You are on page 1of 19

Solar Energy 139 (2016) 676–694

Contents lists available at ScienceDirect

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

Exploitation of thermochemical cycles based on solid oxide redox


systems for thermochemical storage of solar heat. Part 5: Testing of
porous ceramic honeycomb and foam cascades based on cobalt and
manganese oxides for hybrid sensible/thermochemical heat storage
Christos Agrafiotis ⇑, Andreas Becker, Martin Roeb, Christian Sattler
Deutsches Zentrum für Luft- und Raumfahrt/German Aerospace Center (DLR), Linder Höhe, 51147 Köln, Germany

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

Article history: Cascaded ThermoChemical Storage (CTCS) of solar energy is a concept targeted to increase the volumetric
Received 21 April 2016 energy storage density and address the thermocline temperature distribution inside regenerative
Received in revised form 5 September 2016 sensible-only storage systems. CTCS involves the use of cascades consisting of different thermochemical
Accepted 15 September 2016
systems, distributed in a rational pattern inside the storage module tailored to their thermochemical
Available online 30 September 2016
characteristics and to the local heat transfer medium temperature. In the case of air-operated Solar
Thermal Power Plants such cascades can consist of porous structures incorporating different redox pair
Keywords:
oxide materials that can come in direct contact with the air stream used as heat transfer fluid and operate
Solar energy
Cascaded thermochemical heat storage
as compact, hybrid sensible-thermochemical storage modules in series.
Cobalt oxide Having previously identified the Co3O4/CoO and Mn2O3/Mn3O4 redox pairs as the most promising
Manganese oxide single-oxide systems for solar energy thermochemical storage, lab-scale (£ 25 mm), Co3O4- and
Cordierite Mn2O3-coated, porous cordierite honeycombs and foams were prepared and tested with respect to their
Ceramic honeycombs thermochemical characteristics in one- and two-oxides cascaded configurations employing redox oxide
Ceramic foams quantities in the range 15–150 g. For such Co3O4-loaded cascades thermochemical storage was clearly
demonstrated as heat uptake/release at constant temperature under proper testing conditions. Besides,
the additive effect of thermochemical on sensible storage within the same storage volume was visualized.
The operating conditions of cascades including both Co3O4 and Mn2O3 were dictated by the redox beha-
viour of the Mn2O3/Mn3O4 pair. Under proper conditions, such two-oxides-cascades could undergo cyclic
reduction-oxidation and operate in complementary temperature ranges during oxidation; however the
thermal effects of only the CoO oxidation reaction could be materialized into temperature rise of the
air stream exiting the cascade.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction a fluid passing through it – provides for the efficient use of air as
heat transfer medium despite its low heat transfer coefficient. This
A major advantage of Solar Thermal Power Plants (STTPs) oper- receiver technology has been developed on a solar tower since the
ating with Concentrated Solar Power (CSP) vs. other solar energy early 1990s over a series of projects culminating to the first-of-its-
technologies is that electricity production can be decoupled from kind, 1.5 MWel such CSP plant, Solar Tower Jülich (STJ) in Germany.
sunshine hours by storing the heat and converting it to electricity In STJ (Alexopoulos and Hoffschmidt, 2010) air at atmospheric
when needed allowing thus a fully-dispatchable supply. Among pressure flows through a solar-irradiated volumetric receiver made
the various CSP technologies, solar towers offer the potential of of SiC honeycombs, being thus heated up to 700 °C to power a
high temperatures and thus high thermodynamic conversion effi- steam generator. In parallel, a part of the solar-heated air is
ciencies. In the other hand, the technology of volumetric ceramic diverted to a regenerative, sensible heat, Thermal Energy Storage
receivers – where concentrated radiation is absorbed inside the (TES) system (Tamme et al., 1991) where it transfers its enthalpy
volume of a porous structure that is used to transfer this heat to to a solid medium as it flows through it (‘‘charging”). The storage
module consists of a 7 m  7 m  6 m housing (Fig. 1b) filled with
⇑ Corresponding author. flow-through ceramic honeycombs to provide a large air-solid heat
E-mail address: christos.agrafiotis@dlr.de (C. Agrafiotis). exchange surface, according to the results of studies on

http://dx.doi.org/10.1016/j.solener.2016.09.013
0038-092X/Ó 2016 Elsevier Ltd. All rights reserved.
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 677

Fig. 1. (a) Layout of Solar Tower Jülich (STJ) power plant and (b) schematic of its sensible regenerative heat storage subsystem based on ceramic honeycombs (Zunft et al.,
2011); (c) example of air temperature distribution in such a regenerative atmospheric air storage system of an air-operated solar power plant (Fricker, 2004); (d) principle of
cascaded thermochemical storage configuration and operation during charging (left) and discharging (middle) of such a system, tailored to the characteristics of the Co3O4/
CoO and Mn2O3/Mn3O4 redox pairs and exemplary schematic (right) with such oxides coated/shaped in various porous supports with possibilities of porosity variation along
and across the cascade; (e) and (f) porous cordierite honeycombs and foams from various suppliers and of various sizes (<20 mm, £ 25 mm and £ 90 mm for testing in a TGA,
the current testing rig and a solar receiver respectively) before and after their coating with Mn2O3.

optimization of volumetric energy storage density (kWh/m3) and constant air mass flow rate of constant temperature. The heat trans-
pressure drop (Zunft et al., 2014, 2011). During off-sun operation fer from the air to the solid and vice versa must take place with a
(‘‘discharging”) the air flow is reversed: ‘‘warm” air is introduced small temperature difference, i.e. with a small exergy loss. The
through the lower end of the storage medium to be heated by that result is a small temperature drop when discharging and a good
as it flows towards the already ‘‘hotter” top end, before introduced storage material utilization. Initially, during charging for example,
again to the power block. the temperature profile along the solid storage medium is stratified
In a regenerative sensible TES system, operation starts from an as it is shown in Fig. 1c (Fricker, 2004). At first the layer of the solid
isothermal cold or hot core, charging or discharging it with a medium adjacent to the air inlet is heated to the constant, inlet air
678 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

