric slipway along the three directions and met the
requirements for fabricating complex three-dimensional micro structures. The microelectrode system
consisted of a rotary helical WC electrode, electrolytic bath, high speed motorized spindle, fixture,
and lifting platform. The glass workpiece was fixed on the electrolytic bath and placed on the lifting
platform. The processing control and monitoring system had the motion control card and Supereyes.
Supereyes monitored the process and captured images. In this research, electrochemical discharge
drilling, electrochemical discharge milling, and wire ECDM were utilized to machine micro structures
on glass workpieces with rotary helical WC electrodes via an experimental system.

157
Processes 2019, 7, 195

Figure 2. Experimental system of ECDM.

2.2. Establishing of Machining Process Model


To investigate the side gap in the ECDM process, three different types of experiments were carried
out, including electrochemical discharge drilling, electrochemical discharge milling, and wire ECDM.
During a certain specified experiment, only one parameter could be adjusted and the effect on the side
gap recorder, all other parameters remained constant.
Step 1 was the model for electrochemical discharge drilling. Establishing the simplified model of
electrochemical discharge drilling on the side gap needed the following hypothetical conditions:
(a) The mean heat released by the discharges q is linear to the energy for melting material in unit
time, for which the linear coefficient is the constant k.
(b) The hole after drilling is a uniform cylinder.
(c) The distance between the end of the rotary helical electrode and the bottom of the hole is assumed
to be constant and this constant is c, shown in Figure 3.

Figure 3. A schematic view of electrochemical discharge drilling.

The discharge energy q in unit time can be obtained by the equation proposed by Jain [28]:

q = U I − RI 2 (1)

where U is voltage, I is the mean current, and R is the resistance between the cathode and the anode.
According to Assumption (a), the relationship between the discharge energy q and n is:

q = knλ (2)

where n is the amount of substance of melted glass in unit time and λ is the dissolution heat of
ultra-clear glass.

158
Processes 2019, 7, 195

The volume of melted glass, V, is worked out as

nM
V= , (3)
ρ

where M is the molar mass of the glass and ρ is the density of the glass. Therefore, the diameter, D,
of the machined hole could be calculated together with Equation (3) as:

nM
D=2 , (4)
πhρ

where h is the drilling depth in unit time. The relationship between h and the feed rate v is:

h = v + c, (5)

where v is the feed rate of the rotary helical electrode, c is the distance between the end of the rotary
helical electrode and the bottom of the hole, according to Assumption (c).
The side gap ΔS1 can be defined as follows, where the diameter of the rotary helical electrode is d:

D−d
ΔS1 = . (6)
2
The side gap ΔS1 could be solved simultaneously with Equations (1), (2) and (4)–(6).

M (U I − RI 2 ) d
ΔS1 = − (7)
πρkλ(v + c) 2

We concluded that side gap ΔS1 rose with the increasing of the voltage, but decreased with higher
feed rates. In addition, the side gap ΔS1 was affected by material properties.
Step 2 was the model for electrochemical discharge milling and WECDM. The side gap was
different between electrochemical discharge drilling and milling. The model of the side gap in
electrochemical discharge milling ought to be reconstructed. The side gap model in the electrochemical
discharge milling process is shown in Figure 4.

Figure 4. Schematic view of electrochemical discharge milling.

In unit time, the shape of the machined glass was considered rectangular in electrochemical discharge
milling. Therefore, the volume of the machined glass could be obtained in unit time as follows:

V = (d + 2ΔS2 )vh1 . (8)

In Equation (8), v is the feed rate, h1 is mean milling depth, and d is the diameter of the rotary
helical electrode.
Therefore, the side gap was obtained by Equations (1)–(3) and (8).

