Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Applied Energy 279 (2020) 115777

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Efficient generation of hydrogen by two-step thermochemical cycles:


Successive thermal reduction and water splitting reactions using
equal-power microwave irradiation and a high entropy material
Yibo Gao , Yanpeng Mao *, Zhanlong Song , Xiqiang Zhao , Jing Sun , Wenlong Wang ,
Guifang Chen , Shouyan Chen
National Engineering Laboratory for Reducing Emissions from Coal Combustion, Engineering Research Center of Environmental Thermal Technology of Ministry of
Education, Shandong Key Laboratory of Energy Carbon Reduction and Resource Utilization, School of Energy and Power Engineering, Shandong University, Jinan,
Shandong 250061, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Successive thermal reduction and water


splitting reactions are performed using
equal-power microwave irradiation.
• The oxygen vacancy of the high entropy
material could be significantly increased
by short-term (4 min) microwave
irradiation.
• The maximum hydrogen yield is 122
mL/g/cycle at 700 W, exceeding the
thermodynamic limits of state-of-the-art
materials.
• Microwave system could reduce power
consumption by 97%, compared with
conventional thermal process.

A R T I C L E I N F O A B S T R A C T

Keywords: Hydrogen production through two-step thermochemical water splitting cycle based on metal oxide has emerged
Hydrogen as a promising strategy to store dilute and intermittent solar energy. However, a typical reaction time of the two-
Two-step water splitting step thermochemical water splitting cycle is lengthy, with at least 0.5 h for thermal reduction step and 1 h for
Microwave
water splitting step, and the energy required in the thermal reduction process for hydrogen regeneration is higher
SiC foam
High entropy oxide
than the generated hydrogen energy. In this work, to overcome the problem of an energy efficiency imbalance,
we investigated the possibility of the rapid and successive reactions of thermal reduction and water splitting,
using short-term irradiation from a low-energy microwave. To this end, a high entropy material, as a poly-metal
oxide used to generate hydrogen, was produced by simultaneously introducing four cations onto a SiC foam –
(FeMgCoNi)O1.2@SiC. It was found that the oxygen vacancy of the (FeMgCoNi)O1.2@SiC could be significantly
increased by short-term microwave irradiation, and hence the thermal reduction process took only 4 min, which
is much less than normal. High H2 generation rates were achieved by re-oxidation of the Fe (II) to Fe (III) of the
(FeMgCoNi)O1.2@SiC. In addition, the microwave plasma generated by microwave irradiation induced (FeMg­
CoNi)O1.2@SiC discharge could enhance the water splitting process. The maximum hydrogen yield was 122 mL/
g at 700 W, due to the coupling effect of the thermochemical cycle and microwave plasma. In this way, the power

* Corresponding author.
E-mail address: maoyanpeng@sdu.edu.cn (Y. Mao).

https://doi.org/10.1016/j.apenergy.2020.115777
Received 9 June 2020; Received in revised form 13 August 2020; Accepted 21 August 2020
Available online 1 September 2020
0306-2619/© 2020 Elsevier Ltd. All rights reserved.
Y. Gao et al. Applied Energy 279 (2020) 115777

consumption of microwave process is only 3% of that of conventional high-temperature heat treatment during
thermal reduction process.

Nomenclature ω angular frequency

Vi volumetric rate (mL/min/g) Acronyms


xi O2 or H2 mole fraction TWS thermochemical water splitting
Vtotal total volumetric flow rate (mL/min) TR thermal reduction
m the mass of the active material (g) WS water splitting
Yi O2 or H2 yield (mL/g) HTGR high temperature gas-cooled reactor
t reaction time (min) DFT density functional theory
Weletric energy input (J) HEO high entropy oxide
ε* complex dielectric constant HEOS high entropy oxide supported by a SiC foam
ε′ the real component of the complex dielectric constant HHV higher heating value
ε′′ the imaginary component of the complex dielectric MFCs mass flow controllers
constant XRD X-ray diffractometer
δ loss tangent XPS X-ray photoelectron spectroscopy
f microwave frequency SEM scanning electron microscopy
P the ability to convert microwave energy into heat GC gas chromatograph
ε0 the permittivity in a vacuum or air Greek symbol
σ conductivity ƞ energy efficiency
C specific heat capacity

δ
MOx →MOx− δ + O2 (1)
2

is combined with either the water splitting (WS) step


1. Introduction
MOx− δ + δH2 O→MOx + δH2 (2)
Since the industrial revolution, the amount of carbon emissions from
the burning of fossil fuels has created significant climate-change risks or the CO2 splitting step
[1]. Simultaneously, it brings opportunities and challenges to the field of MOx− δ + δCO2 →MOx + δCO (3)
sustainable energy development. Solar energy is the most abundant
energy source and the largest source of renewable energy [2], and The TR step is an endothermic reaction in which a higher-valence
hydrogen production from solar energy has considered as a potential metal oxide is dissociated into a lower-valence metal oxide (or metal)
way of storing solar energy [3]. At present, many hydrogen production and oxygen, at a high temperature (TTR) provided by a solar concen­
methods have been reported, for example, solar-driven electrochemical trator [14]. In the WS step (non-solar exothermic step), the lower-
[4] and photoelectrochemical [5] pathways for the direct conversion of valence metal oxide (or metal) reacts with water at a low temperature
water to hydrogen and oxygen, and solar thermochemical processes for (TWS ≤ 1000 ◦ C) to form hydrogen and a higher-valence metal oxide,
the indirect conversion of CO2 and H2O to fuel [6]. All of these are which is subsequently recycled in the first step, ideally for indefinite
promising technologies for hydrogen production with minimal global reuse [15]. A schematic diagram of two-step TWS cycle is illustrated in
warming potential in the long term [7]. In this field, thermochemical
water splitting (TWS) has attracted a lot of attention because of it has the
potential to achieve large-scale industrialization [8]. One method to
combine multiple-step (two or more) chemical reactions to split water is
the so-called thermochemical cycles with the net reaction being water
decomposition [9]. Thermochemical cycles can be divided into two
main categories, which are two-step cycle of metal oxides and the multi-
step cycle such as the Iodine-Sulfur (three-step) [10] or Iron-chlorine
(four-step) cycles [11]. The direct thermolysis of water is the simplest
reaction to split water. However, water can be directly decomposed into
hydrogen and oxygen only when the temperature exceeds 2500 K,
because of its unfavourable thermodynamics [12]. In most cases, the
required temperature for water splitting decreases when more steps are
employed, but so does the efficiency potential due to the energy losses
associated with heat transfer and products separation in each step [13].
Therefore, solar-driven two-step TWS based on metal oxides has
emerged as an attractive technology for solar fuel production due to its
relatively simplicity and operation convenience. Such cycles commonly
operate in two separate reduction/oxidation steps utilizing a metal
oxide as a reactive intermediate, described by reactions (1)–(3).
The thermal reduction (TR) step Fig. 1. Schematic of two-step thermochemical water splitting cycle.