temperature; the other end of the solid remains practically at its each one coated or shaped to a honeycomb ceramic can be con-
initial temperature, whereas in-between exists a thermocline zone ceived as per the sketch in Fig. 1c.
(TCZ) of the storage material where the heat transfer takes place
2 Co3 O4 þ DH ¢ 6CoO þ O2 Tred  890  C Tox  870  C ð1Þ
and the charging air is cooled. This ‘‘constant high-temperature
profile” moves progressively along the storage medium from the
‘‘hot” to the ‘‘cold” end, until finally the whole medium reaches a 6Mn2 O3 þ DH ¢ 4Mn3 O4 þ O2 Tred  950  C Tox  690  780  C
temperature close to that of the incoming heat transfer fluid. A sim- ð2Þ
ilar procedure is applied for the discharging stage, where the dis-
A drawback of this configuration is that during on-sun opera-
charging air is heated in the TCZ. The longer the zones of
tion the entire cascade has to be heated up above the higher reduc-
practically constant temperature above and below the TCZ are in
tion temperature, that of Mn2O3; in the other hand during off-sun
relation to the storage core flow path, the better is the utilization
operation the ‘‘warm” air entering from the bottom needs to initi-
of the storage material and the longer the time during which the
ate first the Mn3O4 oxidation reaction taking place at much lower
storage module can supply practically constant temperature.
temperature than the CoO oxidation.
Despite the relatively large volume of the storage module
The CTCS concept is independent of the reactor/heat exchanger
(120 m3) its’ full load discharge period is limited to 1.5 h. To
design and can be equally well applied to all possible heat exchan-
extend the storage capacity and achieve round-the-clock power
ger concepts like packed beds or honeycomb/foam-based ones and
supply, rather than further augmenting the volume, the volumetric
to any combination of them. CTCS can be also combined with
energy storage density should be increased. As discussed in the
porosity variation along and across the storage module: for exam-
previous parts of this publication series, regenerative sensible heat
ple honeycomb or foam structures with varying cross–section
storage can be easily ‘‘hybridized” with TCS with redox oxides: in
porosity can be distributed along the cascade (rightmost sketch
addition to sensible-only-effects the heat effects of chemical reac-
of Fig. 1d) to homogenize and ‘‘smooth” the air velocity profile
tions can be exploited within the same storage module volume
achieving thus more uniform heat transfer. Furthermore it can be
increasing thus the volumetric energy storage density. In this per-
combined with different (non-thermochemical) modes of solar
spective, the idea of exploiting redox oxide pairs coated on or
heat storage, like sensible or latent heat, in a configuration where
shaped into porous structures like honeycombs and foams, as
particular stages will employ for example only sensible and others
hybrid sensible-thermochemical solar energy storage means in
hybrid sensible-thermochemical or sensible-latent heat storage
air-operated STTPs was set forth, having currently reached the
(Zanganeh et al., 2015).
level of small-size redox-oxide coated honeycombs and foams
To further advance these concepts, the present work involved
(Agrafiotis et al., 2015b,c; Tescari et al., 2014) as well as larger-
the preparation of lab-scale (£ 25 mm), Co3O4- and Mn2O3-
scale porous structures made entirely of Co3O4 (Karagiannakis
coated, porous cordierite honeycombs and foams and their testing
et al., 2014, 2016; Pagkoura et al., 2015, 2014). This concept of
in a one- or two-oxides cascaded configuration. Studies were per-
exploiting multi-channeled porous ceramics to enhance heat
formed to identify the effects of operational parameters on oxygen
transfer between a heat transfer fluid and a storage medium has
release/uptake and the thermal effects of the reactions on the air
been adopted later in similar ‘‘hybrid” sensible-latent heat storage
stream exit temperature as well as to compare such hybrid sensi-
studies where bricks (de Gracia and Cabeza, 2015) or SiC honey-
ble/thermochemical storage systems vs. respective sensible-only
combs (Li et al., 2015) have been combined with Phase Change
ones.
Materials (PCMs). In parallel, other TCS reactor concepts in CSP
plants are investigated: a recent simulation study (Ströhle et al.,
2016) has addressed TCS with packed and fluidized beds as storage 2. Experimental
media using the Mn2O3/Mn3O4 pair as a model system; ‘‘moving
constant-temperature profiles” similar to the ones of Fig. 1c were 2.1. Materials
shown to describe the operation of such a storage module as well.
However, due to the thermocline temperature distribution The same single oxides of Co3O4 and Mn2O3 from Materion
within the regenerative storage block, if the entire honeycomb Advanced Chemicals as in Part IV of this series were employed.
inventory was coated with a single oxide, the two TCS steps (charg- Cordierite honeycombs of 400 cells per square inch/cpsi (Corning
ing and discharging) would cease to take place in regions where S.A., U.S.A.) and foams of 20 and 30 pores per linear inch/ppi
and when the temperature would drop below the oxide’s reduc- (Ceramiques Techniques et Industrielles, France) as well as of 45
tion/oxidation temperature, respectively. To address such issues, ppi (Fraunhofer-IKTS, Germany), all of 25.4 mm (1 in.) diameter
the concept of Cascaded ThermoChemical Storage (CTCS) were used as porous supports. Their physical properties were pro-
(Agrafiotis and Pitz-Paal, 2013) has been proposed where various vided in Part II of this work (Agrafiotis et al., 2015b); it was also
oxides can be cascaded with decreasing reduction temperature shown therein by XRD analysis after cyclic redox operation of
from the ‘‘hot” to the ‘‘cold” end. Among four such single oxides Co3O4-coated small-size cordierite specimens, that no formation
investigated by the authors in previous works (Agrafiotis et al., of mixed phases between cobalt oxide and cordierite has taken
2015a, 2016a,b) only the Co3O4/CoO and Mn2O3/Mn3O4 pairs were place, justifying the current selection.
capable of reproducible, close-to-stoichiometric, cyclic operation.
Whilst the reaction schemes and thermodynamic calculations for 2.2. Coating of the ceramic structures
a particular redox pair assume one, common ‘‘equilibrium” tem-
perature for reduction and oxidation, this is not the case in prac- Series of such honeycombs and foams were coated with the two
tice: not only there might be some hysteresis phenomena during powders via the approach known with the name ‘‘washcoating”.
operation that induce small differences of the order of few tens This technique, described in details in previous publications
of degrees between reduction and oxidation, but the Mn2O3/ (Agrafiotis and Tsetsekou, 2000a,b) involves impregnation of the
Mn3O4 pair is characterized by a difference between reduction supports in properly prepared low-viscosity, high-solids-content
and oxidation temperature of the order of 100–200 °C. Based on powder slurries and subsequent drying and calcination. The details
the reduction/oxidation temperature ranges of these systems as for the preparation of the Co3O4 slurry were identical to those
shown in the schemes (1) and (2) below, a two-oxides cascade, already reported (Agrafiotis et al., 2015b). The same dispersant
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 679

(DOLAPIX PC75Ò, Zschimmer & Schwarz GmbH & Co KG Chemisc- placed inside a tubular furnace (Gero GmbH & Co. KG, Germany)
she Fabriken, Germany) was employed to stabilize the Mn2O3 of maximum operating temperature 1250 °C programmed through
slurry and shift its pH from its’ ‘‘natural” value of 4.7 which was a Eurotherm 3200 programmer/controller. As shown in Fig. 2, the
very close to its iso-electric point (IEP) reported in the literature heated part of the tube can accommodate cascades of non-coated
(Kosmulski, 2009), to 7.7. Indeed, low-viscosity slurry was or redox-oxide-coated, 25 mm-diameter honeycombs or foams for
formed and no settling was observed during coating. After drying testing their sensible as well as thermochemical storage perfor-
for 12 h at 120 °C, the coated ceramics were calcined in a furnace mance or powder samples contained in ceramic boats. Both ends
at 800 °C for 4 h heated with a rate of 2 °C/min to remove the of the reactor tube were closed with metal seals having ports for
organic dispersant and induce adequate adhesion of the redox gas inlet/outlet as well as for thermocouples that could be slided
coating on the supports. The temperature of 800 °C was selected in and out. The temperature at specific points along the reactor
so that not to exceed the reduction temperature of either oxide. was measured with S-type thermocouples placed at positions corre-
The mass of oxide coated on the ceramic supports, calculated by sponding as shown in the exemplary sketches of Fig. 2d. The oxygen
weighting the dry objects before coating and after calcination, is concentration at the reactor’s exit stream was monitored by an in-
reported in Table 1. A typical set of such supports spanning the size line oxygen sensor (Sensor Technics Model XYA5M, measuring
range from very small ones suitable for the Thermogravimetric accuracy ± 5 mbar, sampling frequency of 0.95 Hz). The air flow rate
Analyzer/TGA, to £ 25 mm ones suitable for the current test rig was controlled by a mass flow controller (MKS INSTRUMENTS).
and to £ 90 mm suitable for a solar receiver (Agrafiotis et al.,
2016a) before and after being coated with Mn2O3 are shown in 2.4. Experiments with oxides in powder form
Fig. 1e and f respectively.
At first Co3O4 and Mn2O3 were tested in powder form, sepa-
2.3. Experimental setup for sensible-thermochemical storage rately and in combination, placed in Al2O3 boats in a single- or
experiments double-boat cascade configuration, as shown in the photo in Fig. 2c
and the sketches in Fig. 2d. All such experiments were performed
The experimental evaluation of thermochemical/sensible heat with an air flow rate of 80 sccm. The experimental protocols were
storage capacity was performed in a dedicated test rig consisting based on the results with these powders already reported in Parts
of an alumina ceramic tube (65 cm long, inside diameter 27 mm) I–IV of this series. Since oxidation of Mn2O3 takes place with a

Table 1
Amount of redox powder coated on the various supports and respective loading percentages.

Co3O4 – coating
Code Weight After 1st calcination After 2nd calcination
Specimen Plain (g) Total weight (g) Total weight gain (g) Total weight gain (%) Total weight (g) Total weight gain (g) Total weight gain (%)
Cordierite honeycombs 400 cpsi, L = 50 mm, £ 25 mm (Corning)
1 11.85 18.98 7.13 60.2 34.61 22.76 192.1
2 12.40 23.18 10.78 86.9 38.70 26.30 212.1
3 11.86 18.77 6.91 58.3 38.25 26.39 222.5
4 11.70 18.94 7.24 61.9 33.84 22.14 189.2
Sum 32.06 97.59

Cordierite foams 20 ppi, L = 25 mm, £ 25 mm (CTI)


1 9.01 13.16 4.15 46.0 23.11 14.10 156.5
2 8.74 13.04 4.30 49.2 22.50 13.76 157.4
3 8.46 13.03 4.57 54.0 26.67 18.21 215.2
4 9.00 13.00 4.00 44.4 25.01 16.01 177.9
Sum 17.02 62.08

Cordierite foams 30 ppi, L = 25 mm, £ 25 mm (CTI)


1 10.13 16.27 6.14 60.6 30.05 19.92 196.7
2 11.43 18.16 6.73 58.9 31.95 20.52 179.5
3 11.59 17.12 5.53 47.7 28.44 16.85 145.4
4 10.51 16.10 5.59 53.2 27.65 17.14 163.1
Sum 23.99 74.43
Sum of 8 foams 41.01 136.51