159
Processes 2019, 7, 195

M (U I − RI 2 ) d
ΔS2 = − (9)
2ρvh1 kλ 2
It was not hard to establish that the side gap ΔS2 became larger with any increase of voltage in
the electrochemical discharge milling. However, the side gap ΔS2 became narrower with higher feed
rate and higher milling depth. The glass properties also influenced the side gap.
The side gap in WECDM could be substituted, approximately, by Equation (9) from Figure 4,
with an h1 thickness of the glass.

3. Experiments and Discussion

3.1. Experimental Arrangement


In this paper, electrochemical discharge drilling, electrochemical discharge milling and wire
ECDM were employed to investigate the side gap during the processing of ECDM. To ensure the
accuracy of the experiments and to avoid accidental influence, each experiment was carried out
repeatedly, at least three times. In all of the following experiments, Φ105 μm tungsten was used as the
rotary helical tool and a 600 μm thick graphite plate was selected as the auxiliary electrode. Workpieces
in the electrochemical discharge drilling and wire ECDM were ultra-glass with a thickness of 300 μm,
while the specifications of the glass workpiece were 46 mm × 25 mm × 1 mm in electrochemical
discharge milling. In addition, all feed depths were 100 μm in electrochemical discharge milling.
The diameter and slit width were measured by a Nikon SMZ1270 microscope (Tokyo, Japan) and
NOVA NANOSEM 450 scanning electron microscope (Hillsboro, OR, USA).
In these experiments, the auxiliary anode (Luhan metal, Shanghai, China), rotary helical electrode
(Union tool, Tokyo Metropolitan, Janpan) and glass workpiece (Citoglas, Haimen, China) were
immersed in electrolytes. When the pulse voltage was applied to the auxiliary anode and the helical
electrode was attached to high speed spindle, a rotary helical electrode moved with a certain feed
speed to machine the glass. The main discharge areas were the bottom, the side wall, and the side wall
of the rotary helical electrode in the electrochemical discharge drilling, the electrochemical discharge
milling, and the wire ECDM, respectively. Therefore, the selected experimental parameters were
different between the three machining methods. The details of the experimental arrangements are
shown in Table 1. In each group of experiments, only one parameter, the pulse voltage, pulse frequency,
duty factor, or feed rate could be adjusted to the desirable range to research the effect on the side gap.
Other variables were kept constant. The effects of the pulse voltage, frequency, duty factor, and feed
rate on the side gap are displayed in the following table.

Table 1. The details of experimental arrangements.

Item ECD-Drilling ECD-Milling Wire ECDM


Pulse voltage 35–41 (V) 34–40 (V) 32–40 (V)
Frequency 400–700 (Hz) 200–500 (Hz) 200–600 (Hz)
Duty factor 60–90 (%) 50–80 (%) 50–90 (%)
Feed velocity 0.5–2 (μm/s) 0.5–2 (μm/s) 0.5–2.5 (μm/s)
Spindle speed 3000 (rpm)
Concentration 3 M (KOH)

3.2. Effect of Pulse Voltage on Side Gap


There have been many experiments conducted to investigate effects of pulse voltage on the side
gap. The side gap was calculated and the influence of the pulse voltage on the side gap is shown in
Figure 5. From Figure 5 and Equations (7) and (9), we concluded that the side gap increased with
the rise of the pulse voltage. At a lower pulse voltage, the bubbles generated by electrolysis were
sparse and thin. Therefore, the thickness of the gas film was thin. The thin gas film and low applied
voltage led to shorter discharge distances, which greatly shortened the side gap. While at higher pulse

160
Processes 2019, 7, 195

voltages, the formation rate of the bubbles increased rapidly. Plenty of bubbles coalesced intensely,
resulting in a thicker gas film. Thus, in this case, the discharge distance was longer, meaning more
material was removed. It was not difficult to conclude that the side gap increased with the rise of the
discharge distance. The diameter of the hole in the electrochemical discharge drilling, the slit width in
the electrochemical discharge milling and the wire ECDM increased with the higher pulse voltage.