2
Y. Gao et al. Applied Energy 279 (2020) 115777

Fig. 1. Although the temperature of the TR step is lower than the direct needed to maintain constant temperatures and hydrogen yield [37].
thermolysis of water, it is still relatively high, ranging from 1600 to However, it is still necessary to development a novel hydrogen regen­
1900 K depending on the type of redox material employed [16]. Thus, eration method that can produce hydrogen repeatedly with a much
two-step TWS are not suitable for coupling to high temperature gas- lower energy.
cooled reactor (HTGR), because the optimal temperature level for In this work, we studied the possibility of the rapid and continuous
HTGR is expected to be around 1100–1250 K, which is suitable for regeneration of hydrogen, using short-term irradiation from a low-
combination with multi-step thermochemical cycles [17]. energy microwave. The reported high-entropy oxide (HEO), (FeMg­
The solar-to-fuel energy conversion efficiency of a two-step TWS, CoNi)Ox, was used as the metal oxide for thermochemical hydrogen
driven by concentrating solar power, is expected to reach 50% because it production. We also investigated how the SiC foam used to support
utilises the entire solar spectrum and does not involve a power genera­ (FeMgCoNi)Ox ((FeMgCoNi)Ox@SiC, hereafter referred to as HEOS)
tion step [18]. It could therefore lead to a future of sustainable and promotes the water splitting cycle during equal-power microwave
secure energy for the world. The two-step TWS cycle must go through irradiation. The hydrogen productivity and overall reaction kinetics of
several stages of development before it can be commercialised, and the HEOS were investigated under different experimental conditions. Also,
first stage is to develop and test suitable metal oxides. The thermo- the energy efficiency of the system was calculated and compared to the
kinetic properties of the metal oxide dictate the viability and effi­ conventional thermal process.
ciency of the two-step TWS process. An optimum metal oxide has not yet
been identified, although active searches have been undertaken. The 2. Experimental section
vast majority of simple binary metal oxides, such as ZnO2/Zn cycle,
Er2O3/ErO cycle [19] and Sm2O3/Sm cycle [20], have been examined 2.1. Preparation of HEOS
through HSC chemistry software and databases under the assumption
that they operate via volatile or non-volatile stoichiometric mechanisms (FeMgCoNi)Ox was synthesised using sol-gel techniques. All the re­
[21]. While the reaction mechanism and the water splitting capability of agents in these experiments were used without further purification.
many complex materials, such as CeO2-δ, doped-ferrite, and ABO3-δ pe­ Portions of iron(III) nitrate nonahydrate (≥98%), magnesium nitrate
rovskites, have been predicted through density functional theory (DFT) hexahydrate (99%), cobalt(II) nitrate hexahydrate (≥98%), and nickel
and experimentally studied [22]. Many studies confirmed that non- (II) nitrate hexahydrate (crystalline), with equimolar cations, were
volatile and non-stoichiometric oxides, such as substituted ceria [23], mixed and dissolved in deionised water with a precursors-to-water mass
spinel ferrites [24], and perovskites [25], had favourable thermo-kinetic ratio of 1:4. Subsequently, citric acid (≥99.5%) and ethyl­
properties for water splitting, moderate reduction temperatures, and enediaminetetraacetic acid (EDTA) were added to the suspension, which
better structural stability than stoichiometric oxides (i.e., ZnO2 and accounted for 75% and 60% of the total molar amount of metal ions,
SnO2) [26]. The most advanced materials, such as CeZrO2-δ and spinel respectively. Sodium hydroxide solution (1 mol/L) was added dropwise
ferrites (specifically NiFe2O4), have good hydrolysis ability at to the suspension with vigorous stirring at room temperature until the
1400–1500 ◦ C [27]. However, given that current large-scale chemical pH reached 11, at which time a dark solution was formed. The resulting
infrastructure relies almost entirely on thermochemical processes at solution was stirred and heated on a hot plate at 100 ◦ C for about 6 h
1100 ◦ C [28], the temperature required for the TR step is an inevitable (gelation), and the cleaned and dried SiC foam was then placed into the
challenge for achieving two-step TWS at an industrial scale. Particularly, generated gel. After being fully coated with gel, the SiC foam was
research on lowering the temperature of TR step is essential for transferred to a beaker and dried overnight at 150 ◦ C overnight. Finally,
increasing the solar-to-fuel energy conversion efficiency. In recent years, the coated and dried SiC foam was moved into a quartz boat, and then
researchers have reported some new materials with a lower temperature calcined in a microwave oven (2.45 GHz) at a power of 900 W for 30 min
of TR step. Ruan et al. [29] proposed a novel CeO2-xSnO2/Ce2Sn2O7 under an air atmosphere (microwave calcination). A novel HEOS device
system, which has a peak O2 evolution rate (2.83 mL/min/g) at 1100 ◦ C. was prepared by supporting HEO powder on the SiC foam with a target
Seo et al. [30] used microwave irradiation to control the oxygen va­ active material loading of approximately 25 wt% (weight percent). For
cancies of metal oxides and achieved continuous regeneration of comparison, the remaining gel was dried overnight at 200 ◦ C overnight,
hydrogen (~750 μmol/g). Shang et al. [28] reported poly-cation oxides, during which time foaming occurred and the gel became solid. The
(FeMgCoNi)Ox (x ≈ 1.2), which consist of four cations. After a cycle at sample was ground into a powder, transferred to a quartz boat, and
1300/800 ◦ C, the hydrogen yield of (FeMgCoNi)Ox was 10.1 ± 0.5 mL/ subjected to the following calcination in air: heating from room tem­
g. Debora et al. [31] discussed the unique water splitting properties of perature to 800 ◦ C at 10 ◦ C/min, and then heating to 1000 ◦ C at 5 ◦ C/
BaCe0.25Mn0.75O3, which has a polymorphic phase transition in the min and maintained for 1 h, and finally, natural cooling (conventional
process of thermal reduction. After a short-term cycle at 1350/850 ◦ C, calcination). In order to explore the influence of microwave intensity on
the hydrogen yield (140 μmol/g) was nearly 3 times higher than ceria. the cycle of thermochemical water splitting, we prepared four samples
Chen et al. [32] studied Sr- and Co-doped LaGaO3-δ perovskite and found (A, B, C, and D) by microwave calcination, which were used to split
that the hydrogen yield of LaGa0.4Co0.6O3-δ was 478 μmol/g at 1350/ water under 300 W, 500 W, 700 W, and 900 W, respectively. The sample
850 ◦ C, which was 15 times that of CeO2-δ (32 μmol/g). In addition, on- parameters are shown in Table 1.
sun thermochemical hydrogen production, at temperatures of
1400–1500 ◦ C, has been verified on a pilot scale. Cho et al. [33] 2.2. Characterization of samples
demonstrated a procedure for hydrogen production that uses CeO2/
MPSZ (MgO-partially stabilised zirconia) foam in a 40-kW reactor, with The crystal phase transformations of HEOS samples were charac­
a temperature for the TR step of 1500–1600 ◦ C. Hoskins et al. [34] used terised via physical and chemical techniques at various stages: as-
a high-flux solar furnace to conduct pilot-scale continuous isothermal synthesised, after reduction, after H2O splitting, and after multicycle
thermochemical hydrogen production. These successful pilot-scale reactive tests. The powder X-ray diffractometer (XRD) on Rigaku Cor­
demonstrations provided important references for further expansion of poration D/max 2500 was used to characterise the structure and
this technique. At the same time, many problems have been encountered composition of the samples, with a scanning scope (2 theta) ranging
during on-sun operations, such as dramatic changes to the reactor from 10◦ to 90◦ , and Cu and Ka radiation generated at 40 mA and 40 kV.
temperature during cloudy weather [35], and the actual efficiency The HEOS oxidation and reduction states during the reaction were
(5.25%) was much lower than the theoretical value [36]. Therefore, characterised using X-ray photoelectron spectroscopy (XPS), which was
strong reactor materials are required to cope with thermal stress performed with a ThermoFischer ESCALAB 250Xi with Al Kα radiation
generated by temperature swings and advanced control systems are (hν = 1486.6 eV), calibrated using the C1s peak at 284.8 eV, with an

3
Y. Gao et al. Applied Energy 279 (2020) 115777

Table 1
HEOS device specifications.
Sample Diameter (mm) Thickness (mm) Cell density (ppi) Porosity (%) Weight of sample (g) Weight of loaded HEO (g) HEO loading (wt%)