Mn2O3 – coating
Cordierite honeycombs 400 cpsi, L = 50 mm, £ 25 mm, (Corning)
1 9.69 15.23 5.54 57.2 27.42 17.73 183.0
2 8.60 11.46 2.85 33.1 20.47 11.87 137.9
3 6.46 13.47 7.00 108.4 27.03 20.56 318.2
4 6.48 8.81 2.33 36.0 20.26 13.78 212.7
5 6.20 11.54 5.33 86.0 21.73 15.50 250.0
6 6.20 9.66 3.45 55.6 16.84 10.61 171.0

Cordierite foams 45 ppi, L = 50 mm, £ 25 mm, (IKTS)


1 4.99 8.84 3.85 77.2 18.58 13.59 272.5
2 5.22 9.29 4.06 77.8 21.34 16.11 308.5
3 6.12 10.50 4.39 71.8 23.73 17.61 288.0
4 4.97 9.59 4.62 92.9 19.93 14.96 300.9
680 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

Fig. 2. Thermochemical/sensible storage capacity test rig: (a) furnace; (b) non-coated cordierite honeycombs ‘‘wrapped” with YSZ sheet before placed in the reactor; (c)
alumina reactor tube and representative configurations that were tested (from top to bottom): cascade of six non-coated honeycombs, of four redox-oxide–coated
honeycombs, of two boats containing loose redox oxide powder(s), of eight small-length foams before and after their coating with Co3O4 and of three redox-oxide–coated
honeycombs; (d) schematic of thermocouples placement in the cases of testing cascades of porous supports and one or two boats.

slower rate and at much lower temperatures than its reduction, to favorable for oxidation (Agrafiotis et al., 2016a,b). Experiments of
ensure completion of oxidation either long dwell times at suitable the first type have been reported previously (Agrafiotis et al.,
low temperatures or slow ramp rates had to be selected so that the 2015a); this work includes primarily experiments of the second
material remains for a long period within the temperature range type.
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 681

2.5. Storage experiments with redox-oxide-coated and non-coated tion/consumption curves from cycle to cycle. In contrast to the
porous supports respective TGA experiments where the weight change of the sam-
ple is recorded within a very short time, oxygen detection takes
Subsequently, experiments with non-coated and redox-oxide- longer and under the particular ramp rate the transition of the oxy-
coated porous supports (honeycombs and foams) were performed gen concentration from reduction to oxidation looks ‘‘quasi-
to compare sensible-only to hybrid sensible-thermochemical heat continuous” and not characterized by a plateau of oxygen concen-
storage under a variety of experimental conditions. For all the tests tration at 21% as is the transition from oxidation to reduction due
involving porous specimens the supports were wrapped with a to the higher temperature range spanned in the latter case.
thin sheet of Zirconium oxide, high-temperature felt insulation Next experiments involved separate tests with Co3O4 and
(ZYF-100, Zircar Zirconia, Inc., U.S.A.) that acts simultaneously as Mn2O3 powders under temperature schedules to ensure reduction
a flow sealant, as an insulating layer and prevents any interaction of both powders. The first such experiments involved one heating
between the redox-oxide-coated specimens and the alumina tube cycle with 5 °C/min to 1000 °C to exceed the reduction tempera-
(Fig. 2b). To exploit the whole heating length of the furnace for ture of both oxides and a dwell of 2 h at this temperature to ensure
the testing of honeycombs, six of them (50 mm-long each) were completion of reduction, before letting then the system cool down
placed in the reactor tube. For the tests of the foams, shorter with the ‘‘natural” rate. For the case of Co3O4 the temperature and
(25 mm-long each) 20 and 30 ppi structures were set together in oxygen concentration evolution curves during two such experi-
a cascade alternating one another with each of them used four ments with quantities of 8 and 10 g are also shown in Fig. 3a, to
times (Fig. 2c). In all cases, the two honeycombs placed on either delineate the effect of Co3O4 mass on the peak height and area.
end of the cascade were never coated but were employed to make The comparative performances of the Co3O4 and Mn2O3 powders
it possible to place thermocouples through their channels inside under exactly the same temperature schedule are shown in Fig. 3b.
the storage system (Fig. 2c and d). Whereas the reduction of the two powders takes place in the same
The tests were performed with the same heating schedules as in temperature range and the behaviour of Co3O4 is fully reversible
the prior tests with powders, however in this case a wide range of upon cooling, the ‘‘natural” cooldown rate (5 °C/min) is too high
air flow rates from 80 to 10,000 sccm was tested. As already men- for Mn3O4 oxidation to take place and no oxygen is recorded to be
tioned, the temperature in the system must reach at least 950 °C uptaken from the reduced powder. Other than that, the shape of
for both the oxidized states of the thermochemical storage oxide the two oxygen release curves is similar, characterized by a sharp
materials to be reduced. However, in contrast to experiments in initial increase followed by a parabolic-like drop, similar to that of
a TGA, the possible reaction zone in the furnace configuration is such oxygen release curves of redox materials used for water/car-
much longer and not all of it is subjected to the same temperature; bon dioxide splitting obtained during thermal reduction (Agrafiotis
even when a fixed temperature set-point is programmed in the fur- et al., 2013; Kostoglou et al., 2014).
nace controller, a non-uniform temperature profile is established In an effort to induce oxidation of Mn3O4 under cycling condi-
along the reactor with higher temperatures near the center of tions, a slower ramp rate of 2 °C/min was tested, within 4 temper-
the heated zone and lower temperatures near both ends. In fact, ature cycles between 1090 and 640 °C and without dwell between
in these experiments the actual temperature reached in each hon- the steps. The last cooling step was performed with 2 °C/min until
eycomb depends on the air flow rate selected. For example, as dis- 640 °C and with the ‘‘natural” cool-down rate thereafter until
cussed later in the text, high flow rates might cool a significant part ambient temperature. The results are shown in Fig. 4a (full
of the entrance length of the cascade to temperature levels below 4-cycle experiments) and b (‘‘magnification” of the first cycle).
those needed to induce reduction; the redox material in this region The absolute value of the O2 curve of the Mn2O3-only experiment
will be practically non-available for thermochemical operation. corresponds to the scale on the right axis; the other two O2 curves
One of the goals of this experimental series was therefore to deter- have been ‘‘shifted” lower for reasons of clarity. Since the oxygen
mine the effect of air flow rate on the temperature distribution of absorbed/released per g of Co3O4 is much higher than that per g
the cascade along the reactor in comparison to the reduction tem- of Mn2O3, different quantities of each material were employed
perature of each particular system. Thereafter, experiments were (as mentioned on the graphs) to ‘‘produce” peaks of approximately
performed with the hybrid sensible-TCS setup with the honey- the same height. Under these conditions reduction and oxidation
combs and foams after their first as well as their last coating at take place for both materials with a clear, repeatable way both in
specific air flow rates to investigate the effect of redox material their single tests as well as when they are together. The reduction
employed. Furthermore, the system was tested with an empty peaks for both materials start to appear upon heating, at around
reactor tube. These tests can be viewed as blank tests under ‘‘no 900–920 °C. During cool-down, oxidation of CoO takes place first,
storage conditions”, either sensible or thermochemical, for com- around 900 °C. However, any oxygen uptake due to oxidation of
parison. Finally, based on the results of the tests above, experi- Mn3O4 is not recorded during further cool-down, but only when
ments comparing Co3O4-only, Mn2O3-only and mixed honeycomb the system is heated up again, around 630 °C (Fig. 4b), lasting until
cascades were performed. a temperature of 830 °C is reached. This observation is in accor-
dance with relevant TGA experiments under a similar schedule
reported in Part IV of the series (Agrafiotis et al., 2016b) where
3. Results – discussion the major weight gain of the reduced Mn3O4 powder due to re-
oxidation was recorded upon heat-up from the lower temperature
3.1. Experiments with Co3O4 and Mn2O3 oxides in powder form limit.

