You are on page 1of 11

Fuel Processing Technology 150 (2016) 30–40

Contents lists available at ScienceDirect

Fuel Processing Technology

journal homepage: www.elsevier.com/locate/fuproc

Investigation of Cu–Fe and Mn–Ni oxides as oxygen carriers for


chemical-looping combustion
Volkmar Frick a,⁎, Magnus Rydén b, Henrik Leion a
a
Department of Chemistry and Chemical Engineering, Division of Environmental Inorganic Chemistry, Chalmers University of Technology, 412 96 Gothenburg, Sweden
b
Department of Energy and Environment, Division of Energy Technology, Chalmers University of Technology, 412 96 Gothenburg, Sweden

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

Article history: Combined Cu–Fe and Mn–Ni oxygen carriers were investigated as bed materials for chemical-looping combus-
Received 29 November 2015 tion. The aim was to identify material combinations which yield oxygen carriers with a high reactivity towards
Received in revised form 16 April 2016 gaseous fuels, such as CO, H2 and CH4, as well as sufficient mechanical durability. For this purpose, 18 different
Accepted 25 April 2016
oxygen carriers were spray-dried and calcined at defined temperatures.
Available online 15 May 2016
Gas conversion as well as release of gaseous oxygen was investigated in a batch fluidized bed reactor setup at
Keywords:
temperatures between 850 and 1050 °C. For testing the mechanical durability, a jet-cup attrition rig was used.
Chemical-looping combustion (CLC) Moreover, properties like specific surface area, oxygen transfer capacity and crystalline phase composition
Chemical-looping with oxygen uncoupling were examined to physically and chemically characterize the oxygen carrier particles.
(CLOU) For both the Cu–Fe and Mn–Ni-based materials, oxygen carriers could be produced which showed a high
Combined oxides oxygen carrier reactivity with gaseous fuels like CO or CH4 while having a sufficiently high mechanical strength. These properties
Copper-iron oxides make them interesting candidates for application in chemical-looping combustion.
Manganese-nickel oxides © 2016 Elsevier B.V. All rights reserved.

1. Introduction chemical-looping combustion is largely compatible with conventional


circulating fluidized bed boilers (CFB's) [2,4,5].
For mitigation of CO2 emissions from combustion of fossil fuels, To date, chemical-looping combustion has been developed for gas-
chemical-looping combustion (CLC) has been proposed as a promising eous, liquid and solid fuels and has been implemented in pilot plants
process for CO2 sequestration. It has the advantage of low cost and of sizes up to 3 MW [6]. The capability of CLC technology has been dem-
energy penalties for CO2-capture as no gas separation or scrubber unit onstrated by research groups around the world and 6700 h of pilot plant
is needed [1,2]. operation has been achieved. Reviews by Fan et al. [7], Adánez et al. [8],
In most CLC related publications, fossil fuel power plants are the and Lyngfelt [4,9] give an overview over recent developments.
target application for this technology. Nevertheless, it might also be In chemical-looping combustion the fuel gets converted to carbon
applied to industrial combustion processes as they represent large dioxide and steam by oxygen provided from a solid metal oxide. In a
point sources of anthropogenic CO2 emissions and account for about subsequent process step, this oxygen carrier is reoxidized to its initial
40% of total global carbon dioxide emissions [3]. state with air. To achieve high rates of heat and mass transfer, the pro-
cess is typically implemented as an interconnected dual fluidized bed
reactor system with the oxygen carrier as bed material, as can be seen
1.1. Chemical-looping combustion in Fig. 1. The oxygen carrying material is oxidized in a separate air reac-
tor, while fuel is converted in a fuel reactor. In this way chemical-
Chemical-looping combustion is a process for efficient utilization of looping combustion enables oxidation of the fuel while avoiding mixing
hydrocarbon fuels with inherent CO2 separation [1]. Compared to pre- of CO2 and N2/O2 in the exhaust gas streams.
or post-combustion carbon capture techniques it has the advantage of Usually, the operating temperatures of the fuel reactor are in the
lower energy demand and lower cost penalty, as CO2 is obtained in a range of 800–1000 °C. The air reactor is typically around 50 °C warmer
separate stream without any active gas separation. In addition, than the fuel reactor due to the exothermic nature of the reactions tak-
ing place here. Too high temperatures may cause agglomeration of the
⁎ Corresponding author. oxygen carrier particles whereas too low temperatures lead to insuffi-
E-mail address: volkmar.frick@chalmers.se (V. Frick). cient fuel conversion. The sum of reactions in fuel and air reactor is

