You are on page 1of 20

Solar Energy xxx (2017) xxx–xxx

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Particle reactors for solar thermochemical processes


Tatsuya Kodama a,c,⇑, Selvan Bellan b,c, Nobuyuki Gokon b,c, Hyun Seok Cho a,c
a
Department of Chemistry & Chemical Engineering, Faculty of Engineering, Niigata University, 8050 Ikarashi 2-nocho, Niigata 950-2181, Japan
b
Center for Transdisciplinary Research, Niigata University, 8050 Ikarashi 2-nocho, Niigata 950-2181, Japan
c
Pacific Rim Solar Fuel System Research Center, Niigata University, 8050 Ikarashi 2-nocho, Niigata 950-2181, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Utilization of solar thermal power for high temperature fuel production has the potential to significantly
Received 10 April 2017 reduce the fossil fuel dependence of our current economy. Over the past two decades, remarkable pro-
Received in revised form 26 May 2017 gress has been made in the development of solar driven thermochemical reactors for the production of
Accepted 28 May 2017
hydrogen and syngas as they are promising energy carriers for transportation, domestic and industrial
Available online xxxx
applications. However, there are solar peculiarities in comparison to conventional thermochemical pro-
cesses – high thermal flux density and frequent thermal transients because of the fluctuating insolation-,
Keywords:
and conventional industrial thermochemical reactors are generally not suitable for solar driven reactors.
Particle reactors
Concentrated solar power
Therefore, solar-specific modifications of reactor design are necessary to realize efficient solar driven
Solar thermochemical processes thermochemical processes. In solar thermochemical reactors, the methods for solar-heating particulate
Solar gasification solid feedstock to high temperatures can be broadly classified as solar ‘‘directly” and ‘‘indirectly” absorb-
ing reactors. On solar thermochemical processes involving reacting solid particles at high temperatures,
such as ‘‘solar two-step water splitting with metal oxides” and ‘‘solar gasification”, various types of solar
directly and indirectly absorbing particle reactors have been developed. In this review, recent develop-
ment of solar particle reactors for the above solar thermochemical processes is described.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction with or geographically matching the incident solar radiation. In


order to efficiently convert solar high temperature heat to chemical
The maximum direct insolation frequently reaches 1 kW m2 in fuels in the ‘‘sunbelt” regions, industrially efficient endothermic
the insolation rich ‘‘sunbelt” regions which include the Southwest- processes are required which can be conducted at temperatures
ern United States, southern Europe, all of Australia, and broad ranging from 500 to 1500 °C (Kodama, 2003). Various important
regions of the developing world (Kodama, 2003). The collection industrial thermochemical processes exist, which may be con-
and concentration of direct insolation can be achieved by sun- ducted utilizing solar high temperature heat as the process heat.
tracking mirrors called collectors or heliostats. Some modern Some high temperature endothermic reactions for converting solar
solar-concentrating systems have maximum concentration factors energy to chemical fuels have been investigated as solar thermo-
in the 1500–5000 range and can provide high-temperature solar chemical processes, such as two-step water splitting cycles with
thermal heat (Kodama, 2003; Johnston et al., 2003; Mills, 2004; metal oxides, natural gas reforming, and coal gasification. There
Kalogirou, 2004; Yogev et al., 1998; Segal and Epstein, 2001; are, however, solar peculiarities in comparison to conventional
Tamaura et al., 2006; Epstein et al., 2006). The concentrated solar thermochemical processes: High thermal flux density (up to the
radiation is focused upon a solar receiver or reactor where maxi- maximums of 1500 and 5000 kW m2 for solar tower and dish type
mum temperatures can exceed 1500 °C, depending upon the con- concentrating systems respectively (Kodama and Gokon, 2007),
figuration of the solar concentrating system. and frequent thermal transients because of the fluctuating insola-
The conversion of solar radiation into chemical fuels is an engi- tion. The latter is caused by astronomical but more drastically by
neering challenge. Chemical fuels offer the advantages of being meteorological factors. Conventional industrial systems for ther-
transportable as well as storable for extended periods of time. This mochemical reactions are generally not suitable for transient oper-
point is important because energy demand is rarely synchronous ation: e.g., increased energetic losses by heat transfer resistance
and catalysts designed for optimum activity in a narrow tempera-
ture range. Although insolation is intense and very stable at day-
⇑ Corresponding author. time in the sunbelt (hence relatively stable reaction operation is
E-mail address: tkodama@eng.niigagta-u.ac.jp (T. Kodama). possible in solar driven chemical processes), still solar-specific

http://dx.doi.org/10.1016/j.solener.2017.05.084
0038-092X/Ó 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
2 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Nomenclature

CIEMAT Centro de Investigaciones Energéticas, Medioambien- PREC Professorship of Renewable Energy Carriers, ETHZ,
tales y Tecnológicas, Spain Switzerland
CNRS National Center for Scientific Research, France PROMES Procédés, Matériaux et Énergie Solaire, Odeillo, France
ETHZ Swiss Federal Institute of Technology, Zurich, Switzer- PSI Paul Scherrer Institute, Switzerland
land STL Solar Technology Laboratory, PSI, Switzerland
NREL National Renewable Energy Laboratory SSPS Small Solar Power Systems

modifications of conventionally used process technologies and operating between 298 and 1000 K is three, as inferred from the
reactor design are necessary to realize efficient solar thermochem- required entropy change.
ical processes. A two-step thermochemical water splitting cycle using a redox
Thermochemical reforming of natural gas is a catalytic reaction system of non-volatile metal oxide is one of the promising pro-
between methane with steam or CO2, and can proceed at temper- cesses for converting concentrated solar high-temperature heat
atures above 850 °C. Solar reformer can be realized by solar indi- into clean hydrogen in sun-belt regions. Nakamura (1977) first
rect heating of the catalyst and reactant gases. Two different proposed a two-step water-splitting cycle by a redox pair of
types of solar reformer using solar indirect heating have been Fe3O4/FeO, and conducted a thermodynamic analysis. The two-
developed: one is a heat exchange type and another is directly irra- step cycle proceeds as follows:
diated reactor tube type (Kodama, 2003). In the former case, solar
Fe3 O4 ! 3FeO þ 1=2O2 ð1Þ
radiant energy is collected by a conventional tube or volumetric
receiver, and the heat carrying medium heats the process gas in
H2 O þ 3FeO ! Fe3 O4 þ H2 ð2Þ
the separate reformer. In the latter case of the directly irradiated
reactor tube type, the concentrated solar radiation is directly The first high-temperature thermal reduction of Fe3O4 is highly
absorbed by a metal wall of the reactor tube and then the catalyst endothermic (DH°298K = 319.5 kJ/mol), and the second low-
bed inside the tube is indirectly heated. temperature hydrolysis by FeO is slightly exothermic (DH
However, some of solar thermochemical processes, such as °298K = 33.6 kJ/mol). The two-step process eliminates the need
solar gasification of coal and the thermal reduction of metal oxides for high-temperature H2/O2 separation. The first solar high-
as part of a two-step water splitting cycle, require a high- temperature step of the thermal reduction of magnetite (Fe3O4)
temperature solar chemical reactor being capable of operating at to wustite (FeO) proceeds at temperatures above 2500 K under
1000–1500 °C. In these cases, it is not easy to retain such high tem- 1 bar, while the second step of the hydrolysis reaction thermody-
peratures on coal or metal-oxide particles in conventional namically proceeds at temperatures below 1000 K (Kodama and
externally-heated reactor systems if concentrated solar radiation Gokon, 2007). Because of the very high reduction temperatures
is used as its energy source, because of high heat resistance of required for the Fe3O4/FeO redox pairs, mixed solid solutions
the exterior solar absorber wall. One effective way to reach and between the redox system Fe3O4/FeO and M3O4/MO have been
retain such high temperatures in a solar chemical reactor is by investigated to reduce the reduction temperature. Partial substitu-
direct irradiation with high fluxes of concentrated solar radiation. tion of iron in Fe3O4 by Ni, Co, etc. is possible to form mixed metal
In solar chemical reactors with reacting particles, it is possible that oxides (Fe1xMx)3O4. The mixed oxide may be reducible at a lower
the particles are heated directly by concentrated solar radiation, temperature than that required for the reduction of Fe3O4, while
resulting in efficient and rapid heating rates. Radiation enters the the reduced phase (Fe1xMx)1yO is still capable of a hydrolysis
gas-particle mixture through a quartz window and due to the reaction. However, another essential problem of the water splitting
low reflectivity of the particles, radiation is absorbed near the sur- cycle with iron-based oxides or ferrites is the deactivation of ferrite
face; heat losses are thus minimized. Direct absorbing particle particles in the cyclic reaction. This is due to melting or sintering of
receiver-reactors are candidates for conducting high-temperature the ferrite particles at temperatures below 1400–1500 °C (Kodama
solar driven processes. and Gokon, 2007). In order to solve the deactivation of ferrites in
This paper reviews the recent developments of particle solar cyclic reaction, ZrO2 supporting for ferrites has been investigated
reactors for thermochemical processes, which can convert solar (Kodama and Gokon, 2007; Kodama et al., 2005, 2008b). The sup-
radiation into chemical fuels. The target processes of particle solar porting ZrO2 alleviated the coagulation and/or sintering of the iron
reactors in this paper are ‘‘solar two-step water splitting cycles” oxides. Finally, Kodama et al. (2005) experimentally showed that
and ‘‘solar gasification”. cobalt-doped and nickel-doped iron oxides (Co-ferrites and Ni-
ferrites) supported on monoclinic ZrO2 (m-ZrO2) are capable of
working below 1400 °C as the reactive redox metal oxide material
(Kodama et al., 2005, 2008b).
2. Solar two-step water-splitting cycles with metal-oxides In recent years, non-stoichiometric cerium oxides have received
much more attention than ferrites as the working material because
2.1. Introduction of the high melting temperature of cerium oxides (Abanades and
Flamant, 2006; Chueh and Haille, 2009). Abanades and Flamant
Two-step reaction cycles are naturally the simplest multi-step (2006) first showed that CeO2/Ce2O3 two-step water splitting cycle
thermochemical water splitting methods. Funk and Reinstrom could produce hydrogen, but the thermal reduction temperature is
(1966) investigated the potential for the first two cycle types at very high (2000 °C) (Abanades and Flamant, 2006). However, when
temperatures up to 1100 °C, but feasible two-step cycles were the thermal reduction temperature is performed at 1500 °C, cerium
not found. This is in agreement with the thermodynamic analysis oxide is still reactive as the redox material for two-step water split-
made by Abraham and Schreiner (1974), which indicates that the ting (Chueh and Haille, 2009; Siegel et al., 2010). In this case, CeO2
minimum number of reaction steps for thermochemical cycles is thermally reduced to non-stoichiometric form in oxygen