The first sets of experiments were performed with boats con- 3.2. Hybrid sensible-thermochemical storage experiments with Co3O4-
taining only Co3O4, based on its redox behaviour when tested in coated porous ceramic supports
the TGA, where complete reduction-oxidation was observed during
cycling between 985 and 785 °C irrespective of the ramp rate. Such The first such experiments were exploratory targeted on identi-
an experiment was performed in the current test rig with a boat fying conditions (with respect to air flow rate, amount of coated
containing 3 g of Co3O4 under a ramp rate of 5 °C/min. The oxygen material etc.) where the thermochemical reactions could induce
traces during test (black curves in Fig. 3a) demonstrate cyclic measurable effects on the oxygen concentration and, most impor-
reduction-oxidation of Co3O4 and reproducible oxygen produc- tantly, on the temperature of the air stream flowing through the
682 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

1200 40
Co3O4

T=990oC Air flow rate = 80 sccm; Ramp rate:5oC/min


1000 mCo O = 3g
3 4

mCo O = 8g
30
3 4

O2 concentration (% in air)
mCo O = 10g
3 4

T=790oC Temperature
800 O2 concentration
Temperature (oC)

600 20

400

10

200

0 0
(a) 0 200 400 600 800 1000
Time (min)

1200 40
Air flow rate = 80 sccm
Tplateau: 1000oC; Dwell time=2 hrs
Heating rate:5oC/min
1000 mCo O = 8 g; mMn O = 9 g
3 4 2 3

O2 concentration (% in air)
30
Temperature (oC)

mOxygen : ~ 70 %
800
of theoretical

600 20

mOxygen : ~ 71 %
Temperature
of theoretical
400 Co3O4
Mn2O3

mOxygen : ~ 86 % O2 concentration 10
Co3O4
of theoretical
200 Mn2O3

(b) 0 0
0 200 400 600 800 1000
Time (min)

Fig. 3. Temperature and O2 concentration evolution curves during experiments with boats containing only Co3O4 or only Mn2O3 powder under an air flow rate of 80 sccm and
heating rate of 5 °C/min: (a) experiments with Co3O4, either cycled between 990 and 790 °C without dwell (black curves) or heated up to 1000 °C, dwell for 2 h and ‘‘natural”
cooldown (blue and red curves); (b) comparison between Co3O4 and Mn2O3 under the latter heating schedule and calculation of conversion extent of thermochemical
reactions from integration of respective O2 concentration curves. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

coated supports. Due to the higher intensity of thermochemical two oxygen curves exhibit nearly the same profile during reduc-
effects of Co3O4 vs. that of Mn2O3, all such experiments were per- tion and oxidation. At higher flow rates the peaks become smaller
formed with Co3O4–coated supports: with cascades of 4 honey- since the same amount of oxygen is produced/consumed per g of
combs (50 mm-long each) and 8 foams (25 mm-long each) redox material but this quantity has a smaller effect on the total
coated only once and both having a total length of 20 cm as per concentration of oxygen in the air stream. The heat effect of reduc-
the photographs in Fig. 2c. This set of experiments was performed tion can be distinguished qualitatively by the small ‘‘deflections” in
under the first schedule above employed with powders, i.e. heating the temperature profile during heat-up: during the time when
with 5 °C/min up to a programmed temperature set point of reduction takes place the increase of temperature is delayed since
1070 °C, dwell for 2 h and then ‘‘naturally” cool-down to ambient heat is consumed in the endothermic reduction reaction and not to
temperature. Low air flow rates, between 30 and 370 sccm were sensibly heat the gas. As soon as reduction is complete the air
selected at first to better visualize any change in oxygen concentra- stream temperature rises again. However during cool-down, the
tion in the air stream due to the thermochemical reactions taking heat generated by the oxidation reaction is not enough to induce
place. a respective clearly visible change in the temperature curve for nei-
Fig. 5a shows the results with respect to O2 concentration and ther of the two flow rates. It should be noted that the apparently
temperature at the end of the reaction zone (i.e. at the end of the low ‘‘reduction” temperature (770 °C) where abrupt oxygen con-
last coated support) for the cascade of 8 foams after their first centration change is recorded (in this as well as in all other subse-
Co3O4–coating under two air flow rates, 180 and 280 sccm. The quent experiments in a similar configuration) is misleading: this
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 683

Fig. 4. Temperature and O2 concentration evolution curves during experiments with the same configuration as in Fig. 3, under a schedule of 4 temperature cycles between
1090 and 640 °C, with a ramp rate of 2 °C/min, without dwell between the steps and ‘‘natural” final cooldown: (a) entire experiments; (b) ‘‘magnification” of the first cycle.

temperature is measured at the end of the reaction zone whereas several experiments. Just like in the case of foams, the heat effect
the oxygen concentration change is due to reduction of redox of reduction is distinguished by the small ‘‘bend” in the tempera-
material being already at higher temperatures within the ‘‘central” ture profile during heat-up, but any temperature increase due to
areas of the reaction zone. Fig. 5b shows the respective results for oxidation taking place during cool-down is very low and not
the case of a cascade of 4 honeycombs after their first Co3O4–coat- detectable in any of the temperature curves during cool-down that
ing for five air flow rates between 30 and 370 sccm. The height and practically coincide.
area of the oxygen peak during reduction is again inversely propor- In all previous experiments with the Co3O4-coated structures
tional to the air flow rate. It is interesting to note that for the par- the oxygen concentration effects during oxidation were clearly
ticular mass of Co3O4 coated, air flow rates lower than 180 sccm do depicted; however any respective temperature effects from the
not provide enough oxygen for the oxidation reaction to fully take enthalpy of the oxidation reaction could not be visualized as a
place and therefore during cool-down oxygen concentration drops recordable increase of the air stream exit temperature. Even
to zero: all oxygen available in the flowing air stream is consumed though the coated material reacts, the heat produced by its oxida-
by the reduced oxide CoO but still is not enough to oxidize it com- tion is not transferred to the flowing air stream but is obviously
pletely. As a result, since the experiments took place consecutively dissipated and lost through the reactor walls. To increase the pos-
without taking the honeycombs out of the furnace but by simply sibility of visualizing the heat effects during oxidation, the honey-
changing the flow rate from experiment to experiment, an oxygen combs and foams were further coated with Co3O4 with consecutive
peak downwards, corresponding to oxidation of the reduced oxide impregnations in the slurry to a much higher final total loading (to
that was not completely oxidized in the previous experiment just generate more heat per unit volume). Subsequently similar exper-
before the peak upwards corresponding to reduction, is visible in iments were performed but employing much higher air flow rates
684 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

Fig. 5. Oxygen concentration evolution and temperature at the end of the reaction zone, i.e. at the end of the last coated support, during experiments under heating with 5 °C/
min up to 1000 °C, dwell for 2.5 h and ‘‘natural” cool-down to ambient temperature: (a) cascade of 8 foams after their first Co3O4–coating under air flow rates of 180 and 280
sccm; (b) cascade of 4 honeycombs after their first Co3O4–coating for five air flow rates between 30 and 370 sccm.

to enhance the transport of the heat produced by the reaction perature profiles shown in Figs. 6b and 7b: at higher flow rates the
enthalpy by convection along the reactor towards the exit of the characteristic ‘‘deflections” during reduction become much more
cascade, before this heat is dissipated. pronounced and a short, but clearly visible ‘‘quasi-isothermal”
The results of these experiments for the two systems, honey- range (‘‘quasi-horizontal” line segment) appears where the tem-
combs and foams, are shown in Figs. 6 and 7 respectively. For both perature remains practically constant for as long as reaction takes
kinds of supports, the effects of increased coated mass (that has place. Upon reaction completion the air stream temperature rises
been approximately tripled) and increased air flow rate and conse- again abruptly to follow the ‘‘straight” line of linear increase with
quently heat transfer by convection, can be clearly observed on the time (exemplary dotted blue line segment). Furthermore, such
temperature profiles comparing Figs. 6a and 7a vs. Fig. 5. At lower quasi-isothermal segments appear now clearly also during the
flow rates the plateau temperature reached during dwell is practi- duration of oxidation in the cooldown step. With increased quan-
cally the same for all flow rates. However beyond a threshold value tity of redox material within the same volume and enhanced con-
of air flow rate the plateau temperature achieved during dwell vection the ‘‘extra” heat produced by oxidation can be transferred
increases. In other words, higher flow rates enhance heat transfer more effectively downstream the reaction zone. These phenomena
by convection and thus the point at the end of the reaction zone were common to both foams and honeycombs. In conclusion, the
reaches higher temperatures (for the same furnace temperature respective ‘‘plateau” segments of the air stream exit temperature
set-point programmed). In addition, the heat effects of the chemi- demonstrate qualitatively the effects of chemical reaction
cal reactions become evident from the ‘‘magnifications” of the tem- enthalpies and the principle of thermochemical storage operation,
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 685

Fig. 6. Experiments with a cascade of 8 foams after their final Co3O4–coating, under the same schedule as those in Fig. 5: (a) effects of air flow rate (range 180–5000 sccm) on
oxygen concentration evolution and temperature at the end of the reaction zone; (b) ‘‘magnification” of temperature profiles to delineate the thermal effects of chemical
reactions.