A 35 15 10 90 8.361 2.260 27.03


B 35 15 10 90 8.742 2.236 25.58
C 35 15 10 90 9.260 2.580 27.86
D 35 15 10 90 9.151 2.360 25.79

accuracy of 0.1 eV. Finally, Scanning Electron Microscopy (SEM) was gas to an N2/H2O gas mixture, which was then introduced into the
performed on a Carl Zeiss AG/SUPRA55TM scanning electron reactor for the WS step. The N2/H2O gas mixture was produced by
microscope. passing N2 gas through distilled water, at a flow rate of 300 mL/min and
a temperature of 80 ◦ C. The partial pressure of steam in the N2/H2O
2.3. Water splitting test mixture was estimated to be 28% (84 mL/min) by considering the water
vapour pressure at 80 ◦ C and 1 bar. The passage for the mixing gas to the
The two-step TWS experiments were performed using a 2.45 GHz reactor was heated by a heater band (250 ◦ C) to prevent condensation.
commercial microwave oven connected to a gas chromatograph (GC, Furthermore, the water vapor that was not decomposed into hydrogen
PerkinElmer Scientific/CLARUS 500), as shown in Fig. 2. A reactor was was separated into a cold trap. The amount of product gases (O2, H2)
made of heat-resistant quartz glass to avoid any loss of microwave en­ generated per minute was measured using a GC. A correction curve for
ergy. A HEOS sample containing approximately 2 g of active material accurate hydrogen generation was obtained using three mixes of stan­
(HEO) was placed in the quartz reactor, and the reactor was closed using dard N2-H2 gases with hydrogen concentrations of 0.101% (mol/mol),
a quartz flange and polytetrafluoroethylene bolts. Inlet gas was passed 1.01% (mol/mol), and 2.02% (mol/mol). The hydrogen productivity
upward through a tray with sieve pores, allowing the gas to make full and reaction kinetics of the materials were investigated under identical
contact with the sample, and then exhausted at the top of the reactor. experimental conditions. At the end of each processing step, the redox
During this thermochemical cycle, the gas flow was regulated with material was cooled in a N2 atmosphere. The correction curve for oxygen
digital mass flow controllers (MFCs). Both the TR and WS steps were generation was obtained with two known concentrations of O2 (2.02%
carried out at the same microwave power and atmospheric pressure. The (mol/mol), 5.05% (mol/mol), balance N2). The transient O2 and H2
TR step was first accomplished by microwave irradiation in a N2 stream evolution rates were calculated using the mass of the loaded active
(99.999%, 300 mL/min). Specifically, the HEOS samples were treated material and Eq. (4):
for 3–4 min at four microwave intensities ranging from 300 to 900 W.
Subsequently, the gas passing through the reactor was switched from N2

Fig. 2. Experimental setup for two-step TWS reaction and discharge light-intensity detection.

4
Y. Gao et al. Applied Energy 279 (2020) 115777

Vi =
xi Vtotal
(4) (a) Conventional calcination
m Microwave calcination
Fe2O3, PDF#33-0664
where Vi (mL/min/g) represents the volumetric rates of O2 or H2 pro­ Fe3O4, PDF#26-1136

Intensity (arb.units)
duced per unit of mass of active material, xi denotes the mole fraction of
O2 or H2 in the produced gas, monitored by the GC, Vtotal (mL/min) is the
total volumetric flow rate regulated by the MFCs, and m (g) is the mass of
the HEO active material. The oxygen and hydrogen yield per cycle was
obtained by numerical integration of the gas evolution rate–time curves,
as shown in Eq. (5):
∫ t2
Yi = Vi dt (5)
t1

where Yi is defined as the yield of O2 or H2 (mL/g), and t (min) is the


reaction time. 20 40 60 80

To investigate the effect of discharge intensity and frequency on 2 theta (degree)


water splitting process, an optical fiber was inserted into the microwave
oven and used to capture the light intensity by using a photoresistance
(b) Conventional calcination
outside the microwave oven. The photoresistance was connected in se­
ries with a DC power supply (12 V) and a resistor (470 kΩ) to detect the
light intensity of the HEOS discharge. The light intensity was recorded at
300–900 W (in 200-W increments) microwave power with a rate of 10
counts per second as a V-t graph using an oscilloscope (Tektronix
TBS1104). The discharge intensity and frequency of HEOS could be
directly exhibited by the variation of voltage.
Microwave calcination
3. Results and discussion

3.1. Physicochemical characterization

Fig. 3a shows the XRD results for the samples calcined by micro­
waves and muffle furnace. Phase identification was conducted using the
MDI Jade software, with the Powder Diffraction Files PDF#33-06644
and PDF#26-1136. The materials obtained by the two calcination Fig. 3. (a) XRD patterns of HEO as-synthesized. (b) SEM images for HEO as-
methods were basically the same, but the individual diffraction peaks synthesized. The XRD patterns of Fe2O3 (PDF#33-06644) and Fe3O4
were different which depended on the preparation method and the (PDF#26-1136) are offered as references.
calcination time. The characteristic lines of Fe2O3 or Fe3O4, as expected,
were hardly detected in the samples, suggesting that Fe ions formed a 3.2. Thermochemical water splitting performance
solid solution with other metal cations. HEO is an equimolar entropy
stabilised oxide [38]. The anionic sub-lattice of HEO can deform to Microwave heating heats the total volume of the reactants by
accommodate differences in cation size and bonding, thereby reducing selectively transferring energy to the microwave-absorbing material,
its enthalpy change [39,40]. The entropy of the HEO controls the tran­ thereby allowing both rapid heating rates and high temperatures to be
sition from multiphase states to a single-phase state: the transition is reached, which in turn drive chemical reactions. In the conventional
endothermic and reversible with temperature [41]. However, Fe, Co, heating process, thermal energy is transferred to the reactant particles
Mg, and Ni have very similar atomic scattering factors and, when mixed, through convection, conduction and radiation of external heat from the
they are difficult to be distinguished with XRD. Shang et al. [28] used outer surfaces of a reactor into the host material itself. In contrast, mi­
magnetite (Fe3O4) and FetO (t ≤ 1) to simulate the spinel and rocksalt crowave energy is delivered directly to the microwave-absorbing
phases of quenched (FeMgCoNi)Ox, respectively, and established the component, such as SiC foam, through molecular interactions. Heat
two-phase models. The value of x can be determined (x ≈ 1.2), assuming transfer by conventional heating is carried out by time-dependent,
that the valence of Fe is +3, Mg and Ni are +2, and Co is a mixture of +2 developing thermal gradients. However, in microwave heating,
and +3 (the same as in Co3O4), from the XPS results discussed later. In through electromagnetic coupling of various charge-dynamical pro­
addition, the material obtained by microwave calcination was looser cesses, electromagnetic energy is directly transferred to the microwave
and had a finer and more uniform particle size, which allowed us to absorbing substance within the sample, and heat is instead generated
produce particles with a small grain size of ~1 nm (Fig. 3b). It was well [43]. This fundamental difference from traditional heating results in
known that the microwave heating effect, or dielectric heating, occurs important advantages, as microwaves can be used to initiate and pro­
when the molecular dipole moments generate vibration and rotational mote the reaction of two-step water splitting by microwave-absorbing
motion under the induction of high-frequency electromagnetic radiation materials or metal oxides. First, this work studied the irradiation time
to achieve rapid heating. The selective heating character of microwaves required for activation of the HEOS in the TR step at microwave in­
offers a shorter reaction time without causing sharp thermal gradients. tensities of 300–900 W (in 200-W increments), using oxygen generation
The conventional calcination method can cause a sharp thermal gradient as an evaluation standard. The reaction temperature was measured at
in the reaction system, which makes the growth process inhomogeneous the reactor wall in the microwave oven by using an infrared thermom­
and consequently broadens the size distribution (Fig. 3b) [42]. eter (Fig. 4a). In the microwave oven, interpretation of heating rates is
complicated by the fact that the temperature is measured at the reactor
wall and is not a direct reflection of the microscopic reaction tempera­
ture. The heating rate (dT/dt) correlates with the real (ε′ ) and imaginary