http://dx.doi.org/10.1016/j.fuproc.2016.04.032
0378-3820/© 2016 Elsevier B.V. All rights reserved.
V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40 31

costs as well as low environmental impact. In CLC, Mn2O3 is reduced


in the fuel reactor to MnO whereas for CLOU, oxygen is released via con-
version of Mn2O3 to Mn3O4. Mn-based materials have a low tendency to
agglomerate as metallic Mn never forms. This is because oxides usually
have higher melting points than metals [9]. A disadvantage of pure Mn-
oxide materials is that oxidation in the air reactor has to be performed at
temperatures b 800 °C as the chemical equilibrium doesn't allow oxida-
tion of Mn3O4 at higher temperatures which is required for CLOU. Any-
how, it has been shown that conversion of gaseous fuels is also possible
if just oxidized to Mn2O3 [17,18]. As a solution to incomplete oxidation,
new materials have been produced which consist of Mn in combination
with other oxides like Cu, Fe, Mg, Ca, Ni and Si [19–21].
Another intensely studied group of materials is the one based on the
CuO‐Cu2O system which has been shown to be suitable for the CLC and
CLOU process. Drawbacks are the high raw material price and low me-
chanical strength [9,22].
When using copper-based materials, high temperatures and high
Fig. 1. Schematic description of the CLC process.
degree of reduction should be avoided as reduction all the way to
metallic Cu can cause agglomeration of the bed material due to the
exothermic and equal to the combustion of the fuel with oxygen. De- low melting temperature of elemental copper [23]. To prevent ag-
pending on the oxygen carrier material and fuel used, the reaction in glomerations, CuO has been combined with Al2O3 as support materi-
the fuel reactor might be exothermic or endothermic, the latter lowers al and no defluidization was observed up to a bed temperature of
the reactor temperature compared to the air reactor [9]. 925 °C [24]. Likewise, it has been shown that impregnated particles
(Cu on alumina) exhibited no agglomeration after 100 oxidation-
1.2. Oxygen carriers for chemical-looping combustion reduction cycles [25] and spray-dried CuO–MgAl2O4 could be oper-
ated up to 960 °C in continuous operation [16]. Nevertheless, it has
Bed materials for CLC are required to exhibit a high fuel conversion been shown that material integrity can be a problem when operating
while showing neither substantial fragmentation and abrasion nor continuously for a longer time [26]. This is confirmed by an extensive
agglomeration at relevant operating temperatures. Moreover, the cost study on supported CuO-based materials performed by Rydén et al.
for raw materials and manufacturing of the oxygen carriers should be [22] Improvements in mechanical properties can possibly be
low. If more expensive materials or production methods are used a achieved by combining with a suitable support material as has
longer lifetime is required. Besides, risks with respect to health, environ- been investigated by Gayán et al. for an Al2O3-supported Cu-based
ment and safety have to be considered. oxygen carrier with addition of NiO [27]. Cu–Fe-based oxygen car-
For combustion of ash-containing fuels, low-cost oxygen carriers are riers have been investigated by TGA for combustion of coal [28]
preferred as ash components might have a negative influence on and coal-derived syngas [29] and operation temperatures up to