i.e. heat storage/release under constant temperature. These results as an air exit temperature higher than that of entrance during oxi-
are in full qualitative agreement with results from a very recent dation; instead the thermochemical effects are manifested as a
study with similar honeycomb structures, either extruded from constant temperature plateau at the exit of the reaction zone dur-
Co3O4-based compositions or Co3O4-coated cordierite ones ing both steps.
(Karagiannakis et al., 2016). The differences are due to the slightly
different experimental testing configuration employed. The 3.3. Sensible-only storage experiments with non-coated porous
authors employed per experiment only one redox oxide honey- ceramic supports
comb of length between 23 and 50 mm and of redox powder
weight between 16 and 31 g. Thus the whole honeycomb was prac- Based on the discussion above, a further issue of concern rele-
tically at the same temperature irrespective of the flow rate, but vant to the particular cascade configuration of coated porous spec-
the amount of redox material and consequently of the enthalpy imens had to do with whether all the quantity of the redox oxide
released/absorbed is small. Therefore, in their study the thermo- loaded on the honeycombs was above the reduction/oxidation
chemical operation of Co3O4 was manifested as a slower increase temperature and was therefore available for thermochemical reac-
of the temperature at the exit of the honeycomb when oxygen tion. To clarify this issue, experiments in sensible-only storage
was produced during heating and as an air exit temperature higher mode (i.e. with non-coated honeycombs) were performed varying
than that of entrance when oxygen was uptaken during cooling. In the flow rate and measuring the temperature along the reactor, to
our study a much higher amount of redox material is employed but determine what extend of the cascade is above the reduction tem-
the length of the redox zone is not under the same temperature peratures of the two oxides examined herein, Mn2O3 and Co3O4,
thus the thermochemical operation of Co3O4 cannot be manifested 950 °C and 890 °C respectively. Characteristic temperatures at four
686 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

(a) 1200 60
4 Coated honeycombs, higher flow rates, higher loadings

Temperature at reaction zone end

Temperature at reaction zone end (C°)


5000 sccm
1000 2500 sccm 50
1000 sccm

O2 concentration (% in air)
370 sccm
Total Co3O4 coated mass = 98 g
O2 concentration
800 5000 sccm 40
2500 sccm
1000 sccm
370 sccm

600 30

400 20

200 10

0 0
0 200 400 600 800 1000
Time (min)

(b) 100
4 Coated honeycombs, higher flow rates, higher loading, total Co3O4 coated mass = 98 g

1100
Temperature at reaction zone end (C°)

O2 concentration (% in air)
75

1000

50
900

25
800

700 0
150 200 250 300 350 400 450
Time (min)

Fig. 7. Experiments with a cascade of 4 honeycombs after their final Co3O4–coating, under the same schedule as those in Fig. 5: (a) effects of air flow rate (range 370–5000
sccm) on oxygen concentration evolution and temperature at the end of the reaction zone; (b) ‘‘magnification” of temperature profiles to delineate the thermal effects of
chemical reactions.

different points along the cascade of six non-coated cordierite hon-  For even higher flow rates, 5000, 7500 and 10,000 sccm even
eycombs as per the sketch in Fig. 8a are shown in Fig. 8b compared the 3rd honeycomb becomes exposed (at least partially) at tem-
to these two reduction temperatures. The trends of the green and peratures lower than the reduction temperatures of both oxides
blue curves point out that increasing the air flow rate ‘‘cools” the and therefore not entirely available for a cascaded TCS concept
entrance of the cascade. The trends of the red and orange curves in this particular rig.
in the other hand, demonstrate that higher air flow rates enhance
From the respective experiments with an empty tube shown in
the transfer of heat produced both resistively by the furnace as
Fig. 8c the effect of sensible-only storage vs. no-storage becomes
well as due to the enthalpy of thermochemical reactions, towards
evident: for the same flow rate, a higher temperature is achieved
the end of the cascade. The following conclusions can be drawn for
both near the center as well as near the exit of the cascade when
the particular test rig configuration from this graph:
the tube is filled with porous honeycombs. This effect is better
delineated in the graph of Fig. 9 where the experiments at air flow
 For flow rates 280 and 380 sccm, 4 honeycombs - the 2nd, 3rd,
rate of 5000 sccm with the four coated honeycombs (hybrid
4th and 5th - are entirely above the reduction temperature of
sensible-thermochemical storage), the non-coated honeycombs
Co3O4 (890 °C) and marginally above that of Mn2O3 (950 °C).
(sensible-only storage) and the empty tube (no storage) are com-
 For higher flow rates1000 and 2500 sccm, three honeycombs –
pared. Even though the first of the four coated honeycombs is
the 3rd, 4th and 5th - are above the reduction temperature of
not entirely above the reduction temperature of Co3O4, the heat
Mn2O3.
effects of the reduction reaction during heat-up and of the
 Flow rates of 1000 and 2500 sccm seem to be the ‘‘optimal”,
oxidation reaction during cool-down are clearly qualitatively
where the temperatures at the entrance of the 3rd and at the
reflected on the temperature of the air stream exiting the reaction
exit of the 5th honeycomb are practically equal.
zone. Cooling of this stream is indeed retarded due to the
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 687

T=1170oC
T5 T2 T1

Air flow Air flow


out T6 in

(a) 14.5 cm 2.5 cm 5 cm 5 cm 5 cm 5 cm 5 cm 5 cm 3 cm 15 cm

1400
6 non-coated cordierite honeycombs 400 cpsi
End of 1st honeycomb; End of 2nd honeycomb; End of 5th honeycomb; End of 6th honeycomb

1200 80
sccm
180
sccm
280
sccm
380
sccm
500
sccm
1000
sccm
2500
sccm
5000
sccm
7500
sccm
10000
sccm

1000 Tred Mn O = 950oC


2 3

Tred Co O = 890oC
Temnperature (oC)

3 4

800

600

400

200

(b) 0
0 2000 4000 6000 8000 10000 12000
Time (min)

1400
Empty tube th
"End of 5 honeycomb";
th
"End of 6 honeycomb";
st
"End of 1 honeycomb"

1200 80 sccm 180 sccm 380 sccm 1000 sccm 2500 sccm 5000 sccm

T = 1045oC T = 1045 C
o
T = 1030oC T = 1020oC
1000 T = 992oC T = 990oC

T = 923oC
Temnperature (oC)

o
T = 895oC T = 900oC o
T = 906 C
T = 875 C

T = 860oC T = 830oC
800 o
T = 795 C

T = 750oC

600 T = 553oC

T = 384oC
400

200

(c) 0
0 1000 2000 3000 4000 5000 6000
Time (min)

Fig. 8. Characteristic temperature profiles at four different points along the reactor in experiments under the same schedule as those in Figs. 5–7: (a) schematic of
arrangement and measurement points; (b) experiments with a cascade of 6 non-coated cordierite honeycombs; (c) experiments with an empty tube.

thermochemical reaction, demonstrating thus the ‘‘proof-of- ule as those depicted in Figs. 5–7, i.e. involved heating with 5 °C/
concept” of the idea of TCS. min up to a programmed temperature set point of 1170 °C and
dwell for 2 h, but cool-down to ambient temperature was per-
3.4. Hybrid sensible-thermochemical storage experiments with Co3O4- formed with a slow, controllable rate of 2 °C/min (similar to that
and Mn2O3-coated porous ceramic supports in the experiments with powders depicted in Fig. 4) instead of nat-
ural cooling, to investigate whether this would have any favorable
The next set of experiments involved comparison among effect on the oxidation of Mn3O4.
Co3O4-, Mn2O3-coated and mixed honeycomb cascades at the The temperature and oxygen concentration evolution curves as
higher range of flow rates already tested. Based on the above, in a function of air flow rate are shown in Fig. 10a for the Co3O4- and
these experiments in one hand only three instead of four redox- 10b for the Mn2O3-loaded honeycombs. The peak temperatures
coated honeycombs were employed, as per the last photograph measured are listed in the graphs. Small discrepancies between
in Fig. 2c. The first experiments were according to the same sched- temperatures at the same location between experiments with
688 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

1200 50
4 Co3O4-coated honeycombs
Total Co3O4 coated mass = 98 g
Air flow rate = 5000 sccm

Temperature at reaction zone end (oC)


1000 Temperature at reaction zone end:
Coated honeycombs 40
Non-coated honeycombs
Empty tube

O2 concentration (% in air)
O2 concentration
800
Hybrid 30
Sensible-TC
Sensible-only
storage effect
600 Storage effect

20
400

10
200

0 0
0 200 400 600 800 1000
Time (min)

Fig. 9. Comparison of temperature at the end of the reaction zone among experiments with a cascade of 4 Co3O4–coated honeycombs (hybrid sensible-thermochemical
storage), non-coated honeycombs (sensible-only storage) and empty tube (no storage) under the same heating-cooling schedule as in Figs. 5–7 and air flow rate of 5000 sccm.