5
Y. Gao et al. Applied Energy 279 (2020) 115777

(c) 12 (d) 160


900 W H2
Microwave power for WS step
700 W 120.7
10 500 W
140 O2
H2 generation rate (mL/min/g)

300 W
120
8
Yield (mL/g) 100
6
80
67.78
4
60

2 40 27.31
8.4
20
0
26.82 8.01 30.01 14.01
0
0 5 10 15 20 25 30 300 500 700 900
Time (min) Microwave power (W)

Fig. 4. (a) Temperature (◦ C) and power (W) profiles of the TR step. (b) Change in oxygen generation rate (mL/min/g) with irradiation time (min) during the TR step
with microwave intensities of 300, 500, 700, and 900 W, respectively. (c) Hydrogen generation rate (mL/min/g) with varying microwave power. (d) Hydrogen yield
(mL/g) with varying microwave power. The experimental conditions were as follows: N2 sweep gas flow rate = 300 mL/min, thermal reduction was performed for 4
min, and water splitting was performed for ~30 min.

(ε′′ ) components of the complex dielectric constant (ε* = ε’ − iε’’) for permittivity.
individual components in reaction mixtures [44]. The ability to convert Fig. 4b shows the amount of generated oxygen according to the
microwave energy into heat (P) is represented by the loss tangent microwave intensity and irradiation time. Intriguingly, the O2 genera­
(tanδ = ε ′ /ε ), which is related to the microwave frequency (f), tem­
′ ′
tion rate shows an initial rapid increase, attaining a peak rate after an
perature, state, and composition of the material. Commonly, the ability irradiation time of 1 min. The O2 generation rate decreased rapidly after
to convert microwave energy into heat (P) is directly proportional to the 4 min of irradiation. Based on this finding, during subsequent tests, the
imaginary part (ε′′ ) of the dielectric constant of the reactant [42], as TR step was performed under microwave irradiation for 4 min, which
shown in Eq. (6): was longer than that reported by Seo et al. [30] (3 min) because of the
different redox materials employed. For thermochemical cycling appli­
P = 2πf ε0 ε’’E2 (6) cations, we compared the transient O2 and H2 generation rates during a
complete water splitting redox cycle, obtained with HEOS. Fig. 4c shows
where E is an applied electric field and ε0 is the permittivity in a vacuum
a comparison of the hydrogen production rates during tests conducted
or air. Furthermore, the rate of heating of the reactants can be described
with different microwave irradiation power conditions (300–900 W, in
by Eq. (7):
200-W increments, for 30 min). During the WS step, the introduction of
dT/dt = σE2 /ρC (7) steam led to immediate production of H2, in all cases. As the microwave
power increased, the amount of generation hydrogen also increased. At
where ρ is the density, C is the specific heat capacity of the reactant, 300 W, the hydrogen production was only 8.40 mL/g, as shown in
and the conductivity (σ ) is positively related to the imaginary part of the Fig. 4d, which was much lower than the oxygen production (26.82 mL/
permittivity, ε∗ = ε′ - iσ /ω (where ω is the angular frequency) [45]. In g). It is worth noting that the temperature required for the WS step is
general, the value of E is dependent on the design of the cavity, and is usually 800 ◦ C, which means that a 300 W microwave power is not
therefore not calculable; however, this expression shows that under applicable to the water splitting process (420 ◦ C). However, at a mi­
microwave radiation, it is possible to rapidly attain a temperature crowave power of 700 W (590 ◦ C), it exhibited an initial rapid increase
capable of activating the reactant by using a reactant with a higher

6
Y. Gao et al. Applied Energy 279 (2020) 115777

to a peak rate of 10.22 mL/min/g, followed by a slower decrease rate, an exponential decline (Fig. 5c). Conversely, at 500 W and 700 W, HEOS
and the total amount of generated hydrogen was 120.70 mL/g, which did not achieve thermodynamic equilibrium, as indicated by the long
was much higher than that of the other microwave power levels. Simi­ tail of H2 generation after re-oxidation. This was due to the comparably
larly, at a microwave power of 500 W, the H2 generation rates of HEOS short re-oxidation period, without sufficient oxidation of HEOS (Fig. 5a
exhibited an initial rapid increase to a peak rate of 2.86 mL/min/g, and 5b), which was further confirmed by the followed XPS character­
followed by a comparably slower decrease rate. In contrast, at 900 W ization (Fig. 9b). As seen in Fig. 5d, the average H2 productivity, after six
(640 ◦ C), the peak H2 generation rate of HEOS was 2.88 mL/min/g, consecutive redox cycles, was 40.56 mL/g for Sample B at 500 W, 63.47
lower than that of 700 W. To assess the stability of the water-splitting mL/g for Sample C at 700 W, and 18.74 mL/g for sample D at 900 W.
performance, HEOS was subjected to six consecutive redox cycles at From the view of kinetics, a proper increase of reaction temperature is
microwave intensities of 500–900 W. At 900 W, the H2 generation rates benefit for WS step. Meanwhile, from the point of thermodynamics, the
of HEOS exhibited an initial rapid increase to the peak rate, followed by WS step is an exothermic reaction, higher reaction temperature is

Fig. 5. (a) The transient O2 and H2 generation rates of Sample B for 500 W. (b) The transient O2 and H2 generation rates of Sample C for 700 W. (c) The transient O2
and H2 generation rates of Sample D for 900 W. (d) O2 and H2 yield of HEOS over six consecutive H2O splitting cycles.