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 3

(CeO2d). In the second step of the hydrolysis step, the CeO2d can et al., 2012; Haussener and Steinfeld, 2012). However, it must be
react with steam at lower temperatures to produce hydrogen. pointed out that such device type receiver-reactor concepts – the
Thus, the two-step reaction can be roughly written as follows: monolithic or foam device type – have a disadvantage that the
mass of the redox materials to be loaded onto the support is very
CeO2 ¼ CeO2d þ d=O2 T > 1500  C ð3Þ
limited and therefore, only a small amount of the redox material,
coated on the restricted surface area of the support, can contribute
CeO2d þ dH2 O ¼ CeO2 þ dH2 T < 1000  C ð4Þ
to the chemical reactions. The active surface area of the redox
The melting point data of CeO2 varies in the literature from metal oxide is very limited due to the small loading of the material
1950 to 2400 °C according to Abanades and Flamant (2006). But in the reactor. Therefore, a small mass of hydrogen will be pro-
in any case, the melting temperature of CeO2 is much higher than duced by ‘‘one run” of the cyclic reaction in the reactor. To realize
those of ferrites. a large mass production of hydrogen by such reactor concepts, a
As two-step water splitting cycle with volatile metal oxide, very large number of runs of the cyclic reaction are required. Dur-
ZnO/Zn cycle has been investigated the most (Palumbo et al., ing the day, the sun availability is limited in time, thus, to realize
1998; Steinfeld and Palumbo, 2001; Steinfeld, 2002, 2005; the very large number of runs for the limited period of time, very
Wegner et al., 2006). The two-step water splitting cycle by a fast kinetics of the reactions is required for the limited surface area
ZnO/Zn redox pair is described below. of the meal oxide in the reactor. In other words, the metal oxide
must realize a very high ‘‘turnover number”. However, the kinetics
ZnO ! ZnðgÞ þ 1=2O2 ð5Þ
of hydrolysis step strongly depends on the active surface area of
the redox materials such as ferrites (Kodama and Gokon, 2007),
Zn þ H2O ! ZnO þ H2 ð6Þ
thus, the large surface area will be required to realize the high
The thermal decomposition of ZnO to Zn vapor and O2 mole- reaction kinetics. As a result, in order to allow the reactor concept
cules proceeds endothermically (DH0298K = 478 kJ), and the temper- to be scaled up to multi-megawatt size, a cluster of the reactors
ature for which the DG0 equals zero is 2235 K (Steinfeld and must be constructed at the top of a solar tower (Roeb et al., 2016).
Palumbo, 2001). The hydrolysis by metallic zinc thermodynamically On the other hand, the idea of direct absorbing ‘‘particles” type
proceeds at temperatures below 1400 K under 1 bar (Wegner et al., receiver-reactors has been examined for more than three decades
2006). The products, zinc vapor and oxygen, eventually must be sep- (Rightley et al., 1992; Meier et al., 1996; Steinfeld et al., 1998). The
arated or quenched to avoid recombination. reactor concept is based on direct solar irradiation of ‘‘suspension
As mentioned above, the first solar endothermic step of thermal of reacting particles in a gas stream” or ‘‘falling particles”, providing
reduction of metal oxide in the two-step water splitting requires a efficient heat transfer directly to a large mass of reacting particles.
high-temperature solar reactor that can be operated at temperatures For solar two-step water splitting with metal oxides, the concept
above 1400 °C. It is difficult to retain such high temperatures of metal using a ‘‘suspension of reacting particles” has been well developed,
oxide particles in conventionally heated reactor systems. Thus, a and several reactors have been demonstrated as mentioned below.
specific, novel solar reactor must be developed for this extremely The difficulty in this concept is how to prevent the solar-heated par-
high-temperature gas-particle reaction (Steinfeld et al., 1999). Normal ticle suspension from contacting the transparent quartz window.
designs of solar reactor for high temperature applications make use of Provided that the reactor is combined with a conventional solar con-
insulated cavity-type configurations in order to obtain isothermal centrating system such as a solar tower (Kodama, 2003; Kodama and
conditions and efficient solar energy absorption (Diver, 1987). How- Gokon, 2007), solar high flux must enter horizontally into the reactor
ever, for operating temperatures above 1300 °C under oxidizing atmo- housing through a side window as illustrated in Fig. 1a. In this case,
spheres, ceramic materials of construction are usually required for conventional fluidized beds of reacting particles cannot be utilized in
lining the inner wall of the cavity. These massive refractory materials the solar reactor, but rather, a ‘‘cloud” of the reacting particles need
are not resistant to the severe thermal shocks that often take place to be created by more swirling gas streams in the reactor to avoid
under intermittent concentrated solar radiation. These problems pro- contact between the particle suspension and the window (Steinfeld
mote the development of the direct absorbing receiver-reactors of et al., 1998), as shown in Fig. 1a. Several reactors using a gas-
‘‘device” or ‘‘particles” types as a novel solar chemical reactor. particle vortex flow confined in a solar cavity receiver with a window
The direct absorbing ‘‘device” type receiver-reactors use honey- have been demonstrated (Steinfeld et al., 1998).
comb or foam ceramic devices coated with or made of some active On the other hand, it has been proposed that a solar chemical
redox materials such as ferrites or cerium oxide. In European reactor can be combined with the newly-developed solar reflective
HYDROSOL projects, multi-channeled honeycomb ceramic sup- tower with beam-down optics (Epstein et al., 2006; Kodama et al.,
ports coated with ferrite powder were used in a solar receiver- 2008a, 2014, 2016a, 2016b; Gokon et al., 2008b, 2009a, 2010,
reactor system, in a configuration similar to that encountered in 2011; Matsubara et al., 2015). The optical path of the beam-
automobile exhaust catalytic converters (Agrafiotis et al., 2005, down is comprised of a heliostat field illuminating a hyperbolic
2006; Roeb et al., 2016; Sattler et al., 2006). The ferrite-coated reflector that is placed on a tower and directs the beams down-
SiC-based honeycombs were directly irradiated by concentrated wards (Yogev et al., 1998; Segal and Epstein, 2001; Tamaura
solar radiation through a quartz window and heated to 1200– et al., 2006; Epstein et al., 2006). This beam-down setup has an
1300 °C in order to perform the thermal reduction step, while pass- advantage over normal tower top arrangements of reactor in that
ing N2 gas through it. Steam was passed through the reactor to per- it allows a large-scale reactor to be built on the ground, with the
form the hydrolysis step while decreasing the amount of solar solar radiation entering the reactor housing through a window in
power input because of the lower temperature of the hydrolysis the ceiling. In such a reactor design, conventional fluidized beds
step than that of the thermal reduction step. By other researchers of the reacting particles can be applied because an interspacing
(Kodama et al., 2007; Chueh et al., 2010; Furler et al., 2012) cera- gap can easily be produced between the fluidized particle bed
mic foam devices coated with or made of the redox materials (fer- and the window to prevent their contact, as shown in Fig. 1b. With
rites or cerium oxide) were also used as the direct absorbing this in mind, Kodama et al. (2008a) proposed and developed a
devices and demonstrated on two-step water splitting in directly novel design concept for chemical receiver-reactors using an inter-
solar-irradiated receiver-reactor systems (Kodama and Gokon, nally circulating fluidized bed of reacting particles (Kodama et al.,
2007; Kodama et al., 2007; Gokon et al., 2008a, 2008c, 2009b; 2008a, 2014; Gokon et al., 2009a, 2011; Matsubara et al., 2015). In
Cho et al., 2015; Petrasch et al., 2008; Chueh et al., 2010; Furler this receiver-reactor concept, the cylindrical reactor body is made

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
4 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Fig. 1. Schematics of solar reactor concepts using (a) a particle cloud and (b) an internally-circulating fluidized bed of reacting particles.

of steel, but a transparent quartz window is installed in the ceiling decades (Steinfeld, 2005; Agrafiotis et al., 2015; Alonso and
of the reactor. A draft tube is centrally inserted into the fluidized Romero, 2015). Many reactor concepts, from packed bed to
bed region. Gases are introduced into the draft tube and annulus quasi-batch arrangements (Schunk et al., 2008; Abanades et al.,
regions of the bed separately. The concentrated solar radiation 2007; Haueter et al., 1999; Chambon et al., 2010; Muller et al.,
passes downwards through the window and directly heats the 2006), have been considered by various researchers worldwide.
internally circulating fluidized bed of reacting particles. In this sys- Heat transfer in a direct absorbing receiver-reactor is predomi-
tem, the particles are always transported upwards in the draft tube nantly transferred to particles by radiative heat transfer and to
and move downwards in the annulus sections. This solid circula- the gas by particle–gas convective heat transfer. Hence, the radia-
tion within the reactor provides solar energy transfer from the tive heat transfer in a solar thermochemical reactor containing ZnO
top of the fluidized bed to the bottom since directly solar-heated particles, O2, and Ar was investigated by Dombrovsky et al. (2009).
particles in the top region always move to the bottom region. This Using the optical constants of ZnO, the local particle volume frac-
creates a more uniform temperature distribution throughout the tion, the particle size distribution and the spectral radiative prop-
solar-irradiated fluidized bed when compared with a normal erties of ZnO particles were calculated using approximation and
solar-irradiated fluidized bed. Another advantage is that the inter- Mie theory. Fig. 2 shows the absorption and transport scattering
nally circulating bed reactor requires a rapid gas flow only in a efficiency factors of ZnO particles for different sizes (10, 100 and
small region (solely in the draft tube), reducing the total gas flow 500 mm) Dombrovsky et al. (2009). For the development of solar
compared to a traditional fluidized bed reactor. This reduces the reactors for the Zn/ZnO thermochemical cycle, The STL in collabo-
energy required for gas flowing to the reactor extensively. ration with PREC at ETHZ has contributed significantly. In 1998,
In this section, recent R&D on particle ‘‘cloud” and ‘‘fluidized” started at the fundamental and laboratory scale reactor (Palumbo
receiver-reactors for solar two-step water splitting cycles are et al., 1998), which culminated with the 100 kW pilot scale of
described. In addition, indirect particle receiver-reactors and other demonstration process in 2014 (Koepf et al., 2016b). Hydrogen
conceptual reactors are also presented. production in zinc based thermochemical cycle has been exten-
sively studied at PSI. A rotary-cavity reactor called ROCA was the
2.2. Particle cloud vortex flow reactors for volatile ZnO process first fully functional ZnO dissociation reactor introduced at PSI
(Haueter et al., 1999), which is shown in Fig. 3. The main aim of
Solar reactors for the thermal reduction of ZnO have been under the reactor was to improve the design principles of the SLOP reac-
development and significant progress has been made for over two tor, which was initially developed with an inclined surface to

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 5

Fig. 2. Radiative characteristics of ZnO particles using approximation and Mie theory (Dombrovsky et al., 2009).

which particulate ZnO could be fed to investigate kinetics and addressed in the next generation solar reactor ZIRRUS, which was
extract practical operation parameters. developed by PSI following the ROCA.
The ROCA reactor featured a windowed rotating cavity-receiver The ZIRRUS reactor was successfully tested on the high flux
and the zinc oxide particles fed through a screw conveyor located solar furnace for dissociation of zinc oxide at 2000 K (Muller
at the rear end of the cavity. The zinc oxide powder was pushed et al., 2006; Müller and Steinfeld, 2007). However, the Hf/HfO2 cav-
towards the reactor wall due to centrifugal force and a ZnO layer ity of the reactor showed mechanical stability problems due to
was formed which served simultaneously as chemical reactant heating–cooling cycles and operation in oxidizing atmosphere
and thermal insulator of the cavity. In order to decrease the cavity (Schunk et al., 2008, 2009a). Thus an improved ZIRRUS reactor
aperture diameter and enhance the concentration of the incoming was designed and developed to operate at a power input of
radiation, compound parabolic concentrator (CPC) was used. The 10 kW as shown in Fig. 4. This reactor was also implemented a
particle-laden product stream was collected by a quench device rotating cavity. The inner walls of the cylindrical cavity were made
mounted at the rear end of the cavity. To reduce deposition, the of sintered Al2O3 composite bricks. A layer of ZnO of a desired
quartz window placed well in front of the cavity aperture and thickness was spread over the surface using a screw feeder. The
flushed with an inert gas stream. With this setup, solar tests were inert gas was used to sweep across the reactor window and carry
carried out on the high flux solar furnace with the nominal 10 kW the products to the quench device as ROCA reactor. Continuous
reactor prototype. ZnO surface temperatures of 2000 K were operation, about 4 h, was demonstrated for the ZIRRUS solar reac-
reached in 2 s and the reaction rates as high as 6 g/min were tor with 9.6 kW solar power input and the Zinc content of the
reported. However, this reactor was suffered from some problems quenched product was reported as 25.4 ± 7% (Schunk et al.,
such as uneven distribution of zinc oxide in the reactor, thermal 2008). Experimental tests around 1807–1907 K were carried out
loss due to water cooled front end, recombination of zinc and oxy- and no cracks or any mechanical stability problems were found.
gen to produce zinc oxide, inability to recover large percentage of The reactor was able to operate with much less water cooling. In
products from the reactor (Muller et al., 2006). These issues were further research, numerical models were developed to simulate