oxygen carrier life time [10]. In contrast, more expensive and reactive 900 °C are suggested.
oxygen carriers could be viable for gaseous fuels [9], from which unde- In the past, much research on oxygen carriers for chemical-looping
sired impurities can be removed at reasonable cost. In CLC, oxides of Mn, combustion was done on Ni-based materials as they have high reactiv-
Fe, Ni and Cu are commonly used as oxygen carrier materials. The metal ity with methane, a phenomenon most likely associated with the cata-
oxide MexOy reacts with the fuel according to Eq. (1). lytic properties of metallic Ni. NiO supported on alumina is well
researched and has been successfully tested in long-term operation
ð2n þ mÞMex Oy þ C n H 2m ↔ð2n þ mÞMex Oy−1 þ mH 2 O þ nCO2 ð1Þ [30].
For pure NiO, the fuel conversion to CO2 and H2O is limited by chem-
ical equilibrium to 99–99.5%, depending on temperature [11]. In the
The reduced metal oxide MexOy − 1 is reoxidized in the air reactor as publication by Shulman et al. [20], fuel conversion and oxygen release
described by Eq. (2). of three particles produced from 80 wt.% Mn3O4 and 20 wt.% NiO are
presented. Interestingly, low formation of CO was detected during com-
bustion of methane which implies nearly complete conversion of the
1
Mex Oy−1 þ O2 ↔Mex Oy ð2Þ methane.
2
A drawback of nickel-based materials is that they are sensitive to
sulphur poisoning which limits their application to sulphur-free fuels
The total heat released by a chemical-looping combustor is the same [31]. Furthermore, the fuel should be ash-free or even gaseous, as ash-
as for normal combustion [11]. rich fuels make frequent replacement of the oxygen carrier material
Some metal oxide systems such as Mn2O3/Mn3O4 and CuO/Cu2O are necessary which is not economically feasible for materials with a high
known to spontaneously release gaseous oxygen when exposed to inert raw material price such as nickel. Besides, for Ni-based particles, health
atmosphere. When applied in CLC, this is referred to as “chemical- and environmental aspects have to be considered [9,32].
looping with oxygen uncoupling” (CLOU). CLOU has been shown to Synthetic oxygen carriers are commonly produced by spray-drying,
improve conversion of solid fuels and it is expected that it will also en- impregnation or freeze-granulation. Spray-drying is often applied be-
hance combustion of liquid or gaseous fuels [12–14]. Furthermore, it has cause it is suitable for production of large quantities of materials. The
been found that the fuel reactor inventory can be significantly reduced if mechanical strength of the particles can be increased by calcination
materials with CLOU capability are used [15]. For CLOU, several material and sintering [9]. The time and temperature required vary between dif-
combinations have been investigated as bed materials and the suitabil- ferent material compositions.
ity of the process has been demonstrated in continuous operation [8,9, The cost of synthetically produced oxygen carriers has been thor-
16]. oughly discussed by Lyngfelt. The author estimated the cost for
An important group of oxygen carriers is Mn-based materials as they spray-drying to be 500–5000 €/t and raw material prices of copper,
have comparatively high reactivity with CO, H2 and CH4, low material manganese and nickel to be 5400 €/t, 150–400 €/t and 13,000 €/t,
32 V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40