7
Y. Gao et al. Applied Energy 279 (2020) 115777

detrimental to its thermodynamic driving force [46]. If the WS step is vapor to produce OH⋅ in a strong discharge environment (Eq. (10)).
conducted at 300 W, the H2 generation rate is too slow to be acceptable. However, in the 6th cycle, the hydrogen production of all samples
While higher reaction temperature (e.g. microwave power ≥ 900 W) is was less than the previous 5 cycles. After 6 cycles, Sample C, with a size
thermodynamically unfavourable for the H2 generation. of 200–500 nm (Fig. 7c), was obviously sintered, when compared with
Notably, during oxidation-reduction reactions, HEOS produced a the as-synthesised samples (Fig. 7a) and the sample after 3 cycles of
strong discharge phenomenon, especially after three cycles. As shown in reactive tests (Fig. 7b). Furthermore, the SiC foam skeleton also suffered
Fig. 6, compared with non-discharge, the differences in the discharge significant losses, mainly caused by the excitation of MP, as shown in
intensity and frequency were obvious. At 900 W, HEOS produced a Fig. 8a. Fig. 8b illustrates the XRD results of the (FeMgCoNi)Ox solid
continuous discharge phenomenon. However, the discharge phenome­ compounds at five different stages: fresh, after the thermal reaction,
non at 700 W and 500 W was pulsed rather than continuous. At 700 W, after H2O splitting, after three cycles, and after six cycles of reactive
the discharge phenomenon was more significant and frequent than that tests. Compared with the fresh and oxidised samples, the reduced sam­
produced at 500 W. The discharge duration (Peak) and interval (Trough) ple showed a significant change in the spinel phase peak, while changes
were 26/2 s at 700 W and 20/10 s at 500 W, respectively. It is widely in the rocksalt phase peak were not detected, which is mainly related to
understood that when microwave interacts with a material, the material the release of lattice oxygen. The XRD results confirm that both the
properties determine the energy absorption characteristics of it for the rocksalt and spinel phases co-exist during thermal cycling. In other
electromagnetic field. In other words, the interactions between micro­ words, (FeMgCoNi)Ox did not fully transition to a single phase but was
wave and a certain material could lead to energy-directed convergence distributed between the rocksalt and spinel phases during thermal
of the relatively dispersed microwave field onto reactive sites of the cycling. In addition, it can be seen from the XRD pattern that with an
material, producing thermal or discharge effects around the material increase in the number of cycles, the XRD diffraction peak becomes
[47]. In fact, this environment of intense discharge can easily induce sharper, indicating that the crystallisation degree of the sample in­
microwave plasma, which is often used to split the hydrogen-bearing creases and the grain grows. After 3 cycles, the sample appeared sintered
liquid compounds to produce hydrogen [48,49]. In general, the high- (Fig. 7a), and the diffraction peak of the spinel phase was weakened.
energy electrons produced by a microwave electric field acceleration After 6 cycles, the sintering of the sample was greatly obvious, as shown
of 2.45 GHz collide with water vapor molecules, resulting in the disso­ in Fig. 7c, and only the diffraction peaks of the rocksalt phase could be
ciation of water vapor in plasma, channelled by vibrational and high- found in the XRD pattern, while the diffraction peaks of the spinel phase
energy electron excitation. Meanwhile, water vapor molecules are ion­ were difficult to distinguish. The results show that sintering destroys the
ised and dissociated into hydrogen and hydroxyl radicals [50], as shown phase transformation ability of HEOS and hinders the release of oxygen.
in Eq. (8)–(10).
H2 O→H - + OH⋅ (8) 3.3. XPS analysis

H− + e→H + 2e (9) The oxidation and reduction states of the HEOS during the reaction
at 700 W were characterised by XPS. The XPS results indicated that the
O2 + 2H2 O→4OH⋅ (10) HEO consists of Fe, Mg, Co, Ni, and O. Mg2+ doping is usually used to
Under the synergetic effects of the thermochemical cycle and increase the oxygen storage capacity of ferrite and does not participate
distributed hot spots formed by the microwaved SiC, a large amount of in the redox reaction [51,52], thus we analysed the XPS spectra of Fe 2P
hydrogen was produced by the HEOS and microwave plasma, especially (Fig. 9a), O 1s (Fig. 9b), Co 2p (Fig. 9c), and Ni 2p (Fig. 9d). As shown in
in the third cycle of Sample C (123.14 mL/g). This also explains why the Fig. 9a, after thermal reduction, the Fe 2p level in (FeMgCoNi)Ox
ratio of hydrogen and oxygen production in some cycles was much exhibited two peaks related to Fe 2p3/2 and Fe 2p1/2, with binding en­
higher than 2:1, because the generated oxygen reacts with excess water ergies at 711.7 eV and 724 eV, respectively, providing clear evidence for
the existence of Fe3+. The presence of satellite peaks of Fe 2p3/2 also

Fig. 6. The discharge light intensity and frequency of HEOS during TR step under different microwave power.

8
Y. Gao et al. Applied Energy 279 (2020) 115777

Fig. 7. Morphology of sol-gel synthesized (FeMgCoNi)Ox. (a) As-synthesized. (b) After three H2O splitting cycles. (c) After six H2O splitting cycles. The H2O splitting
cycle was operated at 700 W.

confirmed the existence of Fe3+ [53]. The component percentage of that Co and Ni sites maintained their initial state during the oxidation-
shakeup peaks increased from 24.2% in the reducing phase to 39.7% in reduction process. In conclusion, Co and Ni were essentially inactive
the oxidising phase. This phenomenon suggested that the composition of during redox, compared to Fe, suggesting that Fe is the only redox-active
Fe (II) and Fe (III) changed during the oxidation-reduction process element in (FeMgCoNi)Ox.
indicating that the Fe site was active. Characteristic peaks in the O 1s
XPS spectra were observed at 530.7 eV and 531.9 eV (Fig. 9b), attributed 3.4. Discussion
to metal oxides and metal carbonate, respectively. The reduction process
of microwave irradiation clearly changed the oxygen-rich HEO to an Preliminary reaction kinetics data indicated that HEOS has the
oxygen-vacancy-rich structure. However, the HEOS obtained by mi­ characteristics of slow water splitting kinetics and easy sintering in the
crowave calcination has a high proportion of metal carbonates. We microwave field, suggesting that there kinetics have potential for
speculated that this was due to the unique heating mode of the micro­ improvement. On the one hand, the activity of HEOS decreased mainly
wave, which caused the carbon source in the SiC foam skeleton to be due to the discharge effect under microwave radiation. It is worth noting
excited and integrated into the lattice of metal oxides. In addition, a that the discharge phenomenon of HEOS appeared to be periodical with
small amount of CO2 was produced during the cycle, which also respect to its intensity and duration after two TWS cycles (Fig. 6).
confirmed our hypothesis. After water splitting, the O 1s level exhibited However, under the same experimental conditions (700 W), there was
a peak with a binding energy of 533.4 eV, related to metal hydroxide, almost non-discharge phenomenon when directly irradiated SiC foam
indicating that the HEOS did not achieve thermodynamic equilibrium. (Fig. 6a), which indicated that the discharge phenomenon may mainly
For the Co 2p XPS spectra, two peaks at 780.08 eV and 796.25 eV could be caused by the uneven distribution of HEO on the SiC foam. Therefore,
be assigned to the spectra of Co 2p3/2 and Co 2p1/2 of Co2+ (Fig. 8d) this could be optimized by improving the synthesis method, or replacing
[54]. Additionally, peaks at 856.22 eV and 873.69 eV could be attrib­ the carrier. On the other hand, the slow water splitting kinetics of HEOS
uted to Ni 2p3/2 and Ni 2p1/2 spectra of Ni2+ (Fig. 8e) [55]. Both the could be improved by changing the reaction conditions, increasing the
profiles of Co 2p3/2 and Ni 2p3/2 in the reducing phase of (FeMgCoNi)Ox water vapour flow, and optimising the reactor. The next section of our
were quite similar to that of the oxidising phase of (FeMgCoNi)Ox, and research will proceed along these lines.
the peak shift was almost undetectable in the shakeup peaks, indicating In addition to CeZrO2-δ and NiFe2O4, a number of other oxides have

9
Y. Gao et al. Applied Energy 279 (2020) 115777

Fig. 8. (a) Image of discharge occurring at 700 W, and image of Sample C after 6 cycles: A: losses in the SiC foam skeleton. (b) XRD patterns of fresh and used
(FeMgCoNi)Ox. All the samples were peeled off from the HEOS and ground before SEM and XRD testing.