Fig. 3. Schematic of the ‘‘rotating-cavity” solar reactor concept for the thermal
dissociation of ZnO called ROCA (Haueter et al., 1999): (1) rotating cavity receiver, Fig. 4. Schematic of the improved reactor design (Schunk et al., 2008, 2009a,
(2) cavity aperture, (3) quartz window, (4) CPC, (5) outside conical shell, (6) 2009b); (1) rotating cavity; (2) Al2O3–SiO2 insulation; (3) Al2O3–Y2O3 insulation; (4)
reactant feeder, (7) ZnO layer, (8) purge gas inlet, (9) product outlet, (10) quench alumina fibers; (5) Al reactor shell; (6) aperture; (7) quartz window; (8) dynamic
device. feeder; (9) conical frustum; and (10) rotary joint.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
6 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

flow fields, heat and mass transfer (Schunk et al., 2009a, 2009b; the window (Koepf et al., 2016a). Due to an excessive use of
Villasmil et al., 2014; Müller and Steinfeld, 2007; Müller et al., quench gas (nearly 2000 l/min), the solar-to-fuel efficiency of the
2008; Dombrovsky et al., 2009, 2007; Coray et al., 2010; Lipinski pilot-scale solar reactor was limited. A maximum solar-to-
et al., 2006). The pre-exponential factor in the Arrhenius rate law chemical efficiency based on the collected products from the disso-
was used as a fitting parameter to match the experimental results. ciation reaction was reported at 3%, but the 100 kW reactor has the
A dynamic heat transfer models were developed for the zinc oxide potential to achieve as high as 16% when the reaction kinetics is
dissociation reactors (Villasmil et al., 2013b; Dombrovsky et al., improved (Villasmil et al., 2016). Apart from PSI’s thermochemical
2009). However, the dissociation kinetics of ZnO exposed to con- reactors for dissociation of ZnO, other researchers across the world
centrated solar radiation is still lacking due to; size of the exposed have been developing various solar reactor concepts for ZnO reduc-
surface and temperature at the reaction surface. The development tion (Abanades et al., 2007; Melchior et al., 2008; Perkins et al.,
work focused on optimizing the reactor’s operational parameters 2008; Haussener et al., 2009; Chambon et al., 2010, 2011; Koepf
for maximum exergy efficiency and improving the features of the et al., 2012, 2015a; Brkic et al., 2016).
previous concepts. Challenging tasks to be addressed are design,
fabrication and incorporation of a Zn/O2 separation device, mini- 2.3. Particle fluidized bed reactors for non-volatile meatal oxide
mization of back-flow of products through the aperture and processes
towards the quartz window.
After laboratory scale operation, the 10 kW ZIRRUS solar reactor Particle ‘‘fluidized bed” receiver-rector concept has been pro-
was up scaled at PROMES-CNRS to demonstrate zinc dissociation in posed and developed for water splitting cycles with non-volatile
a 100 kW pilot plant (Hutter et al., 2011; Villasmil et al., 2013a). On metal oxides such as ferrites and cerium oxide. This type of
sun testing experiments were carried out for more than 60 h water-splitting receiver-reactor concept was first experimentally
(Villasmil et al., 2013a). In-situ flow visualization technique was demonstrated using two kinds of laboratory-scale reactors.
demonstrated by directly visualizing reactant and product parti- Schematics of quartz tube and windowed stainless-steel tube reac-
cles illuminated by sunlight. Results reported that the successful tors are shown in Fig. 6 for the thermal reduction of a ferrite/zirco-
and long term operation of the reactor was acute due to the pro- nia particles (Kodama et al., 2008a; Gokon et al., 2008b, 2009a,
duct particle deposition over the quartz window (Koepf et al., 2010, 2011). The monoclinic ZrO2 supported NiFe2O4 (NiFe2O4/m-
2015b). After a series of proofing experiments conducted, the ZrO2) particle bed was internally fluidized by passing N2 gas
100 kW reactor was first commissioned in 2011. In the initial through from the bottom of the reactor. Then, the particle bed
design, alumina tiles and porous ceramic insulation materials were was irradiated by about 1–3 kWth of concentrated visible light
used to fabricate the cavity but it did not survive due to thermal from a sun-simulator to carry out the thermal reduction step of
shock and rotation. A cavity design, utilized alumina bricks the ferrite. In the case of the windowed stainless tube reactor,
arranged in an interlocking pattern, was demonstrated in 2012 the bed surface temperature on the top of the fluidized bed
(Villasmil et al., 2013a). Subsequently in 2014, the 100 kW reactor reached in the range 1200–1600 °C while the bed bottom temper-
was successfully demonstrated using the solar furnace of CNRS- ature was in the range 1000–1200 °C, depending on the flow rate of
PROMES after successfully addressing the protection of the quartz N2. After the irradiation, the thermally-reduced ferrite particles
window and the occasional clogging of the reactor’s outlet (Koepf were transferred to another fixed bed reactor and were reacted
et al., 2016b). Experiments were carried out over 13 days and with steam at 1000 °C. About 24–44% of the NiFe2O4 on the m-
28 kg of ZnO was dissociated. Fig. 5 shows the 100 kW ZnO solar ZrO2 support was converted to the reduced phase which was com-
reactor pilot plant developed at PSI (Koepf et al., 2016a). The pilot pletely re-oxidized with steam at 1000 °C to produce hydrogen
plant reaction temperature was above 2000 K and sustained for (Gokon et al., 2008b, 2009a, 2010). With this system, pulverization
over 97 h operation. The flow pattern to protect the quartz window of the NiFe2O4/m-ZrO2 particles is not needed following thermal
was optimized by in-situ high temperature flow visualization reduction – similar levels of hydrolysis reactivity were obtained
experiments (Koepf et al., 2015b). Thus, an uninterrupted opera- either with or without pulverizing the particles. This implies that
tion for three consecutive full days was achieved without cleaning the thermal reduction and hydrolysis steps can be conducted in a

Fig. 5. Schematic of the 100 kW ZnO solar pilot plant: (1) quartz window mount, (2) cavity aperture, (3) aluminum shell, (4) hopper containing ZnO, (5) retractable screw
feeder, (6) product hoses, (7) product filters, (8) movable carriage (Koepf et al., 2016b).

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 7

Fig. 6. Schematics of experimental set-ups for an internally circulating fluidized bed laboratory-scale reactors; (a) a ‘‘quartz tube” reactor and (b) a ‘‘windowed stainless steel
tube” reactor.

‘‘single” reactor of this type by switching the feed gas between


inert gas (nitrogen) and steam. Successive two-step water splitting
with NiFe2O4/m-ZrO2 particles in a ‘‘single” fluidized bed receiver-
reactor was also demonstrated using the windowed stainless tube
(Gokon et al., 2011). The internally circulating fluidized bed of
about 140 g of NiFe2O4/m-ZrO2 particles was exposed to concen-
trated visible light from a sun-simulator with an input power of
2.4–2.6 kWth for the thermal reduction step and 1.6–1.7 kWth for
subsequent hydrolysis step. The feed gas was switched from N2
gas in the thermal reduction step to a gas mixture of N2 and steam
in the hydrolysis step. The hydrogen productivity was 951 N cm3
per cycle and the ferrite conversion was about 35% in the succes-
sive two-step water splitting by the windowed stainless tube.
Also for CeO2 process, two-step water splitting in a ‘‘single” flu-
idized bed receiver-reactor was successfully demonstrated in a
large reactor (Omori et al., 2012). A new reactor was fabricated
for 5 kWth input power of concentrated visible light from a sun
simulator. The configuration of the reactor is shown in Fig. 7. It
consists of an internally circulating bed reactor tube (420 mm
length and 85 mm inner diameter with a 7 mm thickness). The
body of the reactor tube was made of Inconel and stainless steel.
A metal foam distributor of porous stainless steel frits was fixed
at the bottom of the fluidized bed region of the reactor tube. The
draft tube was centrally located in the fluidized bed region. The
inner diameter of the draft tube is 28 mm with 3 mm thickness Fig. 7. Schematic of 5 kW internally circulating fluidized bed reactor.
and the tube length was 48 mm. The bottom of the draft tube is
positioned at 18 mm height from the metal foam distributor. The
top of the reactor tube is equipped with a diverging conical funnel for the hydrolysis step. The O2 and H2 productivities reached 2.7
for mounting a quartz window. About 1 kg of CeO2 particle was and 5.0 N dm3 per cycle respectively. The molar ratio of H2/O2
loaded in the reactor. A successive reaction of thermal reduction was about 1.9 which was close to stoichiometric one of water split-
step (Eq. (3)) and hydrolysis step (Eq. (4)) in the single reactor ting. In order to control the fluidization of CeO2 particles, a basic
was performed by switching the feed gas between N2 gas and relationship between pressure drop of the passing gas through
H2O/N2 gas mixture. Simultaneously, the input power of concen- the fluidized bed and the gas flow velocity was experimentally
trated visible light into the reactor from a sun-simulator was also investigated using the CeO2 particles with different particle size
changed from 5.1 kW for the thermal reduction step to 0.4 kW by a small-scale quartz tube (Etori et al., 2015). Particle sizes smal-

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
8 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Fig. 8. (a) 30 kWth beam-down sun-simulator and (b) a bottom view of 19 xenon lamps in the simulator house.

ler than 300 lm were found to be desirable for the stable control of from 1050 to 1200 °C by the visible light irradiation with the aver-
fluidization of CeO2 particles. Then, the thermal reduction step of age heat flux of about 950 kW/m2, depending on the air flow rate.
the fluidized bed of CeO2 particles was investigated in more detail The output air temperature from the receiver reached 1000–
using this 5 kWth windowed receiver-reactor with the sun- 1060 °C (Kodama et al., 2016b). In the second test campaign,
simulator (Etori et al., 2015). CeO2 particles with a particle size two-step water splitting with CeO2 particles has been performed
range of 50–100 lm were used as the fluidized reacting particles using this 30 kWth reactor. About 8 kg of CeO2 particles was loaded
in the receiver-reactor, and the relation between O2 productivity in the reactor. The fluidized bed temperature of CeO2 particles
during the thermal reduction step and the N2 gas flow velocity reached above 1400 °C during the thermal reduction step while
was examined. A numerical computation for this windowed parti- passing N2 gas into the reactor. In subsequent hydrolysis step by
cles fluidized bed receiver-reactor was developed and studied by passing an H2O/N2 gas mixture at around 800 °C, about 10–
Matsubara et al. (2014). 20 N dm3 of hydrogen per cycle were successfully produced. The
In order to carry out laboratory performance tests of the solar detailed results will be presented in upcoming solarPACES confer-
particles fluidized bed receiver-reactor concept in a much larger ence 2017 (September 26–29, 2017 in Santiago, Chile).
scale, a 30–40 kWth big ‘‘beam-down” sun-simulator was installed This fluidized particle receiver-reactor concept needs to be
at Niigata University (Kodama et al., 2016b) as shown in Fig. 8. The combined with a beam-down type solar concentrator. Niigata
big ‘‘beam-down” sun-simulator has 19  7-kW xenon-arc lamps University, University of Miyazaki and Miyazaki prefecture have
with elliptical reflective mirrors. The lamps are fixed on the top started an R&D joint project since 2011, to demonstrate the up-
dome of the simulator house to focus towards downwards scaled fluidized bed receiver-reactor on solar. A new type of
(beam-down orientation). The lamps emit visible light beam 100 kWth beam-down solar concentrating system with a secondly
downwards and irradiate the focal spot with about 20-cm diame- elliptical reflector was built in August 2012 at the campus of
ter. The total beam power on the focal spot is about 30–40 kWth. University of Miyazaki for solar demonstration of the receiver-
The average light flux density within the focal circle area with a reactor. A snapshot of the demonstration plant is shown in Fig. 9
20-cm diameter is about 950 kW/m2. The 30-kWth fluidized bed (Kodama et al., 2014). The new beam-down system consists of an
receiver-reactor was newly fabricated. The basic structure of the elliptical mirror with 4.6 m diameter, a central tower with 16 m
receiver-reactor is similar to the previous 5 kWth receiver-reactor height, and 88 heliostats. Each heliostat consists of 10 small mir-
for CeO2 process (Fig. 7) but the draft tube was removed from this rors with 50 cm diameter. The heliostat field is 60  60 square
receiver-reactor. The gas distributor at the bottom of the fluidized meters, and the total area of mirrors is 176 m2. The beam-down
bed region was changed from metal foam to an Inconel perforated
flat plate. The perforated plate at the bottom of the bed was made
by many holes with 0.5 mm diameter. In order to provide the jet
flow along the central axial region at the bottom-inlet, the density
of holes in the central axial region of the perforated plate was
increased than the near wall region to create an internally-
circulating fluidization. The diameter and length of the static par-
ticles bed were 16 cm  16 cm. The body of the reactor was made
of Inconel and stainless steel. The reactor has a quartz window
with a diameter of 42 cm at the top. In the first test campaign,
quartz sand particles were used as the fluidized particles. The
quartz sand fluidized bed was created by passing air from the bot-
tom distributor of the receiver, and about 30 kWth of high flux vis-
ible light from 19 xenon-arc lamps of the sun-simulator was
directly irradiated on the top of the fluidized bed in the receiver
through a quartz window. The particle bed temperature at the cen-
ter position of the fluidized bed went up to a temperature range
Fig. 9. 100-kWth beam-down solar concentrating system at Miyazaki.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 9