respectively. These material costs are significantly higher than those 2.2. Material characterization
of natural ores such as ilmenite (100 €/t) but might be justified if the
overall process cost can be reduced by higher reactivity and material For engineering of fluidized bed reactor systems, the particles' den-
lifetime [9]. sity strongly affects the minimum fluidization velocity needed to fluid-
ize the particles [34]. Here, the density was measured by weighing a
defined volume of the material. Another important property of bed ma-
1.3. Aim of this study terial is the mechanical durability, which in this study was characterized
by measurement of the crushing strength and the attrition resistance.
This study focuses on investigation of Cu–Fe and Mn–Ni oxygen car- The crushing strength was determined by measuring the force to
riers to be used in chemical-looping combustion. Materials based on Cu, break 30 individual particles in the size range of 180–250 μm using a
Mn or Ni have attracted attention in CLC-related research in the past and digital force gauge (Shimpo FGN-5) and calculating the average value.
are likely to also yield good results when used as combined oxides. The To measure the attrition resistance, a jet cup attrition rig was used
Cu-system has been shown to have remarkable oxygen release proper- where the particles were exposed to a high velocity gas jet. The fines
ties whereas the Ni-system is promising with respect to gas conversion. produced are collected in a filter which is weighed after specific time in-
Combining both systems with other metal oxides might yield bed mate- tervals. To compare the attrition behavior of different particles, the attri-
rials with a lower price and similar or better properties e.g. by formation tion index Ai is calculated according to Eq. (3).
of solid solutions.