been studied for potential TWS application, and a summary of these


Energy out HHV
alternative materials is exhibited in Table 2. These materials have η= = × 100\% (12)
Energy in Welectric
generally been tested using traditional heating methods, with reduction
temperatures ranging from 1300 to 1500 ◦ C. Obviously, larger steam Here, Energy out as the enthalpy of combustion of the hydrogen
concentrations than the concentration we used in our experiments, thus (HHV), and Energy in as the electricity power consumption (Welectric) in
it was difficult to directly compare the properties of these materials. both the microwave and the electric furnace.
However, most of these alternative materials require significantly longer Fig. 10 presents the profile of the two processes in terms of the
reduction and oxidation cycle times (from 30 min to 5 h), due to kinetic electric input power and temperature. It was found that the energy ef­
limitations associated with the redox process [31]. This means that their ficiency of both microwave and electric furnace were very low at the
H2 productivity per unit time consumes more energy than our experi­ laboratory scale set-up, typically, less than 2.21%. Whereas, compared
ments. The microwave system is a rather complex system, particularly with electric furnace, microwave could reduce power consumption by
when combined with HEOS discharge phenomenon, and it is difficult to 97% in the TR process. The energy efficiency is critically dependent on
accurately evaluate the efficiency. Notwithstanding, to understand the the absorbed microwave power. Unfortunately, at present nearly 95% of
system feasibility, we attempted to estimate the energy efficiency of the the microwave energy used in our small-scale experimental laboratory
system by Eq. (12). configuration is lost (Fig. 10a), quite simply because the high-power

10
Y. Gao et al. Applied Energy 279 (2020) 115777

Fig. 9. XPS characterization of (FeMgCoNi)Ox after thermal reduction and after water splitting.

power and, most important, the increasing availability of low-cost car­


Table 2
bon-free electricity suggests that renewable sources of primary elec­
Non-exhaustive list of promising TWS materials reported in the literature for
tricity could be used for generation of microwaves [53].
comparison of performance. The H2 yields of our experiments are based on
average data, and the others are from literature reports.
4. Conclusions
Material H2 yield TR step WS step conditions
(mL/g) conditions
This work investigated a new method for continuous hydrogen
(FeMgCoNi)Ox 40.56 500 W for 4 700 W for 30 min – 28 vol% regeneration based on two-step water splitting, in which both the
(Our) min H2O
(FeMgCoNi)Ox 63.47 700 W for 4 700 W for 30 min – 28 vol%
thermal reduction and water splitting steps were carried out under the
(Our) min H2O same microwave irradiation power. The HEOS could produce a
(FeMgCoNi)Ox 18.74 900 W for 4 700 W for 30 min – 28 vol% maximum hydrogen generation rate of 13.89 mL/min/g at 700 W mi­
(Our) min H2O crowave intensity, and the maximum hydrogen yield was 122 mL/g.
(FeMgCoNi)Ox 10.1 [28] 1300 ◦ C for 5 h 800 ◦ C for 5 h – 9.5 vol%
This result exceeded the thermodynamic limits of state-of-the-art ma­
H2O/91 ppm H2
BaCe0.25Mn0.75O3 3.15 [31] 1350 ◦ C for 5.5 850 ◦ C for 20 min – 40 vol% terials such as spinel ferrites and ceria.
min H2O The microwave plasma induced by HEOS discharge can strengthen
CeO2 4.94 [31] 1500 ◦ C for 1000 ◦ C for 1200 s – 40 vol the water splitting process, but at the same time it also speeds up the
330 s % H2O sintering of the material. However, the discharge process is uniform and
Ce0.85Zr0.15O2-δ 3.21 [32] 1350 ◦ C for 40 850 ◦ C for 400 min – 86 vol
min % H2O
easy to reproduce, and hence this result establishes a foundation for
LaGa0.4Co0.6O3-δ 10.74 [32] 1350 ◦ C for 40 850 ◦ C for 400 min – 86 vol optimizing the discharge intensity and pulse properties by appropriate
min % H2O morphological design. The very short times (4 min) for O2 generation
Ce0.85Mn0.15O2 11.03 [56] 1500 ◦ C for 30 1150 ◦ C for 30 min – pH2O under microwave irradiation, and this highly effective energy transfer
min 0.5–0.84 atm
process in comparison to the conventional thermal reduction process
Ce0.85Ni0.15O2 15.46 [56] 1500 ◦ C for 30 1150 ◦ C for 30 min – pH2O
min 0.5–0.84 atm and the subsequent water splitting process is at the very heart of the
Ce0.85Co0.15O2 17.26 [56] 1500 ◦ C for 30 1150 ◦ C for 30 min – pH2O enhanced efficiencies. Furthermore, compared with conventional ther­
min 0.5–0.84 atm mal process, microwave system could reduce power consumption by
97%. The operated temperatures of microwave system ranging from 500
to 650 ◦ C, which are of practical significance for current industrial
microwave device is presently configured only for our very small sample
equipment. Although most reactor architectures for such reactions have
volume (14 cm3 HEOS/2 g HEO). However, future larger-scale micro­
used solar concentrators as heat sources, the increasing availability of
wave systems would be designed to achieve 99.9% efficient absorbed

11
Y. Gao et al. Applied Energy 279 (2020) 115777

Fig. 10. (a) Comparison of microwave process and


conventional thermal process. (b) The energy effi­
ciency evaluation between microwave process and
conventional thermal process. The conventional
thermal process was refer from [31]: In the TR step,
material was heated from room temperature to
1100 ◦ C with a ramp rate of 20 ◦ C/min, then further
heated to 1350 ◦ C with a ramp rate of 5 ◦ C min-1
and maintained at 1350 ◦ C for 40 min; In the WS
step, the reactor was cooled to 800 ◦ C with a rate of
20 ◦ C/min and maintained at 800 ◦ C for 400 min.

low-cost carbon-free electricity suggests that localised microwave Declaration of Competing Interest
heating could be used for TWS as well.
In the next stage of work, we will optimize high entropy oxides by The authors declare that they have no known competing financial
replacing the carrier or adding a high melting point support substance, interests or personal relationships that could have appeared to influence
and explore the optimal thermochemical cycle operating conditions for the work reported in this paper.
high entropy oxides with different pH or different cations.
Acknowledgement
CRediT authorship contribution statement
We gratefully acknowledge the financial support from the Natural
Yibo Gao: Formal analysis, Investigation, Writing - original draft. Science Foundation of Shandong Province (ZR2018MEE030), and Na­
Yanpeng Mao: Conceptualization, Methodology, Writing - review & tional Natural Science Foundation of China (21307075).
editing. Zhanlong Song: Supervision. Xiqiang Zhao: Project adminis­
tration. Jing Sun: Resources. Wenlong Wang: Validation. Guifang
Chen: Software. Shouyan Chen: Resources.