the radiation at the second focal point, as illustrated in Fig. 10. In


order to increase the concentration of solar fluxes coming through
the inlet of the reactor, a CPC was designed and fabricated for this
new beam-down system, which is a conical frustum shape tunnel
(150 cm height, 75 cm inlet diameter, 45 cm exit diameter) of
reflecting mirrors. The CPC was located at the second focal spot
of the Miyazaki beam-down system. The performance test of con-
centrating solar fluxes of this beam-down system with and with-
out the CPC was done in 2012–2014 (Kodama et al., 2014). The
solar receiver-reactor was newly fabricated as shown in Fig. 11
for solar on-site performance test using the Miyazaki beam-down
system (Kodama et al., 2016a). The receiver has a quartz window
with a diameter of 75 cm at the top. In the first campaign of the
solar performance test, a fluidized bed of quartz sand particles with
air was used for the solar receiver-reactor. About 17 kg of quartz
sand with particle size of 100–500 lm was filled in the receiver-
reactor. The diameter and the height of the static bed was 25 cm.
The experimental system is illustrated in Fig. 12. The first solar test
Fig. 10. Principle of a novel beam-down system with an elliptical secondary
reflector.
campaign was carried out in the middle of December 2015. The
central bed temperature of the sand particles in the fluidized bed
solar receiver-reactor attained 1100 °C, when the direct normal
insolation or DNI was over 950 W/m2 at the peak time, using the
beam-down solar concentrating system. The outlet air temperature
of the receiver-reactor reached 1120 °C. It was predicted that the
central bed temperature could have been higher than 1100 °C if
solar receiver-reactor test had been conducted in other seasons
than winter. The second solar campaign of the particles fluidized
bed receiver-reactor using ‘‘CeO2 particles” for two-step water
splitting is currently going on.

2.4. Indirect particle reactors for metal oxide process

In indirectly heated reactors, the concentrated solar radiation is


first absorbed by a non-transparent wall and then transferred to
the reactants and catalysts by conductive and convective coupled
heat transfer. The flow region of indirect reactor contains the reac-
tants in the form of either packed bed or fluidized bed. The use of
Fig. 11. Photo of the 100-kWth solar particles fluidized bed receiver-reactor. indirectly irradiated solar reactors avoids the use of windows.
However, heat transfer is strongly influenced by several factors
such as packing density of the particles, size of the particles, their
system has the first focal point at 14 m height and the second focal
ability to move through the bed and physical properties of the
point at 10 m above ground level. The heliostat mirrors reflect the
reactants (Adinberg et al., 2004). Absorber material properties play
sunlight towards the first focal point of the elliptic mirror placed
a vital role and the following characteristics are required; chemi-
on the central tower, and then the elliptical mirror concentrates
cally inert and sustainable at high temperature, high at radiative

Fig. 12. Illustration of the experimental system.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
10 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Fig. 13. Multi-tube reflective cavity solar-thermal reactor (Lichty et al., 2012, 2013).

absorbance and thermal conduction, and resistive to thermal peratures and reaction rates in the range of 1780–1975 K
shocks (Piatkowski et al., 2011). In order to enhance the heat trans- (Melchior et al., 2008). A research group at University of Colorado
fer between the wall and the reactants, the use of intermediate tested a multi-tube reflective indirectly irradiated packed bed reac-
working fluids, such as gases or liquids have been used (Adinberg tor for water and carbon dioxide splitting (Lichty et al., 2012,
et al., 2004). 2013). Fig. 13 shows the multi-tube reflective cavity solar-
A lab-scale indirect 5 kW cylindrical cavity reactor containing a thermal reactor, which composed of a cooled reflective cavity with
tubular ceramic absorber, subjected to high-flux solar irradiation 5 reaction tubes. Cycling and robustness of the mixed metal oxide
in the range 448–2125 kW/m2, was fabricated at ETH. The selected particles were tested at National Renewable Energy Laboratory’s
chemical reaction was the thermal dissociation of ZnO into its ele- (NREL) High Flux Solar Furnace (HFSF). Reaction rates as high as
ments, which proceeds endothermically at above 1800 K. A heat 15.2 and 9.8 lmol/s/g were observed for H2O and CO2 splitting
transfer reactor model coupled with Monte Carlo ray-tracing was respectively.
developed and validated with the experimentally measured tem-

Fig. 14. The GRAFSTRR beam-down solar thermochemical reactor: (1) water-cooled window mount and vortex-flow generation, (2) water-cooled cavity aperture, (3) data-
acquisition cavity access ports, (4) alumina-tile reaction surface, (5) annular solid ZnO exit, (6) bulk insulation and cavity-shape support, (7) central product-vapor and gas
exit (Koepf et al., 2012).

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 11

Fig. 15. Schematic of solar reactor for continuous metal oxide processing (Abanades et al., 2007).

2.5. Other concepts of particle reactors for metal oxide processes approaches. The energy necessary to drive gasification reactions is
usually supplied by partial combustion of carbonaceous materials
A thermochemical reactor GRAFSTRR (Gravity-Fed Solar- with oxygen or air, but this releases large amounts of CO2 to the
Thermochemical Receiver/Reactor) was developed and tested for atmosphere (Tart, 1990; McMullan et al., 1997). The clean high tem-
the reduction of ZnO as shown in Fig. 14 (Koepf et al., 2012). The perature energy from concentrated solar radiation in the sunbelt may
reactor features an inverted conical-shaped reaction surface along be used to supply the process heat of gasification reactions. The use
which reactant powder descends continuously as a moving bed, of high temperature solar heat to drive the endothermic reactions
undergoing a thermochemical reaction at high temperature. The associated with gasification has been suggested and investigated
design objective of the GRAFSTRR concept was to simultaneously since the 1980s (Gregg et al., 1980a, 1980b; Aiman et al., 1981;
achieve three goals: (1) create a cavity-type reaction environment Taylor et al., 1983; Beattie et al., 1983; Epstein et al., 1994; Marray
to achieve high and uniform temperature, (2) disperse reactants and Fletcher, 1994; Kodama et al., 1998, 2001; Matunami et al.,
along the cavity walls as evenly and uniformly as possible, and 2000). This process, called ‘‘solar gasification”, produces not only a
(3) provide sufficient residence time for adequate reaction comple- highly useful and transportable end product, but also results in the
tion. Initial experiments using a high-flux solar simulator success- storage of a significant fraction of solar energy in the chemical bonds
fully demonstrated the mechanical stability of the reactor and of the fuel molecules. Thus, solar energy may be transformed into a
primary systems. Carbothermal reduction of ZnO with beech char- form that is both storable and transportable, and which can be used
coal was also performed in a directly irradiated 10 kW solar ther- in existing equipment.
mochemical reactor using a high-flux solar simulator. Solar The chemistry of gasification of coal and biomass involves the
thermochemical energy conversion efficiencies of up to 12.4% were two basic steps: pyrolysis and char gasification. The pyrolysis pro-
recorded, with reactant conversion as high as 14% (Koepf et al., cess can be roughly represented as follows:
2015a).
CNRS-PROMES designed a solar reactor for the continuous Coal or biomass ! charðcarbonÞ þ CO þ CO2 þ H2 þ CH4 þ tars
metal oxide reduction up to 2300 K under controlled atmosphere ð7Þ
as shown in Fig. 15 (Abanades et al., 2007). A computational model
was developed by coupling the fluid flow, heat and mass transfer, This process is slightly endothermic and the energy required for
and the chemical reaction. The model predicted that the initial par- this reaction is relatively small. The pyrolysis process is essentially
ticle size plays a vital role on the conversion rate as shown in complete by the time the coal reaches 1200 K (Gregg et al., 1980b).
Fig. 16. The maximum yield of particles recovery in the filter was
21% and the dissociation yield was up to 87% (Zn weight content
in the final powder) for a 5 NL/min gas flow-rate (Chambon
et al., 2010). A 1 kW lab-scale solar reactor was designed, built,
and operated for continuous dissociation of volatile oxides, ZnO
and SnO2 under reduced pressure (Chambon et al., 2011). The
kinetics of ZnO dissociation and activation energy was reported
by the O2 measurements.

3. Solar gasification

3.1. Introduction

Steam or CO2 gasification of carbonaceous materials, such as


coal, coke, biomass, etc., is highly endothermic, being a process
which is greatly dependent upon energy and high temperature
although the utilization of carbonaceous materials through conver-
sion to gaseous fuels embodies a number of different technical Fig. 16. Conversion rate vs. initial particle diameter (Abanades et al., 2007).

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
12 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

The primary reaction for chemically storing solar energy in pro- heated reaction zone. Scatter of the data was believed to result
duct gas is char gasification with steam, or CO2 in the next step. mainly from ash buildup and ash fusing, and from the different
The basic reaction of steam or CO2 gasification is the water gas positions of the reactor in the focal plane. For the CO2 gasification,
reaction (Eq. (8)), or the Boudouard reaction (Eq. (9)), which pro- a solar power level of 10.7 kW was used, and 27% of this solar
duces syngas: power was chemically stored in the product gas. Taylor et al.
(1983) demonstrated charcoal gasification with steam, or CO2 in
C þ H2 OðlÞ ! CO þ H2 DH 298K ¼ 175 kJ ð8Þ a packed bed solar gasifier with a plunger for using a 2-kW solar
furnace. The structure is illustrated in Fig. 17b. The packed bed
C þ CO2 ! 2CO DH 298K ¼ 172 kJ ð9Þ was pushed up by the plunger. The steam was generated by spray-
Because of these highly endothermic reactions, the product syn- ing water directly onto the surface of the charcoal and at the same
gas has greater calorific value, ideally higher than initial coal by time, heating the charcoal at the focus of a solar furnace. At an
44–45%. Equilibrium gas compositions of carbon-steam and optimum rate of water injection, half of the steam reacted with
carbon-CO2 systems, as a function of temperature, shows that, car- carbon and 30% of the incident solar energy was stored as chemical
bon conversion is completed above 1200–1300 K and at 1 bar, and enthalpy at solar power of 1.2 kW. In the charcoal gasification with
that only CO and H2 are produced in both systems (Kodama, 2003). CO2, the fraction of incident solar energy utilized to produce CO
Gregg et al. (1980b) presented a number of designs for solar ranged in 31–42% with the solar power of 670–1270 W. The chem-
gasification reactors, such as moving-bed and fluidized-bed solar ical storage efficiency range was 35 ± 5%, about the same as that
gasifiers, solar gasifiers using direct heat exchange between coal observed previously by Gregg et al. (1980a). Chemical storage effi-
and solar-heated molten salt, and a gasifier using a closed fluid ciencies of around 30–40% were obtained in the fixed-bed solar
(He) loop for heat transfer to the coal gasifier. It was anticipated coal gasifiers.
that each reactor design would have its own specific advantages However, in an industrial scale up, packed bed coal gasifiers
and technical problems. As mentioned above, in windowed solar involve some technical drawbacks, such as limitations of mass
gasifiers, coal particles can be heated directly by concentrated solar and heat transfer, long solid residence time and ash buildup to pre-
radiation, resulting in efficient and rapid heating rates. Radiation vent the reaction. From this point of view, other concepts of solar
enters the gas-particle mixture through a quartz window and gasifiers must be developed for industrial scale applications. In
due to the low reflectivity of the particles, radiation is absorbed Sections 3.2 and 3.3, the recent R&D on particle ‘‘cloud” and ‘‘flu-
on the surface, heat losses are thus minimized. Therefore, direct idized” solar gasifiers are reviewed. Indirect particle gasifiers are
absorbing particle receiver-reactors are also promising candidates reviewed in Section 3.4.
for solar gasification.
Gregg et al. (1980a) first demonstrated a fixed, packed-bed coal 3.2. Particle cloud vortex flow gasifiers
gasification with steam or CO2 using direct solar irradiation in a
23-kW solar furnace. The solar radiation was focused directly onto SYNPET (Hydrogen Production by Steam Gasification of Pet-
a fixed coal bed, through a quartz window, in an L-shaped gasifica- coke) is a collaborative project between CIEMAT, ETHZ and the
tion reactor. The illustration of the L-shaped reactor is given in industrial partner PDVSA (Venezuela). The main purpose of this
Fig. 17a. The coal bed was contained in the 30-cm diameter vertical project was to demonstrate the technical feasibility of solar gasifi-
section of the L, and a solar beam was admitted at the far end of the cation of waste from extra-heavy oil in the Fajadel Orinoco. It
horizontal member through a 20 cm-diameter quartz window. involves H2 production by solar gasification of petcoke and vacuum
Steam or CO2 was passed through the heated coal bed in the gasi- residue from heavy oil refining. In the initial phase of this project, a
fier. The coal bed temperature was measured at the positions about 5 kW prototype was developed and tested at PSI. From the
2 cm behind the reaction front of the coal bed, which varied from obtained results, the design of a solar reactor with quartz window
1175 to 1425 K. The product gases from the reactor were mainly was scaled up for an evaluation of a 500 kW installation at the PSA
CO and H2. There was a considerable range of scatter of the chem- CRS tower.
ical storage efficiencies from 19% to 48% in the solar power range of Using concentrated solar energy and petroleum coke, the ther-
4–23 kW. They indicated that ash buildup in the reaction zone modynamics and kinetics of the pertinent reactions were analyzed
caused less efficient operation by increasing the fraction of sun- for two types of petroleum coke: Flexicoke and Petrozuata delayed
light reflected and by reducing coal concentration in the solar- coke (Trommer et al., 2005). Based on reaction mechanisms