:(&'0
 (&'0LOOLQJ
(&''ULOOLQJ


7KHVLGHJDS μP








     
3XOVHYROWDJH 9

Figure 5. Effect of pulse voltage on side gap.

3.3. Effect of Duty Factor on Side Gap


To research the effect of the duty factor on the side gap, a series of experiments were carried
out, including electrochemical discharge drilling, electrochemical discharge milling and wire ECDM.
The results are shown in Figure 6. As the picture depicts, the side gap increases as the duty factor
rises, from 40% to 90%. The discharge energy q in unit time increased due to the higher duty factor,
which resulted in more material removal. Therefore, the diameter of the hole in electrochemical
discharge drilling, the slit width in electrochemical discharge milling and the wire ECDM increased
with the rise of the duty factor. The optimal duty factor should be low, but the lower duty factors
reduced material removal rate.

(&''ULOOLQJ
 (&'0LOOLQJ
:(&'0


7KHVLGHJDS μP









    
'XW\IDFWRU 

Figure 6. Effect of duty factor on side gap.

3.4. Effect of Frequency on Side Gap


The influence of frequency on the side gap is shown in Figure 7. In this set of experiments, the duty
factor remained unchanged at 70% and frequency ranged from 200 Hz to 600 Hz. In electrochemical
discharge drilling, electrochemical discharge milling, and wire ECDM the side gap decreased when

161
Processes 2019, 7, 195

the frequency increased, gradually. The number of discharge rose with higher frequency per unit
time, but the pulse width was correspondingly reduced. Therefore, the discharge energy of a single
discharge decreased, resulting in less material removal and smaller side gaps, eventually. The diameter
of the hole in electrochemical discharge drilling, the slit width in electrochemical discharge milling
and the wire ECDM decreased with the rise of frequency. The optimal frequency should be high but
the higher frequency will reduce the material removal rate.

:(&'0
 (&'0LOOLQJ
(&''ULOOLQJ


7KHVLGHJDS μP






  
)UHTXHQF\ +]

Figure 7. Effect of frequency on side gap.

3.5. Effect of Feed Rate on Side Gap


Numerous experiments were conducted to research the effect of feed rate on the side gap. Different
feed rates had different influences on the side gap, as displayed in Figure 8. Better machining location
with s higher feed rate could be obtained with a lower side gap. The shorter discharge time with
the higher feed rate in unit machining distance along the direction of feed, led to less material being
removed. Therefore, the side gap was lower in the electrochemical discharge drilling, electrochemical
discharge milling, and wire ECDM. However, the optimal feed rate was not higher. The rotary helical
electrode collided with the workpiece when the feed rate rose to critical values.



:(&'0
 (&''ULOOLQJ
(&'0LOOLQJ


7KHVLGHJDS μP









   
)HHGUDWH μPV

Figure 8. Effect of feed rate on side gap.

162
Processes 2019, 7, 195

4. Experimental Results
According to the above experiments and analysis, the effects of the parameters, including voltage,
frequency, duty factor and feed rate, on the side gap were worked out in ECDM with rotary helical
electrodes. The parameters, after optimization, were selected based on many experiments exploring
fabricated micro holes, micro grooves, micro channels and complicated three-dimensional features
with lower side gaps. There were some micro structures displayed.

4.1. Electrochemical Discharge Drilling of Array Micro Holes


According to the above discussion about the effect of the parameters on the side gap, the smaller
side gaps needed a low voltage, low duty factor, high frequency, and high feed rate. However,
considering material removal rate and machining stability, the experiments were carried out
to select a set of optimized parameters for electrochemical discharge drilling. The optimized
parameters were: pulse voltage—37 V, frequency—3000 Hz, duty factor—70%, feed rate—1 μm/s,
spindle speed—3000 rpm, and electrolyte—3 mol/L KOH. High quality array micro holes were
successfully fabricated with a lower diameter, as shown in Figure 9. Thickness of the glass was
300 μm. A minimum side gap of 27.2 μm could be obtained with electrochemical discharge drilling.