Important factors to be assessed in this study are reactivity towards m f ;t¼60 min −m f ;t¼30 min 60
gaseous fuels such as H2, CO and CH4, CLOU oxygen release and mechan- Ai ¼ ∙ ∙100 ½wt:%=h ð3Þ
ms 30
ical durability. The aim is to identify robust oxygen carriers while taking
advantage of the high O2 release and gas conversion.
mf,t is the cumulative elutriated mass at time t and ms is the initial
mass of the sample. The attrition index shall not be interpreted as an es-
2. Experimental timation of the expected life-time of oxygen carrier materials in practi-
cal operation, but only as a measure of the relative mechanical
2.1. Assessed oxygen carriers durability of the particles. A detailed overview of the set-up and proce-
dure for attrition measurements is given by Rydén et al. [35].
In this study, oxygen carriers consisting of different ratios of Cu/Fe Furthermore, the specific surface area (BET) of the particles was
and Mn/Ni were studied. The oxygen carriers were manufactured by measured before (fresh state) and after (used state) testing in the fluid-
VITO in Belgium by means of spray-drying. They were produced from ized bed reactor. This was done by N2 adsorption in a Micromeritics
commercial raw materials (CuO: Sigma-Aldrich, Fe2O3: Alfa Aesar, NiO TriStar 3000 apparatus.
(Green grade F): Novamet, Mn3O4 (Trimanox): Chemalloy). First, To determine the phase composition of the materials, they were ana-
these raw materials were milled to powder and dispersed in deionized lyzed before and after fluidized bed reactor testing by powder X-ray dif-
water together with polyvinyl alcohol (PVA) as organic binder. This sus- fractometry (Siemens D5000, Cu Kα1 radiation). Thereby it was possible
pension was then pumped through a nozzle into the spray-drying appa- to draw conclusions about the mechanism responsible for oxygen release.
ratus where the water evaporated. The particles obtained were calcined For quantification of CLOU oxygen release, measurements with a Q500
in air for 4 h. Table 1 gives an overview of the particles and production TA-Instruments thermogravimetric analyzer (TGA) were performed. Dur-
conditions being used. This production method has been used before ing these tests the feed gas volume flows and the sample mass have been
and is further explained elsewhere [33]. varied to minimize mass transfer limitations. A sample of 7 mg oxygen
The Cu–Fe sample with the highest Cu content calcined at 1100 °C carriers, which has already experienced reducing and oxidizing conditions
(Cu30Fe_1100) could not be tested as production of particles failed in a batch fluidized bed reactor, was placed on an alumina crucible and ex-
due to the high calcination temperatures. posed to alternating oxidizing (5 vol.% O2 in N2) and inert conditions. The
material was oxidized for 60 min and exposed to pure N2 for 15 min. To
ensure stable oxygen release, at least five oxidation-reduction cycles
were performed. The oxygen transfer capacity RO was calculated accord-
Table 1
ing to Eq. (4) by relating the difference between the sample mass in oxi-
Properties of oxygen carriers assessed in this study.
dized state mox and the mass in reduced state mred (at the end of the
Oxygen carrier Synthesis composition [wt.%] Calcination temperature [°C] inert period) to the sample mass in oxidized state mox.
Cu5Fe_950 950
Cu5Fe_1025 5% CuO, 95% Fe2O3 1025 mox −mred
Cu5Fe_1100 1100 RO ¼ 100 ½wt:% ð4Þ
mox
Cu15Fe_950 950
Cu15Fe_1025 15% CuO, 85% Fe2O3 1025
Cu15Fe_1100 1100 The maximum absolute rate in weight change of the oxygen carrier
was taken as a measure for the maximum CLOU reaction rate.
Cu30Fe_950 950
Cu30Fe_1025 30% CuO, 70% Fe2O3 1025
Cu30Fe_1100a 1100 2.3. Experimental method and setup for batch reactor testing
Ni5Mn_1000 1000
Ni5Mn_1100 5% NiO, 95% Mn3O4 1100 A batch fluidized bed reactor made of quartz glass with an inner di-
Ni5Mn_1200 1200 ameter of 22 mm and a height of 820 mm has been used to determine
Ni10Mn_1000 1000 the materials' capability to release gaseous oxygen and to convert syn-
Ni10Mn_1100 11% NiO, 89% Mn3O4 1100 gas (50 vol.% H2 in CO) as well as methane. At a height of 370 mm (mea-
Ni10Mn_1200 1200 sured from the bottom), a porous quartz plate was placed to support the
Ni33Mn_1000 1000 bed material. The temperatures of inflowing gas and of the fluidized bed
Ni33Mn_1100 34% NiO, 66% Mn3O4 1100 were measured by thermocouples (Type K) just under the porous plate
Ni33Mn_1200 1200 and in the actual bed. The condition of the fluidized bed was monitored
a
Material formed a solid block during calcination. by measuring the pressure drop over the reactor.
V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40 33