Co3O4- and Mn2O3-loaded honeycombs under the same air flow ‘‘first” at the entrance of the reaction zone, to ensure that the
rate are due to slight shifts of the thermocouples location. The tem- Mn2O3-coated ones will be exposed at higher temperatures.
perature trends are in full accordance with those depicted in Fig. 8; The effect of increasing flow rate on the oxygen concentration
increase of air flow rate renders temperature effects more visible trace is also evident. At higher flow rates the peak height dimin-
and oxygen concentration ones less pronounced. However, even ishes and the relevant curves, even though maintaining their qual-
though the Oxygen traces of the Co3O4-loaded honeycombs are itative characteristics, are characterized by a lot of ‘‘noise”. The
clear both during heat-up and cool-down, oxygen peaks from the oxygen evolution curves are compared for the case of the lower
Mn2O3-loaded ones are visible only during reduction (heat-up), flow rate tested, 500 sccm, where they are most intense, in
even at the lowest flow rate tested where the intensity of any such Fig. 12. The curve of the Co3O4-only experiment corresponds to
peaks was expected to be higher. The ‘‘magnification” of the tem- the absolute scale on the right axis and the other two have been
perature profiles at the exit of the reaction zone for the two sys- ‘‘shifted” to lower absolute values. In this cycling mode the oxida-
tems under the same flow rate of 2500 (Fig. 10c) shows not only tion reaction of Mn3O4 is evident, though not during the cool-down
the effects of Co3O4 reduction–oxidation on the ‘‘deflection” of step but during the next heat-up one, prior to reduction. In com-
the temperature curves in complete analogy with Figs. 6b and 7b, plete analogy to the experiments with powders depicted in
but also the absence of such effects in the case of Mn2O3. It is clear Fig. 4, the curve of the ‘‘mixed” Co3O4-Mn2O3 cascade is a superpo-
therefore that even this low cool-down rate does not suffice to sition of the other two, characterized by overlapping reductions
induce oxidation of Mn3O4 back to Mn2O3; this is also reflected and two distinct oxidations, of CoO during cool-down and of
on the appearance of oxygen peaks downwards during initial Mn3O4 essentially during heat-up.
heat-up in the experiments in Fig. 10b (which were actually per- To delineate the effects of redox oxide material and air flow rate
formed in a sequence opposite than that shown in the Figure, i.e. on temperature, the temperature profiles at the entrance and exit
from the highest to the lowest flow rate), corresponding to oxida- of the reaction zone corresponding to the first cycle of the three
tion of the material that was not oxidized during cool-down in the experiments depicted in Fig. 11, are shown in ‘‘magnification” in
previous experiment. Fig. 13a and b for the two cases of low (500 sccm) and higher
Thus the Mn2O3-loaded honeycombs were also further loaded (2500 sccm) air flow rates respectively. Even though the effects
with Mn2O3 to their final loadings listed in Table 1 in order to are of relatively low intensity, are at the same time qualitatively
increase the heat per unit volume produced and a final set of com- visible and distinct. It is evident that the intensity of any deflection
parative experiments among Co3O4-, Mn2O3-coated and mixed of the temperature curves is ‘‘proportional” to the amount of Co3O4
honeycomb cascades was performed at the higher range of flow present: during heat-up such deflection is obvious in the case of
rates, but this time in cyclic mode: under the same schedule of four Co3O4-only-coated honeycombs, present but less intense in the
(4) cycles between upper and lower temperature set points of 1070 ‘‘mixed” Co3O4-Mn2O3 cascade and inconspicuous in the case of
and 670 °C respectively, with a ramp rate of 2 °C/min and no dwell. Mn2O3-only cascade. It is interesting to note the effect of the Peclet
The respective curves of temperature and oxygen concentration number on the temperature profiles: in the case of low flow rates
evolution are shown in Fig. 11 (in separate plots for reasons of clar- where thermal diffusion is significant, the temperature profiles at
ity) for three characteristic cases: 3 Co3O4-coated honeycombs the entrance of the reaction zone (green curves) exhibit a deflec-
(Fig. 11a), 3 Mn2O3-coated honeycombs (Fig. 11b) and a cascade tion as well, whereas at higher flow rates convection limits this
of 1 Co3O4- and 2 Mn2O3-coated honeycombs (Fig. 11c). Just like effect to the end of the reaction zone (red curves). During cool-
in the previous experiments, concentration effects are enhanced down only in the case of high flow rate and Co3O4-only coating, a
in lower air flow rates and temperature at higher ones. Increasing minute but nevertheless clearly present deflection can be observed
the flow rate to 5000 sccm results in temperatures lower than that in the temperature at the exit of the cascade, shifting it to slightly
of the reduction temperature of Mn2O3 even at the first coated higher values (in the dotted box in Fig. 13b), in analogy to Fig. 9.
honeycomb; for this reason in the experiment with 1 Co3O4- and To summarize, even though the sequential chemical reaction
2 Mn2O3-coated honeycombs the Co3O4-coated one was placed part of a Co3O4-Mn2O3 cascade during oxidation was
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 689

Fig. 10. Temperature and oxygen concentration evolution profiles during experiments under heating with 5 °C/min up to a set-point of 1170 °C, dwell for 2.5 h and cool-
down with a rate of 2 °C/min to ambient temperature: (a) 3 Co3O4-coated honeycombs (mass of Co3O4 = 75 g); (b) 3 Mn2O3-coated honeycombs (mass of Mn2O3 = 15 g); (c)
‘‘magnification” of temperature profiles at reaction zone end for the experiments with air flow rate 2500 sccm.

demonstrated, it was not possible to materialize the thermal effect der grade and quantity and in the specific test rig configuration
of the Mn3O4 oxidation reaction into quantitative heat transfer employed in this study, in contrast to that of CoO. Even though
fluid temperature changes, at least with the particular Mn2O3 pow- the performance of the Mn2O3/Mn3O4 system and especially its
690 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

Fig. 11. Temperature and oxygen concentration evolution profiles during testing for four cycles between upper and lower temperature set points of 1070 and 670 °C
respectively, with a ramp rate of 2 °C/min and no dwell for: (a) 3 Co3O4-coated honeycombs (mass of Co3O4 = 75 g); (b) 3 Mn2O3-coated honeycombs (mass of Mn2O3 = 50 g);
(c) a cascade of 2 Co3O4- and 1 Mn2O3-coated honeycombs.

oxidation reaction was found to depend on the physical/morpho- above not only in studies involving TGA experiments (Carrillo
logical properties of the particular powder source (Carrillo et al., et al., 2014, 2015) but also in larger-scale packed powder bed test-
2014, 2015), the inherent peculiarities of this system mentioned ing rigs (Wokon et al., 2014) constitute a challenge. Obviously
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 691

Fig. 12. Comparison of oxygen concentration evolution profiles among the experiments of Fig. 10 under an air flow rate of 500 sccm (temperature profiles from experiment
with 3 Co3O4-coated honeycombs).

incorporation of this system in large-scale STPP operation needs is obvious that the area under the oxidation oxygen peak during
further elaboration of the operation strategy, especially since dur- cooling is higher than that under the sharp reduction peak during
ing discharging, the oxidation reactions of such redox-oxide-based the previous heat-up. What has really happened is that due to
thermochemical systems should occur upon their cooling (Fricker, shortage of oxygen during cooling in a previous experiment with
2004; Ströhle et al., 2016). For both systems, larger-scale, effi- another flow rate, a quantity of the oxide has remained in its
ciently insulated, multiple-honeycomb/foam systems with high reduced state; when the next experiment with a different air flow
thermal mass can perhaps allow full exploitation of any thermo- rate started, at first during initial heatup an oxidation of this quan-
chemical benefits. These issues are currently under consideration tity of the oxide was observed as an oxygen uptake peak down-
in next experimental campaigns targeted to testing of such cas- wards. However due to the fast heatup, not all the quantity of
caded systems in larger scale. this ‘‘leftover” reduced oxide has been oxidized before the reduc-
tion temperature is reached. Thus the quantity of oxygen evolving
3.5. Oxygen production/consumption beyond this stage corresponds to a mass of oxide being reduced
less than the stoichiometric one (since some quantity of the oxide
In all such experiments as above, from the known quantities of has been already remained reduced from the previous experi-
powder mass and air flow rate, the oxygen produced/consumed ment). Then during cooldown and oxidation the mass of oxygen
per mass of powder (e.g. in lmoles O2/g of powder) with respect uptaken is not only due to the oxidation of the reduced oxide dur-
to the theoretical one expected from a given mass of redox powder ing heat-up but in addition due to the oxidation of the reduced
according to fully stoichiometric reduction/oxidation reactions, oxide that is ‘‘carried” through the previous experiment, as a result
can be calculated by integrating the respective oxygen concentra- the area is higher than that from the peak upwards. This trend is
tion curves. In the first set of experiments with powders (Fig. 3), eliminated at much higher flow rates, where the areas under the
for both redox pairs Co3O4/CoO and Mn2O3/Mn3O4, this amount reduction curves are consistently higher than those under the oxi-
of O2 calculated from such integrations of the curves vs. the max- dation curve. However another problem was introduced: with air
imum one corresponding to stoichiometric reduction/oxidation of flow rates above a threshold value the coated supports in the
1 g of Co3O4 (2076 lmoles of O2) or Mn2O3 (1056 lmoles of O2) entrance of the cascade were cooled below the reduction temper-
during reduction ranged between 70% and 77% whereas that for ature of Co3O4 and thus the powder loaded in them was not avail-
CoO oxidation accounted for 84–86% of the maximum value corre- able for reduction. Since there was no way of knowing the mass of
sponding to stoichiometric reaction (Fig. 3b and Table 2, no reoxi- Co3O4 actually participating in the reaction so that to calculate the
dation of Mn3O4 demonstrated). theoretical number of (moles of O2/g of Co3O4) corresponding to
In the next set of experiments with 8 Co3O4-coated foams and 4 ‘‘stoichiometric” reduction/oxidation, calculation of a percentage
Co3O4-coated honeycombs heated with a constant ramp rate of with respect to the theoretical maximum was not possible either.
5 °C/min to 1070 °C, dwelling for 2 h and then cooled-down to In contrast, though, based on the above, conversion calculations
ambient temperature under different flow rates (Figs. 5–7) calcula- are not meaningless for several of the experiments of the 2nd type
tion of the relevant percentages with respect to the theoretical depicted in Fig. 11, i.e. testing of 3 honeycombs, for four cycles
ones from the known amount of Co3O4 loaded on the honey- between 1070 and 670 °C with a ramp rate of 2 °C/min and no
combs/foams and integration of oxygen curves like in the case of dwell: in particular for these experiments where temperature
the experiments with powders, was problematic. At the lower measurements show that all three oxide-coated honeycombs (with
range of air flow rates (30–370 sccm) as already mentioned, the a known total mass of redox oxide) are above the reduction tem-
experiments took place consecutively without taking the honey- perature (dotted red line) throughout the experiment. The most
combs out of the furnace rig but by simply changing the flow rate characteristic such experiments are these with 3 Co3O4-coated
from experiment to experiment. In several of these experiments it honeycombs (with a known total mass of Co3O4 = 75 g) under air
692 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