Figure 9. Array micro holes and partial magnification

4.2. Electrochemical Discharge Milling of Micro Structures


Electrochemical discharge milling was capable of fabricating micro grooves, micro channels
and micro three-dimensional structures. Some complex micro structures could be machined with
a lower side gap by a set of optimized parameters. The optimized parameters were: pulse
voltage—34 V, frequency—500 Hz, duty factor—50%, feed rate—2 μm/s, spindle speed—3000 rpm,
and electrolyte—3 mol/L KOH. As shown in Figure 10, the micro groove array was milled on the glass.
The mean width was 129.4 μm, the length was 750 μm and depth was about 130 μm. The smallest side
gap reached 11.5 μm.

163
Processes 2019, 7, 195

Figure 10. Micro groove array on glass.

The micro channel machined on the glass by electrochemical discharge milling is displayed in
Figure 11. The groove width was about 135.9 μm and the depth was about 150 μm. The abbreviation
of the university name milled on the glass is shown in Figure 12. The three-dimensional step structure
with vertical sidewalls and high shape accuracy is shown Figure 13. The three-dimensional convex
structure of micro electrochemical discharge milling is shown in Figure 14, which is made of two layers
of convex structures. The width of the upper convex plate was about 75 μm, the length was 260 μm
and the height was about 70 μm.

Figure 11. Complex micro channel on glass.

Figure 12. The abbreviation of the university name on glass.

164
Processes 2019, 7, 195

Figure 13. Three-dimensional step structure.

Figure 14. Three-dimensional convex structure.

4.3. Wire Electrochemical Discharge of Micro Structures


Wire electrochemical discharge with rotary helical electrodes fabricated high aspect ratio
structures. According to the above discussion and experiments, a set of optimized parameters was
selected for machining the micro structures. Long narrow slits were fabricated on 300 μm thick glass,
as shown in Figure 15. The smallest side gap reached 14.95 μm. The optimized parameters were:
pulse voltage—34 V, frequency—600 Hz, duty factor—50%, and feed rate—1 μm/s. The closed micro
structures were machined as displayed in Figure 16. To improve the refreshment of the electrolyte in
the closed micro structures, larger processing parameters were used (40 V, 500 Hz, 50%, 1 μm/s).

Figure 15. Long narrow slits on 300 μm thick glass workpiece.

165
Processes 2019, 7, 195

Figure 16. Closed micro structure on 300 μm thick glass.

The high aspect ratio structure was manufactured on 1060 μm thick glass with wire ECDM,
using a rotary helical electrode. The slit width was about 175.4 μm and the side gap was about 35.2 μm,
as shown in Figure 17 (40 V, 300 Hz, 60%, 1 μm/s). In addition, the micro cantilever beam was
successfully fabricated on a 35 μm thick glass workpiece, as shown in Figure 18. The length of the
micro cantilever beam was 1500 μm and the aspect ratio reached 42:1.

Figure 17. High aspect ratio structure on 1060 μm thick glass.

Figure 18. Micro cantilever beam on 35 μm thick glass workpiece.

5. Conclusions
This research employed ECDM with a rotary helical electrode to fabricate ultra-clear glass.
Using a rotary helical tool in electrochemical discharge drilling, electrochemical discharge milling,
and wire ECDM, the effects of pulse voltage, frequency, duty factor, and feed rate on the side gap were
investigated. The conclusions can be summarized as follows:

166
Processes 2019, 7, 195

(1) The mathematical model for the ECDM process was established to guide the machining of
microstructures on ultra-clear glass.
(2) The side gap increased with the increase in voltage and duty factor and was reduced with a higher
frequency and feed rate in a certain range.
(3) By employing optimized parameters in ECDM, micro holes, micro channels, micro slits and
complicated three-dimensional features with ten several-micron side gaps were successfully
fabricated on ultra-clear glass.