12
Y. Gao et al. Applied Energy 279 (2020) 115777

References hercynite family of water splitting cycles. Energy Environ Sci 2015;8:3687.
https://doi.org/10.1039/c5ee01979f.
[25] Kong Hui, Hao Yong, Jin Hongguang. Isothermal versus two-temperature solar
[1] Chu 1 Steven, Majumdar Arun. Opportunities and challenges for a sustainable
thermochemical fuel synthesis: A comparative study. Appl Energy 2018;228:
energy future. Nature 2012;488:294–303. https://doi.org/10.1038/nature11475.
301–8. https://doi.org/10.1016/j.apenergy.2018.05.099.
[2] Courbon Emilie, D’Ans Pierre, Permyakova Anastasia, Skrylnyk Oleksandr,
[26] Scheffe Jonathan R, Steinfeld Aldo. Oxygen exchange materials for solar
Steunou Nathalie, Degrez Marc, et al. A new composite sorbent based on SrBr 2 and
thermochemical splitting of H2O and CO2: a review. Mater Today 2014;17:341–8.
silica gel for solar energy storage application with high-energy storage density and
https://doi.org/10.1016/j.mattod.2014.04.025.
stability. Appl Energy 2017;190:1184–94. https://doi.org/10.1016/j.
[27] Bulfin B, Vieten J, Agrafiotis C, Roeb M, Sattler C. Applications and limitations of
apenergy.2017.01.041.
two step metal oxide thermochemical redox cycles; a review. J Mater Chem A
[3] Lewis Nathan S. Research opportunities to advance solar energy utilization.
2017;5:18951. https://doi.org/10.1039/c7ta05025a.
Science 2016;351:aad1920. https://doi.org/10.1126/science.aad1920.
[28] Zhai Shang, Rojas Jimmy, Ahlborg Nadia, Lim Kipil, Toney Michael F,
[4] Muhammad Taqi Mehran, Seong-Bin Yu, Dong-Young Lee, Jong-Eun Hong, Seung-
Jin Hyungyu, et al. The use of poly-cation oxides to lower the temperature of two-
Bok Lee, Seok-Joo Park, et al. Production of syngas from H2O/CO2 by high-
step thermochemical water splitting. Roy Soc Chem 2018:2172–8. https://doi.org/
pressure coelectrolysis in tubular solid oxide cells. Appl Energy 2018;212:759–70.
10.1039/C8EE00050F.
Doi: 10.1016/j.apenergy.2017.12.078.
[29] Ruan C, Tan Y, Li L, Wang J, Liu X, Wang X. A novel CeO2-xSnO2/Ce2Sn2O7
[5] He Yan-Rong, Yan Fang-Fang, Yu Han-Qing, Yuan Shi-Jie, Tong Zhong-Hua,
pyrochlore cycle for enhanced solar thermochemical water splitting. Am Inst Chem
Sheng Guo-Ping. Hydrogen production in a light-driven photoelectrochemical cell.
Eng 2017;63:3450–62. https://doi.org/10.1002/aic.15701.
Appl Energy 2014;113:164–8. https://doi.org/10.1016/j.apenergy.2013.07.020.
[30] Seo Keumyoung, Jeong Sang-Mi, Lim Taekyung, Ju Sanghyun. Continuous
[6] Gao Xiang, Liu Guanyu, Zhu Ye, Kreider Peter, Bayon Alicia, Gengenbach Thomas,
hydrogen regeneration through the oxygen vacancy control of metal oxides using
et al. Earth-abundant transition metal oxides with extraordinary reversible oxygen
microwave irradiation. Roy Soc Chem 2018;8:37958–64. https://doi.org/10.1039/
exchange capacity for efficient thermochemical synthesis of solar fuels. Nano
C8RA08055K.
Energy 2018;50:347–58. https://doi.org/10.1016/j.nanoen.2018.05.045.
[31] Barcellos Debora R, Sanders Michael D, Tong Jianhua, McDaniel Anthony H,
[7] Safari Farid, Dincer Ibrahim. A review and comparative evaluation of
O’Hayre Ryan P. BaCe0.25Mn0.75O3-d – a promising perovskite-type oxide for
thermochemical water splitting cycles for hydrogen production. Energy Convers
solar thermochemical hydrogen production. Energy Environ Sci 2018;11:3256.
Manage 2020;205:112182. https://doi.org/10.1016/j.enconman.2019.112182.
https://doi.org/10.1039/c8ee01989d.
[8] Miller JE, McDaniel AH, Allendorf MD. Considerations in the design of materials
[32] Chen Zhenpan, Jiang Qingqing, Cheng Feng, Tong Jinhui, Yang Min,
for solar-driven fuel production using metal-oxide thermochemical cycles. Adv
Jiang Zongxuan, et al. Sr- and Co-doped LaGaO3-δ with high O2 and H2 yields in
Energy Mater 2013;4(2):1300469. https://doi.org/10.1002/aenm.201300469.
solar thermochemical water splitting. J Mater Chem A 2019;7:6099–112. https://
[9] Muhich CL, Ehrhart BD, Al-Shankiti I, Ward BJ, Musgrave CB, Weimer AW.
doi.org/10.1039/C8TA11957K.
A review and perspective of efficient hydrogen generation via solar thermal water
[33] Cho HS, Gokon N, Kodama T, Kang YH, Lee HJ. Improved operation of solar
splitting. WIREs Energy Environ 2016;5:261–87. https://doi.org/10.1002/
reactor for two-step water-splitting H2 production by ceria-coated ceramic foam
wene.174.
device. Int J Hydrogen Energy 2015;40:114–24. https://doi.org/10.1016/j.
[10] Kubo Shinji, Kasahara Seiji, Okuda Hiroyuki, Terada Atsuhiko, Tanaka Nobuyuki,
ijhydene.2014.10.084.
Inaba Yoshitomo, et al. A pilot test plan of the thermochemical water-splitting
[34] Hoskinsa Amanda L, Millicana Samantha L, Czernika Caitlin E,
iodine–sulfur process. Nucl Eng Des 2004;233:355–62. https://doi.org/10.1016/j.
Alshankitia Ibraheam, Netterb Judy C, Wendelinb Timothy J, et al. Continuous on-
nucengdes.2004.08.018.
sun solar thermochemical hydrogen production via an isothermal redox cycle. Appl
[11] Safari Farid, Dincer Ibrahim. A study on the Fe-Cl thermochemical water splitting
Energy 2019;249:368–76. https://doi.org/10.1016/j.apenergy.2019.04.169.
cycle for hydrogen production. Int J Hydrogen Energy 2020;40:18867–75. https://
[35] Koepf E, Villasmil W, Meier A. Pilot-scale solar reactor operation and
doi.org/10.1016/j.ijhydene.2020.04.208.
characterization for fuel production via the Zn/ZnO thermochemical cycle. Appl
[12] Roeb Martin, Neises Martina, Monnerie Nathalie, Call Friedemann, Simon Heike,
Energy 2016;165:1004–23. https://doi.org/10.1016/j.apenergy.2015.12.106.
Sattler Christian, et al. Materials-related aspects of thermochemical water and
[36] Marxer Daniel, Furler Philipp, Takacs Michael, Steinfeld Aldo. Solar
carbon dioxide splitting: a review. Materials 2012;5:2015–54. https://doi.org/
thermochemical splitting of CO2 into separate streams of CO and O2 with high
10.3390/ma5112015.
selectivity, stability, conversion, and efficiency. Energy Environ Sci 2017;10:1142.
[13] Yadav Deepak, Banerjee Rangan. A review of solar thermochemical processes.
https://doi.org/10.1039/c6ee03776c.
Renew Sustain Energy Rev 2016;54:497–532. https://doi.org/10.1016/j.
[37] Arribas L, Aguilar JG, Romero M. Solar-driven thermochemical water-splitting by
rser.2015.10.026.
cerium oxide: determination of operational conditions in a directly irradiated fixed
[14] Romero Manuel, Steinfeld Aldo. Concentrating solar thermal power and
bed reactor. Energies 2018;11:1–15. https://doi.org/10.3390/en11092451.
thermochemical fuel. Energy Environ Sci 2012;5:9234. https://doi.org/10.1039/
[38] Oses C, Toher C, Curtarolo S. High-entropy ceramics. Nat Rev Mater 2020. https://
c2ee21275g.
doi.org/10.1038/s41578-019-0170-8.
[15] Chaubey Rashmi, Sahu Satanand, James Olusola O, Maity Sudip. A review on
[39] Rost Christina M, Sachet Edward, Borman Trent, Moballegh Ali, Dickey Elizabeth
development of industrial processes and emerging techniques for production of
C, Hou Dong, et al. Entropy-stabilized oxides. Nat Commun 2015;6:8485. https://
hydrogen from renewable and sustainable sources. Renew Sustain Energy Rev
doi.org/10.1038/ncomms9485.
2013;23:443–62. https://doi.org/10.1016/j.rser.2013.02.019.
[40] Anand G, Wynn AP, Handley CM, Freeman CL. Phase stability and distortion in
[16] Lu Youjun, Zhu Liya, Agrafiotis Christos, Vieten Josua, Roeb Martin,
high-entropy oxides. Acta Mater 2018;146:119–25. https://doi.org/10.1016/j.
Sattler Christian. Solar fuels production: Two-step thermochemical cycles with
actamat.2017.12.037.
cerium-based oxides. Prog Energy Combust Sci 2019;75:100785. https://doi.org/
[41] Sarkar Abhishek, Wang Qingsong, Schiele Alexander, Chellali Mohammed Reda,
10.1016/j.pecs.2019.100785.
Bhattacharya Subramshu S, Wang Di, et al. High-entropy oxides: fundamental
[17] Mao Yanpeng, Gao Yibo, Dong Wei, Wu Han, Song Zhanlong, Zhao Xiqiang, et al.
aspects and electrochemical properties. Adv Mater 2019;31:1806236. https://doi.
Hydrogen production via a two-step water splitting thermochemical cycle based on
org/10.1002/adma.201806236.
metal oxide – A review. Appl Energy 2020;267:114860. https://doi.org/10.1016/j.
[42] Gerbec Jeffrey A, Magana Donny, Washington Aaron, Strouse Geoffrey F.
apenergy.2020.114860.
Microwave-enhanced reaction rates for nanoparticle synthesis. J Am Chem Soc
[18] Hosseini Seyed Ehsan, Wahid Mazlan Abdul. Hydrogen production from renewable
2005;127(45):15791–800. https://doi.org/10.1021/ja052463g.
and sustainable energy resources: Promising green energy carrier for clean
[43] Jie Xiangyu, Gonzalez-Cortes Sergio, Xiao Tiancun, Yao Benzhen, Wang Jiale,
development. Renew Sustain Energy Rev 2016;57:850–66. https://doi.org/
Slocombe Daniel R, et al. The decarbonisation of petroleum and other fossil
10.1016/j.rser.2015.12.112.
hydrocarbon fuels for the facile production and safe storage of hydrogen. Energy
[19] Bhosale Rahul R, Sutar Parag, Kumar Anand, AlMomani Fares, Ali Moustafa
Environ Sci 2019;12:238. https://doi.org/10.1039/c8ee02444h.
Hussein, Ghosh Ujjal, et al.. Solar hydrogen production via erbium oxide based
[44] Thostenson ET, Chou T-W. Microwave processing: fundamentals and applications.
thermochemical water splitting cycle. J Renew Sustain Energy 2016;8:034702.
Compos A Appl Sci Manuf 1999;30(9):1055–71. https://doi.org/10.1016/s1359-
Doi: 10.1063/1.4953166.
835x(99)00020-2.
[20] Bhosale Rahul, Kumar Anand, AlMomani Fares, Ghosh Ujjal, Anis Mohammad
[45] Sautbekov Seil, Bourgiotis Sotiris, Chrysostomou Ariadni, Frangos Panayiotis.
Saad, Kakosimos Konstantinos, et al. Solar hydrogen production via a samarium
A Novel Asymptotic Solution to the Sommerfeld Radiation Problem: Analytic field
oxide-based thermochemical water splitting cycle. Energies 2016;9(316). Doi:
expressions and the emergence of the Surface Waves. Prog Electromagn Res M
10.3390/en9050316.
2018;64:9–22. https://doi.org/10.2528/pierm17082806.
[21] Bhosale Rahul, Kumar Anand, AlMomani Fares. Solar thermochemical hydrogen
[46] Demont Antoine, Abanades Stéphane, Beche Eric. Investigation of Perovskite
production via terbium oxide based redox reactions. Int J Photoenergy 2016:
structures as oxygen-exchange redox materials for hydrogen production from
9727895. https://doi.org/10.1155/2016/9727895.
thermochemical two-step water-splitting cycles. J Phys Chem C 2014;118(24):
[22] Diver Richard B, Siegel Nathan P, Miller James E. Two-step water splitting using
12682–92. https://doi.org/10.1021/jp5034849.
mixed-metal ferrites: thermodynamic analysis and characterization of synthesized
[47] Xue Chao, Mao Yanpeng, Wang Wenlong, Song Zhanlong, Zhao Xiqiang, Sun Jing,
materials. Energy Fuels 2008;22:4115–24. https://doi.org/10.1021/ef8005004.
et al. Current status of applying microwave-associated catalysis for the degradation
[23] Davenport Timothy C, Yang Chih-Kai, Kucharczyk Christopher J,
of organics in aqueous phase – A review. J Environ Sci 2019;81:119–35. https://
Ignatowich Michael J, Haile Sossina M. Maximizing fuel production rates in
doi.org/10.1016/j.jes.2019.01.019.
isothermal solar thermochemical fuel production. Appl Energy 2016;183:
[48] Nomura S, Toyota H, Mukasa S, Yamashita H, Maehara T, Kawashima A.
1098–111. https://doi.org/10.1016/j.apenergy.2016.09.012.
Production of hydrogen in a conventional microwave oven. J Appl Phys 2009;106:
[24] Muhich Christopher L, Ehrhart Brian D, Witte Vanessa A, Miller Samantha L,
73306. https://doi.org/10.1063/1.3236575.
Coker Eric N, Musgrave Charles B, et al. Predicting the solar thermochemical water
splitting ability and reaction mechanism of metal oxides: a case study of the