Fig. 17. Illustrations of (a) L-shaped, packed bed solar gasifier and (b) packed bed solar gasifiers with a plunger.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 13

(describing reversible adsorption/desorption processes) and irre- temperature range 1420–1567 K (Z’Graggen et al., 2008). The feed-
versible surface chemistry, kinetic rate laws were formulated. An stock was continuously injected as a liquid spray at 423 K into the
experimental quartz tubular reactor containing a fluidized bed of reactor’s cavity along with a coaxial steam flow. This reactor was
petroleum coke in steam was directly exposed to concentrated further used for upscaling the 5 kW prototype to a 300 kW and
radiation to determine the kinetic parameters and their solar gasification with petroleum vacuum residue and steam
Arrhenius-type temperature dependence. Syngas containing (Z’Graggen and Steinfeld, 2008). A two-phase reactor model that
approximately an equimolar mixture of H2 and CO and with a rel- couples radiative, convective, and conductive heat transfer to the
ative CO2 content of less than 5% was produced at above 1350 and chemical kinetics was developed to optimize the reactor geometri-
1550 K for Flexicoke and Petrozuata delayed coke respectively. Fol- cal configuration and operational parameters such as particle size,
lowing this study, design of a 5 kW prototype solar reactor was feeding rates, and solar power input. For the 300 kW reactor, com-
presented with modeling and testing results (Z’Graggen et al., plete reaction extent was predicted for an initial feedstock particle
2006). As shown in Fig. 18, the reactor features a continuous vortex size up to 35 lm at residence time of less than 10 s and peak tem-
flow of steam laden with petcoke particles confined to a cavity peratures of 1818 K. The upscaling of this reactor concept to a
receiver and directly exposed to concentrated solar radiation. The 500 kW at the SSPS-tower (PSA, Spain) was presented (Vidal
reactor was tested in a high-flux solar furnace at PSI in the range et al., 2010). The reactor was designed to produce about 100–
of 1300–1800 K, and up to 87% petcoke conversion was obtained 180 kg/h of syngas from 50 kg/h of petcoke.
in a single pass of 1 s residence time. The reactant mass flow rate
was in the range of 1.85–4.45 g petcoke/min and 3.68–9.04 g
3.3. Particle fluidized bed gasifiers
steam/min. In order to prevent the steam condensation and protect
it from radiation spillage, the window was oil cooled and kept in
Because of some technical drawbacks of a fixed-bed coal gasi-
the range 393–453 K. The window was kept clean by an aerody-
fier, such as limitations of mass and heat transfer, long solid resi-
namic protection created by a tangential flow through four tangen-
dence time and ash buildup to prevent the reaction, Epstein
tial nozzles (Z’Graggen et al., 2006).
suggested in 1994 that the circulating, or spouted fluidized-bed
A petcoke–water slurry was continuously injected into a solar
reactors are desirable (Epstein et al., 1994). The solid residence
cavity-receiver, to create a vortex flow directly exposed to concen-
time must be accomplished in minutes, depending on the chemical
trated radiation, and the effect of particle size (range 8.5–200 mm)
reactivity and temperature level. Earlier, Taylor et al. (1983)
and slurry stoichiometry (range 2.1–6.3) on the degree of chemical
demonstrated the fluidized-bed coal gasifier with CO2 under direct
conversion and energy conversion efficiency was examined
solar irradiation in their 2-kW solar furnace. The reactor was
(Z’Graggen et al., 2007). The solar energy conversion efficiency,
essentially a vertical silica-glass tube (5-cm diameter) located at
defined as the portion of solar energy absorbed as chemical energy
the focus of the solar furnace. The downward high-flux solar beam
and sensible heat, was about 17%. For solar inputs between 3.3 and
illuminated the upper surface of the fluidized bed in the silica-glass
5.1 kW, the energy conversion efficiency was between 0.5% and
tube. At a solar power of about 1.1 kW, the maximum chemical
4.7%. The average composition of the syngas production was: 62%
storage efficiency was 10% with 30% CO2 conversion. Marray and
H2, 25% CO, 12% CO2 and 1% CH4. The maximum degree of carbon
Fletcher (1994) also investigated the steam gasification in a flu-
conversion obtained was 87% at 1500 K, with water-petcoke molar
idized bed of cellulose using a quartz reactor and a solar furnace.
ratio of 4.8 and 2.4 s of residence time. Subsequently, the combined
The present author (Kodama et al., 2002) also tested a quartz tube
pyrolysis and steam-gasification of petroleum vacuum residue was
reactor for CO2 gasification of bituminous coal under direct irradi-
performed using a 5 kW aerosol-flow solar chemical reactor in the
ation of the fluidized coal bed with 0.3 kWth of concentrated visible

Fig. 18. Scheme of a vortex-flow solar thermochemical reactor for the steam-gasification of petcoke (Z’Graggen et al., 2006).

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
14 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

light from a sun-simulator. Von Zedtwitz and Steinfeld (2005) trated visible light could be reduced by optimizing the passing gas
modeled and investigated a quartz tube reactor containing a flu- velocities of steam, and the almost uniform temperature distribu-
idized bed of coal particles for steam generation. Muller et al. tion within the bed was created at around a temperature between
(2003) studied the reaction kinetics for steam gasification of a flu- 1000 and 1200 °C depending on the gas velocities. The test cam-
idized bed of coal using a quartz tube reactor. paign for coal coke gasification was performed by two operation
For about a decade, the present author (Kodama et al., 2008a) modes of the windowed gasifier. The first operation mode was a
has developed a ‘‘window-type” solar gasifier containing a flu- batch-type reaction mode. Steam was passed through the bed of
idized bed of coal particles exposed to concentrated visible light the mixture of coal coke and quartz sand to prepare a fluidized
(Kodama et al., 2008a, 2010; Gokon et al., 2012, 2014, 2015). bed and simultaneously gasify the fluidizing coal coke particles
Laboratory-scale prototype windowed gasifiers were fabricated by heating with direct light irradiation. The coke conversion after
for 1–3 kWth input power scales of concentrated visible light from 120 min of irradiation reached 60–88%. The peak light-to-
sun-simulators in order to test their reactor performances on chemical energy conversion was 5–13%. The second operation
steam or CO2 gasification of coal or coke particles. The structures mode involved a series of processes. This mode was essentially a
of the prototype windowed gasifiers are similar to the particle flu- batch-type reaction mode but carried out for future application
idized receiver-reactor for CeO2 process of water splitting (Fig. 7). to an operation mode in which coal coke particles are continuously
In the typical design of the prototype windowed gasifier for 3 kWth fed to the fluidized sand bed (by using screw feeder with hopper)
input power of concentrated visible light, the fluidized bed reactor at a high temperature above 1000 °C in the solar windowed gasi-
tube has a 420 mm length, and 62 mm inner diameter and 7 mm fier. A mixture of coal coke and quartz sand was initially fluidized
thickness. The reactor tube was made of a stainless steel by passing N2 gas while irradiating the bed top surface with con-
(SUS310S). A metal foam distributor of porous stainless steel frits centrated light in order to create a stable particle fluidization
was fixed at the bottom of the fluidized bed region in the reactor before gasification reaction. After reaching the steady state of flu-
tube. The internal centrally located draft tube has an inner diame- idization at around 1000 °C, passing gas was switched from N2 to
ter of 20 mm and is 3 mm thick and 44 mm long. The bottom of the steam to start the gasification reaction at the high temperature.
draft tube was positioned 28 mm above the distributor. The top of Immediately after switching the passing gas from N2 to steam,
the reactor tube was equipped with a diverging conical funnel for the strong peaks of CO and H2 evolution appeared. The gasification
mounting a quartz window. of coal coke was completed in 10–30 min after switching gas to
In the first concept of gasification by this prototype windowed steam. The coke conversion was about 93% and the light-to-
gasifier, coal or coke particles were solely loaded (without other chemical energy conversion was 13% at the peak.
fluidizing media such as sand particles) in the fluidized bed region, As mentioned above, at the new beam-down solar concentrat-
and the fluidized bed of the particles was directly exposed to con- ing system at Miyazaki, Japan, 17 kg of quart sand particles was
centrated visible light. Two designs of the fluidized bed reactor – fluidized in the 100-kWth windowed receiver-reactor and the top
internally circulating type with a draft tube and normal type with- surface of the fluidized bed was exposed to concentrated solar
out a draft tube – were compared on steam and CO2 gasification radiation from the beam down system through the quartz window
performances of coal coke (Kodama et al., 2008a, 2010; Gokon at the top of the receiver-reactor (Kodama et al., 2016a). The cen-
et al., 2012, 2014, 2015). The design with the draft tube, that tral bed temperature of the sand particles in the fluidized bed
enhances an internally circulation of the particles within the flu- exceeded 1100 °C even in the middle of December in Miyazaki,
idized bed, showed a much more uniform temperature distribution Japan. The fluidized bed temperature above 1000 °C is sufficient
within the fluidized bed along with the depth direction of the bed, for operating gasification reaction. This implies that by combining
superior gasification rates, coke conversion, and light-to-chemical the new beam-down system and the new windowed solar gasifier,
energy conversion (chemical storage efficiency) compared with solar gasification can be probably operated if coal particles are fed
the normal fluidized gasifier design without a draft tube. For steam to solar-heated fluidized bed of sand particles.
gasification of coal coke particles, the windowed, internally circu-
lating fluidized bed gasifier with a draft tube for 3 kWth input 3.4. Indirect particle gasifiers
power of light irradiation showed 10% of the peak light-to-
chemical energy conversion (Gokon et al., 2015). The coke conver- An indirectly heated 5 kW packed-bed reactor prototype was
sion was about 60% in 120 min of light irradiation. For CO2 gasifi- designed as shown in Fig. 19 for a beam-down system
cation of coal coke, a peak light-to-chemical energy conversion of (Piatkowski and Steinfeld, 2008). In which, the sunlight collected
12% was obtained by the same windowed gasifier with a draft tube by the heliostat field was redirected, by a hyperbolic reflector, to
for 3 kWth input power of light irradiation (Gokon et al., 2012). The the receiver at ground level. The reactor consisted of two cavities
coal coke conversion reached 73% in 120 min of light irradiation. A in series separated by SiC-coated graphite plate as a radiant emit-
kinetics analysis based on homogeneous and shrinking core mod- ter plate. A CPC (compound parabolic concentrator) was incorpo-
els was also done for CO2 gasification for the windowed, fluidized rated in the reactor’s aperture to increase the incident solar flux.
bed gasifier (Gokon et al., 2012). The upper cavity contained a small windowed aperture served as
In the second concept for the gasification by this windowed the solar absorber. The lower one was the reaction chamber and
gasifier, in order to enhance the gasification rate in the fluidized contained the reacting packed-bed. The reactor was tested with
bed, quartz sand particles were investigated as a chemically inert beech charcoal up to 1440 K packed-bed temperature. A quasi-1D
fluidized bed material which improves heat transfer within the transient thermal model was developed by coupling combined
bed and thermal storage of the bed during direct light irradiation radiative–conductive heat transfer with chemical reaction kinetics
of the bed (Gokon et al., 2014). The mixture of coal coke and quartz for steam gasification of coal. Model validation was accomplished
sand particles was loaded in the windowed gasifier and the steam by experimental measurements of bed temperatures, gasified mass
gasification of coke particles was performed by irradiating the top and bed shrink rates for beech charcoal. The overall reaction rate
surface of the fluidized bed of the mixture with about 3 kWth input was determined experimentally by thermogravimetry, while the
power from a sun-simulator. 96–320 g of quartz sand and 32– effective thermal conductivity was determined experimentally in
112 g of coke particles were used for the fluidized bed. The static a radial heat flow oven. The authors concluded that above
bed height was about 110 mm. The bed temperature distribution 1300 K, radiative transfer was the predominant mode of heat
during irradiating the top surface of the fluidized bed with concen- transport within the bed, consequently the effective thermal con-