Author Contributions: Conceptualization, Y.L. and C.Z.; investigation, S.L. and Z.W.; methodology, C.G.; project
administration, Y.L.; resources, Z.W.; writing—original draft, S.L.; and writing—review and editing, Y.L. and C.G.
Acknowledgments: The authors acknowledge financial support from the Shandong Provincial Natural
Science Foundation (Nos. ZR2018MEE018, ZR2017BEE012), the China Postdoctoral Science Foundation (No.
2018M630772), the Young Scholars Program of Shandong University, Weihai (2015WHWLJH03), and the Shenzhen
Science and Technology Project (JCYJ20170818103826176).
Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Allesu, K.; Ghosh, A.; Wuju, M.K. A preliminary qualitative approach of a proposed mechanism of material
removal in electrical machining of glass. Eur. J. Mech. Eng. 1991, 36, 201–207.
2. Zheng, Z.P.; Cheng, W.H.; Huang, F.Y.; Yan, B. 3D microstructuring of Pyrex glass using the electrochemical
discharge machining process. J. Micromech. Microeng. 2007, 17, 960–966. [CrossRef]
3. Díaz-Tena, E.; Gallastegui, E.; Hipperdinger, M.; Donati, E.R.; Ramírez, M. New advances in copper
biomachining by iron-oxidizing bacteria. Corros. Sci. 2016, 112, 385–392. [CrossRef]
4. Paredes-Sanchez, J.P.; Lopez-Ochoa, L.M.; Lopez-Gonzalez, L.M.; Las-Heras-Casas, J.; Xiberta-Bernat, J.
Evolution and perspectives of the bioenergy applications in Spain. J. Clean. Prod. 2019, 213, 553–568.
[CrossRef]
5. Sanchez, J.A.; Plaza, S.; De Lacalle, L.N.; Lamikiz, A. Computer simulation of wire-EDM taper-cutting. Int. J.
Comput. Integr. Manuf. 2006, 19, 727–735. [CrossRef]
6. Nasim, S.; Mohammad, R.R.; Mansour, H. Experimental investigation of surfactant-mixed electrolyte into
electro chemical discharge machining (ECDM) process. J. Mater. Process. Technol. 2017, 250, 190–202.
7. Zhang, Z.Y.; Huang, L.; Jiang, Y.J.; Liu, G.; Nie, X.; Lu, H.Q.; Zhuang, H.W. A study to explore the properties
of electrochemical discharge effect based on pulse power supply. Int. J. Adv. Manuf. Technol. 2016, 85,
2107–2114. [CrossRef]
8. Sathisha, N.; Somashekhar, S.H.; Shivakumar, J. Prediction of material removal Rrate using regression
analysis and artificial neural network of ECDM process. Int. J. Recent Adv. Mech. Eng. 2014, 3, 69–81.
9. Jawalkar, C.S.; Sharma, A.P.; Kumar, P. Micromachining with ECDM: Research Potentials and Experimental
Investigations. Channels 2012, 6, 340–345.
10. Cao, X.D.; Kim, B.H.; Chu, C.N. Hybrid Micromachining of Glass Using ECDM and Micro Grinding. Int. J.
Precis. Eng. Manuf. 2016, 14, 5–10. [CrossRef]
11. Elhami, S.; Razfar, M.R. Effect of ultrasonic vibration on the single discharge of electrochemical discharge
machining. Mater. Manuf. Process. 2017, 33, 444–451. [CrossRef]
12. Elhami, S.; Razfar, M.R. Analytical and experimental study on the integration of ultrasonicallyvibrated tool
into the micro electro-chemical discharge drilling. Precis. Eng. 2017, 47, 424–433. [CrossRef]
13. Singh, T.; Dvivedi, A. Developments in electrochemical discharge machining: A review on electrochemical
discharge machining, process variants and their hybrid methods. Int. J. Mach. Tools Manuf. 2016, 105.
[CrossRef]
14. Elhami, S.; Razfar, M.R. Study of the current signal and material removal during ultrasonic-assisted electro
chemical discharge machining. Int. J. Adv. Manuf. Technol. 2017, 92, 1591–1599. [CrossRef]
15. Han, M.S.; Min, B.K.; Lee, S.J. Modeling gas film formation in electrochemical discharge machining processes
using a side-insulated electrode. J. Micromech. Microeng. 2008, 18, 19–26. [CrossRef]
16. Furutani, K.; Maeda, H. Machining a glass rod with a lathe-type electro-chemical discharge machine.
J. Micromech. Microeng. 2008, 18, 6–13. [CrossRef]