1150
o
3 Co3O4- vs. 3 Mn2O3- vs. 2 Mn2O3+1 Co3O4 honeycombs, 2 C/min , 500 sccm
nd th
Co3O4: T end of 2 , T end of 5
1100 Co3O4+ Mn2O3: T end of 2 ,
nd
T end of 5
th

nd th
Mn2O3: T end of 2 , T end of 5

1050

Temperature ( C)
o

1000

o
Tred Mn O = 950 C
950
2 3

900

850

(a) 800
350 400 450 500 550 600 650 700 750
Time (min)

1150
o
3 Co3O4- vs. 3 Mn2O3- vs. 2 Mn2O3+1 Co3O4 honeycombs, 2 C/min, 2500 sccm

nd th
Co3O4: T end of 2 , T end of 5
1100 Co3O4+ Mn2O3:
nd
T end of 2 , T end of 5
th

nd th
Mn2O3: T end of 2 , T end of 5

1050
Temperature ( C)

1000
o

o
Tred Mn O = 950 C
950
2 3

900

850

(b)
800
5350 5400 5450 5500 5550 5600 5650 5700 5750
Time (min)

Fig. 13. Comparison of temperature evolution profiles during the first cycle among the experiments of Fig. 11 under an air flow rate of (a) 500 sccm; (b) 2500 sccm.

flow rates of 500, 1000 and 2500 sccm (gray, blue and brown TGA ones; first in studies of our group employing solar-irradiated
curves in Fig. 11a, respectively). These calculations with respect Co3O4 powders in a rotary kiln (Neises et al., 2012) where they
to lmoles O2 released/consumed per g of Co3O4 vs. the maximum were attributed to partial reaction of the solid under the particular
one corresponding to stoichiometric reduction/oxidation of 1 g of experimental conditions. As already mentioned, similar values not
Co3O4 (2076 lmoles of O2) are shown in Table 2. Even being exceeding 60% of theoretical were reported by another research
approximate, the values of Table 2 show a definite trend: they group for Co3O4- and Mn2O3-made porous pellets of mass in the
are higher in case of small quantities of powders, and with loaded range of 8–10 g (Karagiannakis et al., 2014; Pagkoura et al., 2014)
porous structures show a tendency for reduction with increasing in test rigs similar to the one tested in this study as well as in
flow rate. The highest among the values of coated honeycombs the latest work of this group with Co3O4-based extruded and
for an air flow rate of 500 sccm (1086–1354 lmoles of O2/g of coated honeycombs. Since in all these studies similar oxygen sen-
Co3O4) lie in exactly the same range with those reported recently sors were employed for the oxygen measurements, there is a pos-
(Karagiannakis et al., 2016) for a variety of Co3O4-based porous sibility that this consistent lower-than-stoichiometric conversions
honeycombs (1090–1210 lmoles of O2/g of Co3O4). could be due to the particular way of measurement which is char-
All the values of lmoles of O2 produced/consumed per g of acterized by a high ‘‘noise” level as well (Fig. 11); however in the
redox oxide are lower than the theoretically maximum ones. It is last two references above the sensors’ signal was cross-checked
worth mentioning that such lower-than-stoichiometry values have vs. that of other measurement techniques like oxygen analyzers
been reported consistently in the literature in studies other than and no discrepancy was found. Thus the difference between TGA
C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694 693

Table 2
Calculated amount of oxygen evolved/uptaken per gram of redox powder as percentage of the theoretical from complete stoichiometric reduction/oxidation reactions, occurring
from integration of the respective oxygen concentration curves in experiments with powders and with Co3O4-coated supports.

Powder Mass (g) Air flow rate (sccm) Dwell (h) Oxygen
Reduction Oxidation
lmoles O2/g oxide (%) Of stoichiometric lmoles O2/g oxide (%) Of stoichiometric
Powders, 1 cycle, dwell (Fig. 3)
Co3O4 8 80 2 1477 71 1794 86
Co3O4 10 80 2 1614 77 1757 84
Mn2O3 9 80 2 730 70 X X

Peak no Air flow rate (sccm)


500 1000 2500
lmoles O2 g Co3O4 (%) Of stoichiometric lmoles O2/g Co3O4 (%) Of stoichiometric lmoles O2/g Co3O4 (%) Of stoichiometric
3 Co3O4 – coated honeycombs (m Co3O4 = 75 g), 4 cycles, no dwell (Fig. 11a)
1 (red) 1145 55 827 40 625 30
2 (ox) 1107 53 1112 53 877 42
3 (red) 1210 58 839 40 729 35
4 (ox) 1086 52 1119 54 892 43
5 (red) 1294 62 863 42 536 26
6 (ox) 1110 53 1130 54 833 40
7 (red) 1354 65 880 42 565 27
8 (ox) 1148 55 1130 54 788 38

results that demonstrate stoichiometric O2 production/consump- The operating conditions of a cascade based on these two oxides
tion for a variety of conditions and redox-oxide-coated and –made are dictated by the redox behaviour of the Mn2O3/Mn3O4 pair since
porous supports for the two systems has to be attributed else- this system is characterized by a higher reduction temperature,
where. Future experiments with porous objects of varying porosity lower oxygen uptake/release and reaction heat effects, slower oxi-
(powders, granules, pellets, honeycombs and foams) are scheduled dation rate and a large temperature gap between reduction and
to investigate this issue. oxidation temperature. Under properly selected conditions the
two oxide systems can operate in a complementary mode. During
heating their reduction reactions overlap whereas during cooling
4. Conclusions
and cycling the oxidation of the reduced state of the oxides takes
place in a consecutive way: that of CoO during cooling and that
The target of this work was to advance further the concept of
of Mn3O4 almost exclusively during re-heating, in different tem-
regenerative ‘‘hybrid” sensible/thermochemical storage of solar
perature ranges. However, even though this proof-of-concept of
energy proposed for air-operated Solar Thermal Power Plants,
the cascaded configuration was in principle demonstrated, any
exploring the concept of Cascaded ThermoChemical Storage
beneficial thermal effects of the Mn2O3/Mn3O4 pair system were
(CTCS). In this approach, cascades of porous structures incorporat-
not evident, at least in the present configuration tested. The pecu-
ing different redox oxide materials are rationally distributed along
liarities of this system need further elaboration of the operation
the storage module to maximize its energy storage density.
strategy in combination with larger-scale, properly insulated sys-
In previous works the Co3O4/CoO and Mn2O3/Mn3O4 redox pairs
tems to allow full exploitation of any thermochemical benefits.
have been identified as the most promising single-oxide systems.
In parallel, small-scale Co3O4-coated porous honeycombs and
foams have demonstrated stoichiometric, reversible, cyclic redox Acknowledgements
operation up to 100 cycles in a thermogravimetric analyzer. Fol-
lowing up these facts, the present work involved the preparation Dr. C. Agrafiotis would like to thank the European Commission
of larger, lab-scale (£ 25 mm), Co3O4- and Mn2O3-coated porous for funding of this work within the Project ‘‘FP7-PEOPLE-2011-IEF,
cordierite honeycombs and foams and their testing in one- or 300194; Thermochemical storage of solar heat via advanced reac-
two-oxides cascaded configurations. Such structures could be tors/heat exchangers based on ceramic foams (STOLARFOAM)”
coated with a high amount of redox oxides, exceeding 100% and under the Intra-European Fellowship Marie Curie Actions of the
in some cases reaching even 300% of their initial weight. 7th Framework Programme and to DLR Programmdirektion
Parametric studies involved tests with redox oxide quantities in Energie (PD-E) for funding through the Project ‘‘Thermochemical
the range 15–150 g, investigating the effects of air flow rate and storage for CSP-applications based on Redox-Reactions – from
kind and quantity of oxide loaded on the porous supports on the materials to processes (REDOXSTORE)”.
oxygen release/uptake and the thermal effects of the thermochem-
ical reactions on the air stream exit temperatures. In the case of References
Co3O4-coated porous honeycombs and foams, thermochemical
Agrafiotis, C., Becker, A., Roeb, M., Sattler, C., 2015a. Hybrid sensible/
storage was clearly demonstrated as heat uptake/release at con- thermochemical storage of solar energy in cascades of redox-oxide-pair-based
stant temperature level, under proper testing conditions (adequate porous ceramics. In: ASME 2015 9th International Conference on Energy
redox material quantity and high flow rates to overcome heat Sustainability. American Society of Mechanical Engineers, Paper Number
ES2015-49334.
losses).
Agrafiotis, C., de Oliveira, L., Roeb, M., Sattler, C., 2016a. A solar receiver-storage
Furthermore, such hybrid sensible/thermochemical storage sys- modular cascade based on porous ceramic structures for hybrid sensible/
tems were compared to respective sensible-only ones. Again, for thermochemical solar energy storage. AIP Conf. Proc. 1734, 050001.
the Co3O4-coated honeycombs the additive effect of thermochem- Agrafiotis, C., Pitz-Paal, R., 2013. Cascaded thermochemical storage of solar heat via
redox oxide materials/Stufenweiser thermochemischer Speicher von
ical on sensible storage could be qualitatively visualized on the Solarwärme durch Redox-Materialien. German Patent Organization, Patent
temperature of the air exiting the thermochemical reaction zone. disclosure No DE102013211249A1, Germany.
694 C. Agrafiotis et al. / Solar Energy 139 (2016) 676–694