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 15

Fig. 19. Indirectly irradiated solar packed-bed reactor for steam gasification of coal (Piatkowski and Steinfeld, 2008).

ductivity was significantly increased. This prototype reactor was entrained-flow solar reactor, which consisted of a 5 cm diameter
further used to test solar steam gasification of industrial and sew- cylindrical cavity receiver and a single 2.5 cm diameter SiC concen-
age sludge, scrap tire powder, fluff, south african coal and beech tric tubular absorber. The cavity receiver made of 10 wt% YO2-
charcoal (Piatkowski et al., 2009). It was reported that the reactor stabilized ZrO2 lined with Al2O3 insulation. The maximum thermal
was operated up to 1490 K and yielded high-quality syngas with energy conversion efficiency was 1.53%. A maximum carbon con-
typical molar ratios of H2/CO = 1.5 and CO2/CO = 0.2. The calorific version was 26% at 1425 K. A reactor model that couples radiative,
content was upgraded up to 30%. Solar-to-chemical energy conver- convective and conductive heat transfer to the chemical kinetics
sion efficiencies varied between 17.3% and 29%. A multiple pseudo- was formulated and validated by comparing with experimentally
component first-order reaction model was formulated to describe measured temperatures and carbon conversions. The developed
the rates of the combined pyrolysis and gasification processes by model was further applied to examine the thermal performance
Piatkowski and Steinfeld (2010). Arrhenius-type kinetic parame- of 100 kW and 1 MW scaled-up solar reactor containing multiple
ters were determined by dynamic thermogravimetric experimen- tubular absorbers. The predicted results showed that a maximum
tal runs conducted in the temperature range of 473–1476 K. solar-to-chemical energy conversion efficiency was about 39%
Further efforts were made by the same group (Piatkowski and and 50% for 100 kW and 1 MW respectively. The same model
Steinfeld, 2011) to develop a more complex dynamic model for was further improved and applied to estimate the performance
the solar-driven packed-bed reactor for the steam gasification of of a 10 MW scaled-up reactor for a solar tower system (Maag
heterogeneous, high-volatile content, carbonaceous waste feed- et al., 2010; Maag and Steinfeld, 2010). It was solved numerically
stocks. Model validation was accomplished in terms of converted by Monte Carlo and finite volume techniques. For an optimized
mass, reactor temperatures and efficiency based on experiments reactor geometry and a desired outlet temperature of 1500 K, a
with an 8-kW reactor (packed bed temperature up to 1490 K). solar-to-chemical energy conversion was 37%.
The model was further applied to analyze a 200 kW upscaled solar High temperature biomass gasification was performed in a pro-
gasifier of sludge operating under a simulated solar day; for a reac- totype concentrated solar reactor (Lichty et al., 2010). Fig. 22
tor having a 4 m diameter packed bed, energy conversion efficiency shows the schematic of the indirectly irradiated entrained-flow
was up to 89.4%. solar reactor which consisted of a cylindrical cavity receiver
A two-phase biomass char (biochar) steam gasification fluidized (18 cm diameter) with specular reflective inner surfaces containing
bed model was developed, as shown in Fig. 20, with concentrated 5 absorbing tubes. It was installed and tested at NREL’s high flux
solar heat as source of energy (Gordillo and Belghit, 2011). The solar furnace. For 7.5 kW solar power with an interchangeable
Rosseland approximation was used to calculate the radiative trans- cooling plate, the central tube temperature was 1660 K for alumina
fer within the bed (Gordillo and Belghit, 2011; Bardos et al., 1987). tubes. The highest temperature was achieved at the central tube
Their trends and predictions were in good agreement with the while the front and rear tubes were 350 and 250 K respectively.
reported results (Piatkowski et al., 2009). The steam-gasification Experiments were conducted to determine the effects of tempera-
of biochar with concentrated solar radiation was experimentally ture, gas flow rate, and feed type on the gasification performance. A
investigated using a 3 kW solar reactor consisting of a cylindrical tubular reactor was constructed and tested for hydrogen produc-
cavity-receiver containing an opaque tubular absorber (Melchior tion by biomass gasification in supercritical water using concen-
et al., 2009). Fig. 21 shows the scheme of the indirectly irradiated trated solar energy Chen et al. (2010). Fig. 23 shows the scheme

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
16 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Fig. 20. (a) Packed bed gasifier (Piatkowski et al., 2009) converted into (b) bubbling fluidized bed gasifier (b) (Gordillo and Belghit, 2011).

of the tubular reactor which was made of SS 316 stainless steel and transparent. Thus, it has been considered as a critical and trou-
with 10 mm outer diameter, 6 mm internal diameter and 18 m blesome component under high-pressure environments and large-
length. 6 kW experimental setup was fabricated. Glucose and bio- scale designs. Moreover, for operating temperatures above 1300 °C
mass (corn meal and wheat stalk) were gasified. The reactor was under oxidizing atmospheres, ceramic materials are usually
exposed to concentrated solar irradiation entering through the required for lining the inner wall of the cavity. These refractory
aperture of cavity. The complete gasification was achieved with materials are not resistant to the severe thermal shocks that often
0.2 M glucose as substrate, and the maximum yield ratio (the mass take place under intermittent concentrated solar radiation. These
of product gas/the mass of feedstock) reached up to 110%. This problems promote the R&D on the ‘‘direct solar absorbing parti-
study was further continued by a technical and economic analysis cles” reactor. Particles cloud vortex flow, windowed reactors have
and the cost for hydrogen production was estimated (Lu et al., been developed for volatile ZnO process of two-step water split-
2011). ting. In recent years, fluidized bed windowed reactor has been
developed for non-volatile ferrite or CeO2 process. These reactor
concepts have also been applied for solar gasification of carbona-
4. Summary ceous materials such as coal, coke, biomass, etc. and demonstrated
for the solar power input scales above 100 kWth.
In solar thermochemical particle reactors, directly irradiated In solar indirectly absorbing or irradiated reactors, the main
reactor is more efficient than the indirectly irradiated reactor. advantage is to avoid the use of window. However, heat transfer
However, one of the challenging tasks is to keep the window clean rate is less due to high thermal resistance of non-transparent wall.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 17

Fig. 22. Indirectly irradiated entrained-flow solar reactor for gasification of biomass
at high temperatures (Lichty et al., 2010).

in the field of directly irradiated reactors. As far as the indirectly


Fig. 21. Indirectly irradiated entrained-flow solar reactor for steam-gasification of
irradiated reactors is concerned, some examples of packed-bed
biochar (Melchior et al., 2009). and entrained-flow reactors can be found; a few reactors have been
developed for thermal reduction of metal oxides; most of the indi-
rectly heated reactors have been designed and tested for solar gasi-
fication of biomass and other carbonaceous materials like coke,
In addition, heat transfer is also affected by several other factors coal and petroleum residues.
such as packing density of the particles and physical properties From this review, it can be concluded that these studies on solar
of the reactants. To enhance the heat transfer between the wall particle reactors have been mostly done at laboratory or pilot plant
and the reactants, appropriate intermediate working fluids should scale levels. Further research has to be conducted to scale up theses
be used. According to the literature, more research has been done reactors towards the commercial scale level.

Fig. 23. Serpentine tubular reactor configuration for biomass gasification in supercritical water (Chen et al., 2010).