content measured does not represent oxygen released from the materi-
al but rather oxygen from the previous oxidation period.
After operation at 1050 °C, reactivity with methane and CLOU oxy-
gen release have been assessed again at 950 °C to check if the material
suffered damage due to high process temperatures.
The values for the gas flowrates were chosen significantly above the
minimum fluidization velocity Umf to ensure sufficient fluidization of
the particles. For oxidizing conditions, a flowrate of 900 ml/min was
chosen which corresponds to a fluidization velocity 10–14 times higher
than the minimum fluidization velocity (0.014 m/s b Umf,oxidizing b
0.016 m/s, depending on operating temperature). For the inert period
and fuel period (syngas/methane), flow rates of 600 ml/min and
Fig. 2. Scheme of the experimental setup for batch reactor testing. 450 ml/min were chosen which correspond to 7–9 and 4–5 times the
minimum fluidization velocity (0.014 m/s b Umf,inert b 0.016 m/s,
0.019 m/s b Umf,fuel b 0.023 m/s) [34].
The feed gas flows were regulated by mass flow controllers. Comput- Characteristic numbers for comparison of oxygen carriers are the
er controlled electric valves were switched according to a pre-defined fuel conversion which is expressed by the CO2 yield γCO2 and the degree
sequence to provide oxidizing, inert or reducing gas atmospheres. The of oxygen carrier conversion ω which are calculated according to
product gases were cooled after leaving the reactor and steam was con- Eqs. (5) and (6).
densed. The dry gas composition together with the gas flow was mea-
sured by a Rosemount NGA-2000 analyzer. Nondispersive infrared
yCO2
detectors (NDIR) were used for measurement of CO, CO2 and CH4, γCO2 ¼ ½− ð5Þ
yCO2 þ yCO þ yCH4
whereas a thermal conductivity detector (TCD) and a paramagnetic
sensor were used for H2 and O2, respectively. A sketch of the experimen-
tal setup is given in Fig. 2.
yi is the molar fraction of component i in the off-gas stream.
Of each oxygen carrier material, 15 g particles of size 125–180 μm
were placed on the porous plate inside the reactor. The particles were
fluidized in oxidizing conditions (5 vol.% O2 in N2) while the reactor m
ω¼ ½−  ð6Þ
was heated to the operating temperature (850–1050 °C). This specific mox
oxygen concentration was chosen because it resembles the target con-
centration at the outlet of a continuous CLC air reactor. In a commercial
unit, higher oxygen partial pressures would cause larger exhaust gas m is the current mass and mox the mass of the oxygen carrier in its
flows along with a decreased thermal efficiency [9]. fully oxidized condition.
The particles were exposed to oxidizing, inert and reducing condi- The oxygen carrier conversion ω can't directly be measured in a fluid-
tions after having reached a specific bed temperature according to the ized bed but it can be calculated as a function of time from the effluent
procedure given in Table 2. A sequence of reducing conditions followed concentrations by applying an oxygen mass balance over the reactor.
by oxidizing conditions is referred to as a cycle. All cycles were conduct- Eq. (7) is valid for combustion of syngas and Eq. (8) for combustion of
ed at least twice to guarantee a steady cyclical behavior. Before and after methane.
combustible gases were fed into the reactor, it was flushed for 60 s with
nitrogen to avoid measurement errors due to combustion of fuel with
t n_ M O
remaining oxygen. For re-oxidation, the particles were exposed to oxi- ωt ¼ ωt−1 −∫ t−1 ∙ð2yCO2 þ yCO −yH2 þ 2yO2 Þdt ð7Þ
mox
dizing conditions until the oxygen concentration in the product gas
was stable and reached again 5 vol.%.
Pure methane and a mixture of 50 vol.% CO in H2 (syngas) were ap- t n_ M O
plied for investigation of the oxygen carriers' reactivity with gaseous ωt ¼ ωt−1 −∫ t−1 ∙ð4yCO2 þ 3yCO −yH2 þ 2yO2 Þdt ð8Þ
mox
fuels. The period for feeding syngas was chosen 4 times longer than
the period for feeding of methane since CH4 consumes four times
more oxygen per mole of gas. MO is the molar mass of oxygen and n_ the molar flow rate at the re-
Assessment of the CLOU oxygen release was done by exposing the actor outlet.
oxygen carrier particles to N2 for 360 s. The oxygen content in the off- For comparison of different oxygen carriers at different operating
gas was recorded and the average value during the last 300 s was conditions, average values for the CO2 yield have been calculated for ω
used for comparison. The first 60 s were neglected because the oxygen ranging from 1 to 0.99.