13
Y. Gao et al. Applied Energy 279 (2020) 115777

[49] Czylkowski D, Hrycak B, Miotk R, Jasiński M, Dors M, Mizeraczyk J. [53] Zhou Z, Zhang Y, Wang ZY, Wei W, Tang WF, Shi J, et al. Electronic structure
Hydrogenenriched gas production from kerosene using an atmospheric pressure studies of the spinel CoFe2O4 by X-ray photoelectron spectroscopy. Appl Surf Sci
microwave plasma system. Fuel 2017;215:686–94. https://doi.org/10.1016/j. 2008;254:6972–5. https://doi.org/10.1016/j.apsusc.2008.05.067.
fuel.2017.11.137. [54] Zong M, Huang Y, Wu HW, Zhao Y, Wang QF, Sun X. One-pot hydrothermal
[50] Chehade Ghassan, Lytle Spencer, Ishaq Haris, Dincer Ibrahim. Hydrogen synthesis of RGO/CoFe2O4 composite and its excellent microwave absorption
production by microwave based plasma dissociation of water. Fuel 2020;264: properties. Mater Lett 2014;114:52–5. https://doi.org/10.1016/j.
116831. https://doi.org/10.1016/j.fuel.2019.116831. matlet.2013.09.113.
[51] Han Sang Bum, Kang Tae Bum, Joo Oh Shim, Jung Kwang Deog. Water splitting for [55] Zong M, Huang Y, Ding X, Zhang N, Qu CH, Wang YL. One-step hydrothermal
hydrogen production with ferrites. Sol Energy 2007;81:623–8. Doi: 10.1016/j. synthesis and microwave electromagnetic properties of RGO/NiFe2O4 composite.
solener.2006.08.012. Ceram Int 2014;40:6821–8. https://doi.org/10.1016/j.ceramint.2013.11.145.
[52] Kuhn M, Bishop SR, Rupp JLM, Tuller HL. Structural characterization and oxygen [56] Gokon N, Suda T, Kodama T. Thermochemical reactivity of 5–15 mol% Fe Co, Ni,
nonstoichiometry of ceria-zirconia (Ce1− xZrxO2− δ) solid solutions. Acta Mater Mn-doped cerium oxides in two-step water-splitting cycle for solar hydrogen
2013;61:4277–88. https://doi.org/10.1016/j.actamat.2013.04.001. production. Thermochim Acta 2015;617:179–90. https://doi.org/10.1016/j.
tca.2015.08.036.

14

You might also like