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
18 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Conflict of interest Gokon, N., Mizuno, T., Nakamuro, Y., Kodama, T., 2008c. Iron-containing yttria-
stabilized zirconia system for two-step thermochemical water splitting. J. Sol.
Energy Eng. 130, 011018.
The authors declared that there is no conflict of interest. Gokon, N., Hasegawa, T., Takahashi, S., Kodama, T., 2008a. Thermochemical two-
step water-splitting for hydrogen production using Fe-YSZ particles and a
ceramic foam device. Energy 33, 1407–1416.
References Gokon, N., Takahashi, S., Yamamoto, H., Kodama, T., 2008b. Thermochemical two-
step water-splitting reactor with internally circulating fluidized bed for thermal
reduction of ferrite particles. Int. J. Hydrogen Energy 33, 2189–2199.
Abanades, S., Flamant, G., 2006. Thermochemical hydrogen production from a two-
Gokon, N., Murayama, H., Nagasaki, A., Kodama, T., 2009b. Thermochemical two-
step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy 80,
step water splitting cycles by monoclinic ZrO2-supported NiFe2O4 and Fe3O4
1611–1623.
powders and ceramic foam devices. Sol. Energy 83, 527–537.
Abanades, S., Charvin, P., Flamant, G., 2007. Design and simulation of a solar
Gokon, N., Takahashi, S., Yamamoto, H., Kodama, T., 2009a. New solar water-
chemical reactor for the thermal reduction of metal oxides: case study of zinc
splitting reactor with ferrite particles in an internally circulating fluidized bed. J.
oxide dissociation. Chem. Eng. Sci. 62, 6323–6333.
Sol. Energy Eng. 131. 011007-1–011007-9.
Abraham, B.M., Schreiner, F., 1974. General principles underlying chemical cycles
Gokon, N., Yamamoto, H., Kondo, N., Kodama, T., 2010. Internally circulating
which thermally decompose water into the elements. Ind. Eng. Chem. Fundam.
fluidized bed reactor using m-ZrO2m-ZrO2 supported NiFe2O4NiFe2O4 particles
13 (4), 305–310.
for thermochemical two-step water splitting. J. Sol. Energy Eng. 132. 021102-1–
Adinberg, R., Epstein, M., Karni, J., 2004. Solar gasification of biomass: a molten salt
02112-10.
pyrolysis study. J. Sol. Energy Eng. 126, 850–857.
Gokon, N., Mataga, T., Kondo, N., Kodama, T., 2011. Thermochemical two-step water
Agrafiotis, C., Roeb, M., Konstandopoulos, A.G., Nalbandian, L., Zaspalis, V.T., Sattler,
splitting by internally circulating fluidized bed of NiFe2O4 particles: successive
C., Stobbe, P., Steele, A.M., 2005. Solar water splitting for hydrogen production
reaction of thermal-reduction and water-decomposition steps. Int. J. Hydrogen
with monolithic reactors. Sol. Energy 79, 409–421.
Energy 36, 4757–4767.
Agrafiotis, C., Lorentzou, S., Ragkoura, C., Kostoglou, M., Konstandopoulos, A.G.,
Gokon, N., Ono, R., Hatamachi, T., Li, L., Kim, H.J., Kodama, T., 2012. CO2 gasification
2006. Advanced monolithic reactors for hydrogen generation from solar water
of coal cokes using internally circulating fluidized bed reactor by concentrated
splitting. In: Proceedings of 13th International Symposium on Concentrated
Xe-light irradiation for solar gasification. Int. J. Hydrogen Energy 37 (17),
Solar Power and Chemical Energy Technologies, Seville, Spain.
12128–12137.
Aiman, W., Thorsness, C., Gregg, D., 1981. Solar Coal Gasification: Plant Design and
Gokon, N., Izawa, T., Abe, T., Kodama, T., 2014. Steam gasification of coal cokes in an
Economics. Lawrence Livemore Laboratory, Livemore, CA, USA. UCRL Preprint
internally circulating fluidized bed of thermal storage material for solar
84610.
thermochemical processes. Int. J. Hydrogen Energy 39, 11082–11093.
Agrafiotis, C., Roeb, M., Sattler, C., 2015. A review on solar thermal syngas
Gokon, N., Izawa, T., Kodama, T., 2015. Steam gasification of coal cokes by internally
production via redox pair-based water/carbon dioxide splitting
circulating fluidized-bed reactor by concentrated Xe-light radiation for solar
thermochemical cycles. Renew. Sustain. Energy Rev. 42, 254–285.
syngas production. Energy 79, 264–272.
Alonso, E., Romero, M., 2015. Review of experimental investigation on directly
Gordillo, E.D., Belghit, A., 2011. A bubbling fluidized bed solar reactor model of
irradiated particles solar reactors. Renew. Sustain. Energy Rev. 41, 53–67.
biomass char high temperature steam-only gasification. Fuel Process. Technol.
Bardos, C., Golse, F., Perthame, B., 1987. The rosseland approximation for the
92, 314–321.
radiative transfer equations. Commun. Pure Appl. Math. 40, 691–721.
Gregg, D., Aiman, W., Otuki, H., Thorsness, C., 1980b. Solar coal gasification. Sol.
Beattie, W., Berjoan, R., Coutures, J., 1983. High-temperature solar pyrolysis of coal.
Energy 24, 313–321.
Sol. Energy 31 (2), 137–143.
Gregg, D., Taylor, R., Campbell, J., Taylor, J., Cotton, A., 1980a. Solar gasification of
Brkic, M., Koepf, E., Alxneit, I., Meier, A., 2016. Vacuum powder feeding and
coal, activated carbon, coke and coal and biomass mixtures. Sol. Energy 25,
dispersion analysis for a solar thermochemical drop-tube reactor. Chem. Eng.
353–364.
Sci. 152, 280–292.
Haueter, P., Moeller, S., Palumbo, R., Steinfeld, A., 1999. The production of zinc by
Chambon, M., Abanades, S., Flamant, G., 2010. Design of a lab-scale rotary cavity-
thermal dissociation of zinc oxide—solar chemical reactor design. Sol. Energy
type solar reactor for continuous thermal dissociation of volatile oxides under
67, 161–167.
reduced pressure. J. Sol. Energy Eng. 132, 021006.
Haussener, S., Hirsch, D., Perkins, C., Weimer, A., Lewandowski, A., Steinfeld, A.,
Chambon, M., Abanades, S., Flamant, G., 2011. Thermal dissociation of compressed
2009. Modeling of a multitube high-temperature solar thermochemical reactor
ZnO and SnO2 powders in a moving-front solar thermochemical reactor. AIChE J.
for hydrogen production. J. Sol. Energy Eng. 131, 024503.
57, 2264–2273.
Haussener, S., Steinfeld, A., 2012. Effective heat and mass transport properties of
Chen, J., Lu, Y., Guo, L., Zhang, X., Xiao, P., 2010. Hydrogen production by biomass
anisotropic porous ceria for solar thermochemical fuel generation. Materials 5,
gasification in supercritical water using concentrated solar energy: system
192–209.
development and proof of concept. Int. J. Hydrogen Energy 35, 7134–7141.
Hutter, C., Villasmil, W., Chambon, M., Meier, A., 2011. Operational experience with
Cho, H.S., Gokon, N., Kodama, T., Kang, Y.H., Lee, H.J., 2015. Improved operation of
a 100 kW solar pilot plant for thermal dissociation of zinc oxide. In: Proceedings
solar reactor for two-step water-splitting H2 production by ceria-coated
of the 17th solarPACES Conference, Granada, Spain.
ceramic foam device. Int. J. Hydrogen Energy 40, 114–124.
Johnston, G., Lovegrove, K., Luzzi, A., 2003. Optical performance of spherical
Chueh, W.C., Haille, S.M., 2009. Ceria as a thermochemical reaction medium for
reflecting elements for use with paraboloidal dish concentrators. Sol. Energy 74,
selectively generating syngas or methane from H2O and CO2. ChemSusChem 2
133–140.
(8), 735–739.
Kalogirou, A.S., 2004. Solar thermal collectors and applications. Prog. Energy
Chueh, W.C., Falter, C., Abbott, M., Scipio, D., Furler, P., Haile, S.M., Steinfeld, A.,
Combust. Sci. 30, 231–295.
2010. High-flux solar-driven thermochemical dissociation of CO2 and H2O using
Kodama, T., Aoki, A., Shimizu, T., Kitayama, Y., Komarneni, S., 1998. Efficient
nonstoichiometric ceria. Science 330, 1797–1801.
thermochemical cycle for CO2 gasification of coal using a redox system of
Coray, P., Lipiński, W., Steinfeld, A., 2010. Experimental and numerical
reactive iron-based oxide. Energy Fuels 12, 775–781.
determination of thermal radiative properties of ZnO particulate media. J.
Kodama, T., Funatoh, A., Simizu, T., Kitayama, Y., 2001. Kinetics of metal oxide-
Heat Transfer 132, 012701.
catalyzed CO2 gasification of coal in a fluidized-bed reactor for solar
Diver, R., 1987. Receiver/reactor concepts for thermochemical transport of solar
thermochemical process. Energy Fuels 15, 1200–1206.
energy. Sol. Energy Eng. 109, 199–204.
Kodama, T., Kondoh, Y., Tamagawa, T., Funatoh, A., Shimizu, K.-I., Kitayama, Y., 2002.
Dombrovsky, L.A., Lipiński, W., Steinfeld, A., 2007. A diffusion-based approximate
Fluidized bed coal gasification with CO2 under direct irradiation with
model for radiation heat transfer in a solar thermochemical reactor. J. Quant.
concentrated visible light. Energy Fuels 16, 1264–1270.
Spectrosc. Radiat. Transf. 103, 601–610.
Kodama, T., 2003. High-temperature solar chemistry for converting solar heat to
Dombrovsky, L., Schunk, L., Lipiński, W., Steinfeld, A., 2009. An ablation model for
chemical fuels. Prog. Energy Combust. Sci. 29, 567–597.
the thermal decomposition of porous zinc oxide layer heated by concentrated
Kodama, T., Kondoh, Y., Yamamoto, R., Andou, H., Satoh, N., 2005. Thermochemical
solar radiation. Int. J. Heat Mass Transf. 52, 2444–2452.
hydrogen production by a redox system of ZrO2-supported Co(II)-ferrite. Sol.
Epstein, M., Spiewak, I., Funken, K., Ortner, J., 1994. Review of the technology for
Energy 78, 623–631.
solar gasification of carbonaceous materials. In: Proc. Solar Engineering Conf.
Kodama, T., Gokon, N., 2007. Thermochemical cycles for high-temperature solar
The ASME/JSME/JSES International Solar Energy Conference, San Francisco, CA,
hydrogen production. Chem. Rev. 107, 4048–4077.
USA, pp. 79–91.
Kodama, T., Hasegawa, T., Nagasaki, A., Gokon, N., 2007. Reactive Fe-YSZ coated
Epstein, M., Odalde, G., Santén, S., Steinfeld, A., Wieckert, C., 2006. Towards an
foam devices for solar two-step water splitting. In: Proceedings of the 2007
industrial solar carbothermic production of zinc. In: Proceedings of 13th
ASME Energy Sustainability Conference (ES2007), Long Beach, California. ASME,
International Symposium on Concentrated Solar Power and Chemical Energy
New York.
Technologies, Seville, Spain.
Kodama, T., Gokon, N., Yamamoto, R., 2008b. Thermochemical two-step water
Etori, T., Gokon, N., Takeuchi, A., Miki, T., Yokota, M., Kodama, T., 2015. Flowability
splitting by ZrO2-supported NixFe3xO4 for solar hydrogen production. Sol.
control of bed materials in a fluidized bed reactor for solar thermochemical
Energy 28, 73–79.
process. Energy Proc. 69, 1741–1749.
Kodama, T., Enomoto, S., Hatamachi, T., Gokon, N., 2008a. Application of an
Funk, J.E., Reinstrom, R.M., 1966. Energy requirements in the production of
internally circulating fluidized bed for windowed solar chemical reactor with
hydrogen from water. I&EC Prc. Des. Devel. 5 (3), 336.
direct irradiation of reacting particles. J. Sol. Energy Eng. 130. 014504-1–
Furler, P., Scheffea, J.R., Steinfeld, A., 2012. Syngas production by simultaneous
014504-4.
splitting of H2O and CO2 via ceria redox reactions in a high-temperature solar
reactor. Energy Environ. Sci. 5, 6098–6103.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx 19