Table 2
Experimental procedure used to investigate oxygen release and gas conversion.

Gas composition Bed No. of cycles Duration Duration Duration


temperature [°C] inert period (CLOU) [s] N2-purge [s] reduction period [s]

50 vol.% CO in H2 950 2 – 60 80
100 vol.% CH4 850 2 – 60 20
100 vol.% N2 850 2 360 – –
100 vol.% CH4 950 2 – 60 20
100 vol.% N2 950 2 360 – –
100 vol.% CH4 1050 2 – 60 20
100 vol.% N2 1050 2 360 – –
100 vol.% CH4 950 2 – 60 20
100 vol.% N2 950 2 360 – –
34 V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40

Table 3 Table 4
Bulk density (tapped) and specific surface area (BET) for investigated particles. Oxygen transfer capacity (CLOU) and maximum absolute weight change (=max. oxygen
release rate) for Cu–Fe and Mn–Ni particles at 950 °C.
Oxygen carrier Bulk density [g/cm3] Specific surface area
(BET) [m2/g] Oxygen O2 transfer capacity at Max. abs. weight change at 950 °C
carrier 950 °C [wt.%] [wt.%/min]
Fresh Used Fresh Used
Cu5Fe_950 0.4 0.3
Cu5Fe_950 1.9 2.0 0.8 0.3
Cu5Fe_1025 0.4 0.2
Cu5Fe_1025 2.1 2.0 0.1 0.2
Cu5Fe_1100 0.4 0.2
Cu5Fe_1100 2.1 1.9 0.1 0.2
Cu15Fe_950 1.2 0.6
Cu15Fe_950 1.5 1.4 0.8 0.5
Cu15Fe_1025 1.3 0.7
Cu15Fe_1025 1.8 1.5 0.4 0.4
Cu15Fe_1100 1.2 0.6
Cu15Fe_1100 2.0 1.8 0.1 0.2
Cu30Fe_950 2.6 3.3
Cu30Fe_950 1.4 1.6 0.6 0.2
Cu30Fe_1025 2.1 1.0
Cu30Fe_1025 2.0 1.7 0.1 0.1
Ni33Mn_1000 1.2 0.9
Ni5Mn_1000 1.5 1.6 0.6 0.4
Ni33Mn_1100 1.2 0.7
Ni5Mn_1100 2.1 2.1 0.1 0.1
Ni33Mn_1200 1.1 0.4
Ni5Mn_1200 2.3 2.2 0.1 0.1
Ni10Mn_1000 1.6 1.7 0.4 0.3
Ni10Mn_1100 2.0 2.0 0.2 0.1
Ni10Mn_1200 2.2 2.2 0.1 0.1 oxidizing to inert atmosphere. In Fig. 3, the average oxygen concentra-
Ni33Mn_1000 1.5 1.7 0.6 0.3 tion during the last 300 s of the inert period is depicted for three differ-
Ni33Mn_1100 2.1 2.1 0.3 0.2 ent bed temperatures. The values for experiments done at 950 °C after
Ni33Mn_1200 2.4 2.4 0.0 0.1
operation at 1050 °C are also included. It can be seen that higher opera-
tion temperatures led to higher oxygen concentrations in the exhaust
3. Results gas.
For Cu–Fe-based materials (Fig. 3, left side), it is clear that those with
3.1. Physical properties the highest copper content, Cu30Fe_950 and Cu30Fe_1025, performed
significantly better than the other six materials. Moreover, it can be
The specific surface area and the bulk density of the Cu–Fe and Mn–Ni seen that the average oxygen concentration was proportional to the
oxygen carrier particles are given in Table 3. It can be seen that for Mn–Ni copper content of the oxygen carrier material. The difference between
based materials the density increased with increasing calcination temper- materials with the same material composition but different calcination
ature. For Cu–Fe based materials, this trend was more pronounced at temperature was marginal.
higher copper content. During reactivity testing in the batch fluidized bed When looking at the data for Mn–Ni materials (Fig. 3, right side), it
reactor, the bulk density changed just for Cu–Fe and there seems to be a becomes obvious that just the materials produced from 34 wt.% NiO
trend towards a similar density for all particles with a specific Cu–Fe ratio. showed a noteworthy oxygen release. With respect to calcination tem-
Regarding the specific surface area, it can be noted that it decreased perature during manufacturing of the oxygen carriers, it can be said that
for both material groups with increasing calcination temperature which lower calcination temperatures led to higher oxygen concentrations.
may indicate a decreasing porosity. Comparing the fresh and used state, This effect was more pronounced at lower fluidized bed temperatures.
particles with low calcination temperature underwent a decrease in To measure the maximum oxygen release rate as well as the oxygen
specific surface area whereas those with higher calcination temperature transfer capacity RO, CLOU cycles were performed in a thermogravimet-
tended to remain rather constant. ric analyzer (TGA) at 950 °C as explained in Section 2.2. The correspond-
With respect to different copper contents, no major effect on surface ing results are summarized in Table 4 and TGA curves of selected
area could be observed for particles calcined at the same temperature. carriers (one for each synthesis composition) are shown Fig. 4. For
The same is valid for the nickel content which neither had a strong influ- Mn–Ni-based materials, just those who showed significant oxygen re-
ence on surface area nor bulk density. lease in the fluidized bed experiment were investigated by TGA.
It can be seen that the oxygen transfer capacity and the maximum
3.2. CLOU oxygen release release rate increased with increasing copper content. For
Cu30Fe_950, RO was 6.5 times higher than for the material produced
The spontaneous oxygen release due to chemical-looping with oxy- from 5 wt.% CuO, e.g. Cu5Fe_950. Regarding the max. O2 uncoupling
gen uncoupling has been assessed by switching the reactor from rate, a value 11 times higher could be achieved for Cu30Fe_950

Fig. 3. Average oxygen concentration during inert period for Cu–Fe (left) and Mn–Ni oxygen carriers (right).
V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40 35