167
Processes 2019, 7, 195

17. Kun, L.W.; Hsin, M.L.; Kuan, H.C. Application of Electrochemical Discharge Machining to Micro-Machining
of Quartz. Adv. Mater. Res. 2004, 939, 161–168.
18. Tsuchiya, H.; Inoue, T.; Miyazaiki, M. Wire electro-chemical discharge machining of glasses and ceramics.
Bull. Jpn. Soc. Precis. Eng. 1985, 19, 73–74.
19. Jain, V.K.; Rao, P.S.; Choudhary, S.K.; Rajurkar, K.P. Experimental Investigations into Traveling Wire
Electrochemical Spark Machining (TW-ECSM) of Composites. J. Eng. Ind. 1991, 113, 75–84. [CrossRef]
20. Panda, M.C.; Yadava, V. Finite element prediction of material removal rate due to traveling wire
electrochemical spark machining. Int. J. Adv. Manuf. Technol. 2009, 45, 506–520. [CrossRef]
21. Kuo, K.Y.; Wu, K.L.; Yang, C.K.; Yan, B.H. Wire electrochemical discharge machining (WECDM) of quartz
glass with titrated electrolyte flow. Int. J. Mach. Tools Manuf. 2013, 72, 50–57. [CrossRef]
22. Wang, J.; Guo, Y.B.; Fu, C.; Jia, Z.X. Surface integrity of alumina machined by electrochemical discharge
assisted diamond wire sawing. J. Manuf. Process. 2018, 31, 96–102. [CrossRef]
23. Rattan, N.; Mulik, R.S. Improvement in material removal rate (MRR) using magnetic field in TW-ECSM
process. Mater. Manuf. Process. 2017, 32, 101–107. [CrossRef]
24. Huang, S.F.; Liu, Y.; Li, J.; Hu, H.X.; Sun, L.Y. Electrochemical Discharge Machining Micro-Hole in Stainless
Steel with Tool Electrode High-Speed Rotating. Mater. Manuf. Process. 2014, 29, 634–637. [CrossRef]
25. Han, M.S.; Min, B.K.; Lee, S.J. Geometric improvement of electrochemical discharge micro-drilling using
an ultrasonic-vibrated electrolyte. J. Micromech. Microeng. 2009, 19, 65004. [CrossRef]
26. Fang, X.L.; Zhang, P.F.; Zeng, Y.B.; Qu, N.S.; Zhu, D. Enhancement of performance of wire electrochemical
micromachining using a rotary helical electrode. J. Mater. Process. Technol. 2016, 227, 129–137.
27. Wang, M.Y.; Zhang, J.H.; Liu, Y.; Li, M.H. Investigation of Micro Electrochemical Discharge Machining Tool
with High Efficiency. Recent Pat. Eng. 2016, 10, 146–153. [CrossRef]
28. Jain, V.K.; Dixit, P.M.; Pandey, P.M. On the analysis of the electrochemical spark machining process. Int. J.
Mach. Tools Manuf. 1999, 39, 165–186. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

168
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Processes Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/processes
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel: +41 61 683 77 34
Fax: +41 61 302 89 18
www.mdpi.com ISBN 978-3-03936-387-2

You might also like