Agrafiotis, C., Roeb, M., Sattler, C., 2016b. Exploitation of thermochemical cycles assessment and semi-quantitative heat effects estimations. Sol. Energy 133,
based on solid oxide redox systems for thermochemical storage of solar heat. 394–407.
Part 4: Screening of oxides for use in cascaded thermochemical storage Kosmulski, M., 2009. Compilation of PZC and IEP of sparingly soluble metal oxides
concepts. Sol. Energy 139, 695–710. and hydroxides from literature. Adv. Colloid Interf. Sci. 152, 14–25.
Agrafiotis, C., Roeb, M., Schmücker, M., Sattler, C., 2015b. Exploitation of Kostoglou, M., Lorentzou, S., Konstandopoulos, A.G., 2014. Improved kinetic model
thermochemical cycles based on solid oxide redox systems for for water splitting thermochemical cycles using nickel ferrite. Int. J. Hydrogen
thermochemical storage of solar heat. Part 2: Redox oxide-coated porous Energy 39, 6317–6327.
ceramic structures as integrated thermochemical reactors/heat exchangers. Sol. Li, Y., Guo, B., Huang, G., Kubo, S., Shu, P., 2015. Characterization and thermal
Energy 114, 440–458. performance of nitrate mixture/SiC ceramic honeycomb composite phase
Agrafiotis, C., Tescari, S., Roeb, M., Schmücker, M., Sattler, C., 2015c. Exploitation of change materials for thermal energy storage. Appl. Therm. Eng. 81, 193–197.
thermochemical cycles based on solid oxide redox systems for thermochemical Neises, M., Tescari, S., Oliveira, L.d., Roeb, M., Sattler, C., Wong, B., 2012. Solar-
storage of solar heat. Part 3: Cobalt oxide monolithic porous structures as heated rotary kiln for thermochemical energy storage. Sol. Energy 86, 3040–
integrated thermochemical reactors/heat exchangers. Sol. Energy 114, 459–475. 3048.
Agrafiotis, C., Tsetsekou, A., 2000a. The effect of powder characteristics on washcoat Pagkoura, C., Karagiannakis, G., Zygogianni, A., Lorentzou, S., Konstandopoulos, A.G.,
quality. Part I: Alumina washcoats. J. Eur. Ceram. Soc. 20, 815–824. 2015. Cobalt oxide based honeycombs as reactors/heat exchangers for redox
Agrafiotis, C., Tsetsekou, A., 2000b. The effect of powder characteristics on washcoat thermochemical heat storage in future CSP plants. Energy Proc. 69, 978–987.
quality. Part II: Zirconia, titania washcoats—multilayered structures. J. Eur. Pagkoura, C., Karagiannakis, G., Zygogianni, A., Lorentzou, S., Kostoglou, M.,
Ceram. Soc. 20, 825–834. Konstandopoulos, A.G., Rattenburry, M., Woodhead, J.W., 2014. Cobalt oxide
Agrafiotis, C., Zygogianni, A., Pagkoura, C., Kostoglou, M., Konstandopoulos, A.G., based structured bodies as redox thermochemical heat storage medium for
2013. Hydrogen production via solar-aided water splitting thermochemical future CSP plants. Sol. Energy 108, 146–163.
cycles with nickel ferrite: experiments and modeling. AIChE J. 59, 1213– Ströhle, S., Haselbacher, A., Jovanovic, Z.R., Steinfeld, A., 2016. The effect of the gas-
1225. solid contacting pattern in a high-temperature thermochemical energy storage
Alexopoulos, S., Hoffschmidt, B., 2010. Solar tower power plant in Germany and on the performance of a concentrated solar power plant. Energy Environ. Sci. 9,
future perspectives of the development of the technology in Greece and Cyprus. 1375–1389.
Renew. Energy 35, 1352–1356. Tamme, R., Taut, U., Streuber, C., Kalfa, H., 1991. Energy storage development for
Carrillo, A.J., Serrano, D.P., Pizarro, P., Coronado, J.M., 2014. Thermochemical heat solar thermal processes. Sol. Energ. Mater. 24, 386–396.
storage based on the Mn2O3/Mn3O4 redox couple: influence of the initial Tescari, S., Agrafiotis, C., Breuer, S., de Oliveira, L., Neises-von Puttkamer, M., Roeb,
particle size on the morphological evolution and cyclability. J. Mater. Chem. A 2, M., Sattler, C., 2014. Thermochemical solar energy storage via redox oxides:
19435–19443. materials and reactor/heat exchanger concepts. Energy Proc. 49, 1034–1043.
Carrillo, A.J., Serrano, D.P., Pizarro, P., Coronado, J.M., 2015. Improving the Wokon, M., Kohzer, A., Benzarti, A., Bauer, T., Linder, M., Wörner, A., Thess, A., 2014.
thermochemical energy storage performance of the Mn2O3/Mn3O4 redox Thermochemical energy storage based on the reversible reaction of metal
couple by the incorporation of iron. ChemSusChem 8, 1947–1954. oxides. In: 3rd International Conference on Chemical Looping, September 9–11,
de Gracia, A., Cabeza, L.F., 2015. Phase change materials and thermal energy storage Göteborg, Sweden.
for buildings. Energy Build. 103, 414–419. Zanganeh, G., Khanna, R., Walser, C., Pedretti, A., Haselbacher, A., Steinfeld, A., 2015.
Fricker, H.W., 2004. Regenerative thermal storage in atmospheric air system solar Experimental and numerical investigation of combined sensible–latent heat for
power plants. Energy 29, 871–881. thermal energy storage at 575 °C and above. Sol. Energy 114, 77–90.
Karagiannakis, G., Pagkoura, C., Zygogianni, A., Lorentzou, S., Konstandopoulos, A., Zunft, S., Hänel, M., Krüger, M., Dreißigacker, V., 2014. A design study for
2014. Monolithic Ceramic redox materials for thermochemical heat storage regenerator-type heat storage in solar tower plants – results and conclusions
applications in CSP plants. Energy Proc. 49, 820–829. of the HOTSPOT project. Energy Proc. 49, 1088–1096.
Karagiannakis, G., Pagkoura, C., Halevas, A., Baltzopoulou, P., Konstandopoulos, A.G., Zunft, S., Hänel, M., Krüger, M., Dreißigacker, V., Göhring, F., Wahl, E., 2011. Jülich
2016. Cobalt/cobaltous oxide based honeycombs for thermochemical heat solar power tower—experimental evaluation of the storage subsystem and
storage in future concentrated solar power installations: multi-cyclic performance calculation. J. Sol. Energy Eng. 133, 031019.

You might also like