Kodama, T., Gokon, N., Enomoto, S., Itoh, S., Hatamachi, T., 2010. Coal coke Müller, R., Lipin´ ski, W., Steinfeld, A., 2008. Transient heat transfer in a directly-
gasification in a windowed solar chemical reactor for beam-down optics. J. Sol. irradiated solar chemical reactor for the thermal dissociation of ZnO. Appl.
Energy Eng. 132. 041004-1–041001-6. Therm. Eng. 28, 524–531.
Kodama, T., Gokon, N., Matsubara, K., Yoshida, K., Koikari, S., Nagase, Y., Nakamura, Nakamura, T., 1977. Hydrogen production from water utilizing solar heat at high
K., 2014. Flux measurement of a new beam-down solar concentrating system in temperatures. Sol. Energy 19, 467–475.
Miyazaki for demonstration of thermochemical water splitting reactors. Energy Omori, K., Gokon, N., Hatamachi, T., Kodama, T., 2012. Two-step switching test of
Proc. 49, 1990–1998. water-splitting process using internally-circulating fluidized bed reactor. In:
Kodama, T., Gokon, N., Cho, H.S., Matsubara, K., Etori, T., Takeuchi, A., Yokota, S., Ito, SolarPACES Conference 2012, Marrakech, Morocco, September, 11–14.
S., 2016b. Particles fluidized bed receiver/reactor with a beam-down solar Palumbo, R., Léde, J., Boutin, O., Elorza, R.E., Steinfeld, A., Möller, S., 1998. The
concentrating optics: 30-kWth performance test using a big sun-simulator. AIP production of Zn from ZnO in a high-temperature solar decomposition quench
Conf. Proc. 1734, 120004. process—I. The scientific framework for the process. Chem. Eng. Sci. 53, 2503–
Kodama, T., Gokon, N., Cho, H.S., Matsubara, K., Kaneko, H., Senuma, K., Itoh, S., 2517.
Yokota, S., 2016a. Particles fluidized bed receiver/reactor with a beam-down Perkins, C., Lichty, P., Weimer, A., 2008. Thermal ZnO dissociation in a rapid aerosol
solar concentrating optics: quartz sand particles receiver test using an 100- reactor as part of a solar hydrogen production cycle. Int. J. Hydrogen Energy 33,
kWth solar concentrating system at Miyazaki. In: SolarPACES Conference 2016, 499–510.
Abu Dhabi, United Arab Emirates, October 11–14. Petrasch, J., Meier, F., Friess, H.M., Steinfeld, A., 2008. Tomography based
Koepf, E.E., Advani, S.G., Steinfeld, A., Prasad, A.K., 2012. A novel beam-down, determination of permeability, Dupuit-Forchheimer coefficient, and interfacial
gravity-fed, solar thermochemical receiver/reactor for direct solid particle heat transfer coefficient in reticulate porous ceramics. Int. J. Heat Fluid Flow 29,
decomposition: design, modeling, and experimentation. Int. J. Hydrogen Energy 315–326.
37, 16871–16887. Piatkowski, N., Steinfeld, A., 2008. Solar-driven coal gasification in a thermally
Koepf, E., Villasmil, W., Meier, A., 2015b. High temperature flow visualization and irradiated packed-bed reactor. Energy Fuels 22, 2043–2052.
aerodynamic window protection of a 100-kWth solar thermochemical Piatkowski, N., Wieckert, C., Steinfeld, A., 2009. Experimental investigation of a
receiver–reactor for ZnO dissociation. Energy Proc. 69, 1780–1789. packed-bed solar reactor for the steam-gasification of carbonaceous feedstocks.
Koepf, E., Advani, S.G., Prasad, A.K., Steinfeld, A., 2015a. Experimental investigation Fuel Process. Technol. 90, 360–366.
of the carbothermal reduction of ZnO using a beam-down, gravity-fed solar Piatkowski, N., Steinfeld, A., 2010. Reaction kinetics of the combined pyrolysis and
reactor. Ind. Eng. Chem. Res. 54, 8319–8332. steam-gasification of carbonaceous waste materials. Fuel 89, 1133–1140.
Koepf, E., Villasmil, W., Meier, A., 2016b. Pilot-scale solar reactor operation and Piatkowski, N., Steinfeld, A., 2011. Solar gasification of carbonaceous waste
characterization for fuel production via the Zn/ZnO thermochemical cycle. Appl. feedstocks in a packed-bed reactor – dynamic modeling and experimental
Energy 165, 1004–1023. validation. AIChE J. 57, 3522–3533.
Koepf, E., Villasmil, W., Meier, A., 2016a. Demonstration of a 100-kWth high- Piatkowski, N., Steinfeld, A., Wieckert, C., Steinfeld, A., Weimer, A.W., 2011. Solar-
temperature solar thermochemical reactor pilot plant for ZnO dissociation. AIP driven gasification of carbonaceous feedstock – a review. Energy Environ. Sci. 4,
Proc. 120005. 73–82.
Lichty, P., Perkins, C., Woodruff, B., Bingham, C., Weimer, A.W., 2010. Rapid high Rightley, M., Matthews, L., Mulholland, G., 1992. Experimental characterization of
temperature solar thermal biomass gasification in a prototype cavity reactor. J. the heat transfer in a free-falling-particle receiver. Sol. Energy 48, 363–374.
Sol. Energy Eng. 132, 011012. Roeb, M., Sattler, C., Klüser, R., Monnerie, N., de Oliveira, L., Konstandopoulos, A.G.,
Lichty, P., Liang, X., Muhich, C., Evanko, B., Bingham, C., Weimer, A.W., 2012. Atomic Agrafiotis, C., Zaspalis, V.T., Nalbandian, L., Steele, A., Stobbe, P., 2016. Solar
layer deposited thin film metal oxides for fuel production in a solar cavity hydrogen production by a two-step cycle based on mixed iron oxides. J. Sol.
reactor. Int. J. Hydrogen Energy 37, 16888–16894. Energy Eng. 128, 125–133.
Lichty, P., Muhich, C., Arifin, D., Weim, S.A., 2013. Methods and apparatus for gas Sattler, C., Roeb, M., Monnerie, N., Graf, D., Möller, S., 2006. Efficient solar thermal
phase reduction/oxidation processes. 13/857,951, US 2013/0266502 A1. processes from carbon based to carbon free hydrogen production. In:
Lipinski, W., Thommen, D., Steinfeld, A., 2006. Unsteady radiative heat transfer Proceedings of the ASME International Solar Energy Conference, Solar
within a suspension of ZnO particles undergoing thermal dissociation. Chem. Engineering. ASME, New York.
Eng. Sci. 61, 7029–7035. Schunk, L.O., Haeberling, P., Wepf, S., Wuillemin, D., Meier, A., Steinfeld, A., 2008. A
Lu, Y., Zhao, L., Guo, L., 2011. Technical and economic evaluation of solar hydrogen receiver-reactor for the solar thermal dissociation of zinc oxide. J. Sol. Energy
production by supercritical water gasification of biomass in China. Int. J. Eng. 130, 21009.
Hydrogen Energy 36, 14349–14359. Schunk, L.O., Lipiński, W., Steinfeld, A., 2009a. Heat transfer model of a solar
Maag, G., Rodat, S., Flamant, G., Steinfeld, A., 2010. Heat transfer model and scale-up receiver– reactor for the thermal dissociation of ZnO—experimental validation
of an entrained-flow solar reactor for the thermal decomposition of methane. at 10 kW and scale-up to 1 MW. Chem. Eng. J. 150, 502–508.
Int. J. Hydrogen Energy 35, 13232–13241. Schunk, L.O., Lipin´ ski, W., Steinfeld, A., 2009b. Ablative heat transfer in a shrinking
Maag, G., Steinfeld, A., 2010. Design of a 10 MW particle-flow reactor for syngas packed-bed of ZnO undergoing solar thermal dissociation. AIChE J. 55, 1659–
production by steam-gasification of carbonaceous feedstock using concentrated 1666.
solar energy. Energy Fuels 24, 6540–6547. Segal, A., Epstein, M., 2001. The optics of the solar tower reflector. Sol. Energy 69,
Marray, J.P., Fletcher, E.A., 1994. Reaction of steam with cellulose in a fluidized bed 229–241.
using concentrated sunlight. Energy 19, 1083–1098. Siegel, N.P., Garino, T., Coker, E.N., Livers, S., Ambrosini, A., Miller, J.E., Diver, R.,
Matsubara, K., Kazuma, Y., Sakurai, A., Suzuki, S., Soon-Jae, L., Kodama, T., Gokon, T., Bobek, M., 2010. Implementation of cerium oxide structures in solar fuel
Cho, H.S., Yoshida, Y., 2014. High-temperature fluidized receiver for production systems. In: Proc. 2009 SolarPACES Symposium, September 21–24,
concentrated solar radiation by a beam-down reflector system. Energy Proc. 2010, Perpignan, France.
49, 447–456. Steinfeld, A., Brack, M., Meier, A., Weidenkaff, A., Wuillemin, D., 1998. A solar
Matsubara, K., Sakai, H., Kazuma, Y., Sakurai, A., Kodama, T., Gokon, N., Cho, H.S., chemical reactor for co-production of zinc and synthesis gas. Energy 23, 803–
Yoshida, K., 2015. Numerical modeling of a two-tower type fluidized receiver 814.
for high temperature solar concentration by a beam-down reflector system. Steinfeld, A., Sanders, S., Ralumbo, R., 1999. Design aspects of solar thermochemical
Energy Proc. 69, 487–496. engineering – A case study: two-step water-splitting cycle using the Fe3O4/FeO
Matunami, J., Yoshida, S., Oku, Y., Yokota, O., Tamaura, Y., Kitamura, M., 2000. Coal redox system. Sol. Energy 65, 43–53.
gasification by CO2 gas bubbling in molten salt for solar/fossil energy Steinfeld, A., Palumbo, R., 2001. Solar thermochemical process technology. In:
hybridization. Sol. Energy 68, 257–261. Meyer, R.A. (Ed.), Encyclopedia of Physical Science & Technology, vol. 15.
McMullan, J., Williams, B., Sloan, E., 1997. Clean coal technologies. Proc. Inst. Mech. Academic Press, pp. 237–256.
Eng. 211 (Part A), 95–107. Steinfeld, A., 2002. Solar hydrogen production via a two-step water-splitting
Meier, A., Ganz, J., Steinfeld, A., 1996. Modeling of a novel high-temperature solar thermochemical cycle based on Zn/ZnO redox reactions. Int. J. Hydrogen Energy
chemical reactor. Chem. Eng. Sci. 51, 3181–3186. 27, 611–619.
Melchior, T., Perkins, C., Weimer, A., Steinfeld, A., 2008. A cavity-receiver containing Steinfeld, A., 2005. Solar thermochemical production of hydrogen––a review. Sol.
a tubular absorber for high-temperature thermochemical processing using Energy 78, 603–615.
concentrated solar energy. Int. J. Therm. Sci. 47, 1496–1503. Tamaura, Y., Utamura, M., Kaneko, H., Hasuike, H., Doming, M., Rellso, S., 2006. A
Melchior, T., Perkins, C., Lichty, P., Weimer, A.W., Steinfeld, A., 2009. Solar-driven novel beam-down system for solar power generation with multi-ring central
biochar gasification in a particle-flow reactor. Chem. Eng. Process.: Process. reflectors and molten salt thermal storage. In: Proceedings of 13th International
Intensificat. 48, 1279–1287. Symposium on Concentrated Solar Power and Chemical Energy Technologies,
Mills, D., 2004. Advances in solar thermal electricity technology. Sol. Energy 76, 19– Seville, Spain, 2006.
31. Taylor, R., Berjoan, R., Coutures, J., 1983. Solar gasification of carbonaceous
Muller, R., Zedtwitz, P.v., Wokaun, A., Steinfeld, A., 2003. Kinetic investigation on materials. Sol. Energy 30, 513–525.
steam gasification of charcoal under direct high-flux irradiation. Chem. Eng. Sci. Trommer, D., Noembrini, F., Fasciana, M., Rodriguez, D., Morales, A., Romero, M.,
58, 5111–5119. Steinfeld, A., 2005. Hydrogen production by steam- gasification of petroleum
Muller, R., Haeberling, P., Palumbo, R.D., 2006. Further advances toward the coke using concentrated solar power—I. Thermodynamic and kinetic analyses.
development of a direct heating solar thermal chemical reactor for the thermal Int. J. Hydrogen Energy 30, 605–618.
dissociation of ZnO(s). Sol. Energy 80, 500–511. Tart, K., 1990. Energy conversion through coal gasification. Energy World June, 15–
Müller, R., Steinfeld, A., 2007. Band-approximated radiative heat transfer analysis of 18.
a solar chemical reactor for the thermal dissociation of zinc oxide. Sol. Energy
81, 1285–1294.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084
20 T. Kodama et al. / Solar Energy xxx (2017) xxx–xxx

Vidal, A., Denk, T., Steinfeld, L., Zacarias, L., 2010. Upscaling of a 500 kW solar Yogev, A., Kribus, A., Epstein, M., Gogan, A., 1998. Solar ‘‘tower reflector” systems: a
gasification plant. 18th World Hydrogen Energy Conference. 17–20 May 2010, new approach for high-temperature solar plants. Int. J. Hydrogen Energy 23,
Essen, Germany, vol. 3, pp. 177–182. 239–245.
Villasmil, W., Meier, A., Steinfeld, A., 2013b. Dynamic modeling of a solar reactor for Zedtwitz, P.v., Steinfeld, A., 2005. Steam-gasification of coal in a fluidized-
zinc oxide thermal dissociation and experimental validation using IR bed/packed-bed reactor exposed to concentrated thermal radiation modeling
thermography. J. Sol. Energy Eng. 136, 011015. and experimental validation. Ind. Eng. Chem. Res. 44, 3852–3861.
Villasmil, W., Brkic, M., Wuillemin, D., Meier, A., Steinfeld, A., 2013a. Pilot scale Z’Graggen, A., Haueter, P., Trommer, D., Romero, M., de Jesus, J.C., Steinfeld, A., 2006.
demonstration of a 100-kWth solar thermochemical plant for the thermal Hydrogen production by steam-gasification of petroleum coke using
dissociation of ZnO. J. Sol. Energy Eng. 136, 011017. concentrated solar power—II. Reactor design, testing, and modeling. Int. J.
Villasmil, W., Meier, A., Steinfeld, A., 2014. Dynamic modeling of a solar reactor for Hydrogen Energy 31, 797–811.
zinc oxide thermal dissociation and experimental validation using IR Z’Graggen, A., Haueter, P., Maag, G., Vidal, A., Romero, M., Steinfeld, A., 2007.
thermography. J. Sol. Energy Eng. 136, 011015. Hydrogen production by steam-gasification of petroleum coke using
Villasmil, W., Copper, T., Koepf, E., Meier, A., Steinfeld, A., 2016. Coupled concentrated solar power—III. Reactor experimentation with slurry feeding.
concentrating optics, heat transfer, and thermochemical modeling of a 100- Int. J. Hydrogen Energy 32, 992–996.
kWth high-temperature solar reactor for the thermal dissociation of ZnO. J. Sol. Z’Graggen, A., Steinfeld, A., 2008. Hydrogen production by steam-gasification of
Energy Eng. 139, 021015. carbonaceous materials using concentrated solar energy – V. Reactor modeling,
Wegner, K., Ly, H.C., Weiss, R.J., Pratsinis, S.E., Steinfeld, A., 2006. In situ formation optimization, and scale-up. Int. J. Hydrogen Energy 33, 5484–5492.
and hydrolysis of Zn nanoparticles for H2 production by the 2-step ZnO/Zn Z’Graggen, A., Haueter, P., Maag, G., Romero, M., Steinfeld, A., 2008. Hydrogen
water-splitting thermochemical cycle. Int. J. Hydrogen Energy 31, 55–61. production by steam-gasification of carbonaceous materials using concentrated
solar energy—IV. Reactor experimentation with vacuum residue. Int. J.
Hydrogen Energy 33, 679–684.

Please cite this article in press as: Kodama, T., et al. Particle reactors for solar thermochemical processes. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.084

You might also like