15 wt.% CuO and Mn–Ni based materials produced from 34 wt.% NiO ex-
hibited a high and relatively stable methane conversion. For Mn–Ni, the
reactivity towards methane seems to be related to the nickel content.
With respect to the impact of calcination temperature, it is clear that
lower calcination temperatures promoted reactivity towards methane.
Regarding conversion of methane, the average CO2 yield as a func-
tion of temperature is shown in Fig. 7. In most cases, higher tempera-
tures promoted gas conversion. At a bed temperature of 1050 °C, all
oxygen carriers but Ni5Mn_1200 and Ni10Mn_1000 achieved a CO2
yield above 0.9. Exceptional reactivity was found for Ni33Mn_1000
and Ni33Mn_1100 with an average CO2 yield greater than 0.95 for all
temperatures investigated.
Comparison of methane conversion before and after operation at
1050 °C shows that most materials retained their reactivity. A significant
deviation can be seen for Ni10Mn_1000 which exhibited a lower gas
conversion at 1050 °C and even lower at 950 °C. Reactivity deteriorated
also for Cu5Fe_950, Cu15Fe_1100 and Cu30Fe_950 due to high temper-
Fig. 4. TGA-curves for CLOU oxygen release at 950 °C for selected oxygen carriers. ature operation. High reactor temperatures together with strongly re-
ducing conditions are known to cause structural changes especially in
compared to Cu5Fe_950. With respect to calcination temperature, it can materials based on elements with low melting point such as copper.
be seen that differences in transfer capacity and release rate for mate- During activity testing, most materials tended to form agglomera-
rials with low copper content (Cu5Fe, Cu15Fe) were marginal whereas tions. Only Cu5Fe_950 did not show any signs of agglomeration. All ag-
for materials produced from 30 wt.% CuO, both values decreased for in- glomerations had in common that they were rather soft and fell apart
creasing calcination temperature. when some force was applied.
For the Mn-Ni-based materials, it can be concluded that there was
just a marginal change in transfer capacity when increasing calcination 3.4. Mechanical durability
temperature. In contrast, the oxygen release rate decreased almost lin-
early for increasing calcination temperature. The mechanical durability was assessed by measuring the crushing
Concerning the evolution of CLOU reaction rate with time (s. Fig. 4), strength of fresh particles and the attrition rate for used particles. A
it is apparent that the mass change of the oxygen carriers was highest low attrition rate and high crushing strength are preferred properties
within the first minute of exposure to N2 and decreased with time. of oxygen carriers in fluidized bed applications. The measurement re-
However, the materials still showed a measurable weight change even sults for Cu–Fe-based materials are summarized in Fig. 8.
after 5 min. Whether this behavior limits a chemical-looping process It is apparent that the crushing strength of the fresh particles in-
depends on the specific residence time distribution of the particles in creased for increasing calcination temperature indicating a higher me-
the fuel reactor. chanical durability. Interestingly, only the used particles produced
from 15 wt.% CuO followed this trend i.e. they showed a lower attrition
3.3. Fuel conversion index for materials with higher calcination temperature. All other used
Cu–Fe-based particles exhibited a fairly low attrition rate which was in-
Research on combustion of gaseous fuels was done to assess the ma- dependent from calcination temperature. This suggests that their me-
terials' capability to provide oxygen. In a first stage, combustion of syn- chanical properties must have significantly been altered due to
gas was investigated as H2 and CO are both easy to convert. Fig. 5 shows reactivity testing.
the conversion of CO which is expressed by the CO2 yield as a function of To determine whether an oxygen carrier might be interesting for
temperature. further research, the average CO2 yield at 950 °C is plotted versus the at-
It is clear that all materials showed a high degree of syngas conver- trition index for all Cu-Fe-based materials being part of the investigation
sion at 950 °C. Concerning hydrogen, full conversion to H2O was (s. Fig. 8, right side). As the attrition rate resembles the mechanical du-
achieved for all materials. rability after reactivity testing, values for the CO2 yield at 950 °C after
In contrast, conversion of methane was more difficult than conver- operation at 1050 °C are used likewise. With respect to methane conver-
sion of syngas. This becomes quite obvious when comparing Fig. 6 sion and attrition index, the oxygen carriers Cu30Fe_950 and
with Fig. 5. Interestingly, only Cu–Fe based materials produced from Cu30Fe_1025 are highly interesting candidates as they both showed

Fig. 5. CO2 yield as a function of oxygen carrier conversion for combustion of syngas at 950 °C for Cu–Fe oxygen carriers (left) and Mn–Ni oxygen carriers (right).
36 V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40

Fig. 6. CO2 yield as a function of oxygen carrier conversion at 950 °C for combustion of methane.

high gas conversion and a low attrition index. Not equally good but still 3.5. X-ray diffraction results
remarkable are the oxygen carriers Cu5Fe_950, Cu5Fe_1025 and
Cu15Fe_1100. They had a somewhat higher attrition rate but still fairly The crystalline phases have been analyzed by X-ray diffraction to
high CO2 yields. draw conclusions about the mechanism responsible for CLOU oxygen
For oxygen carriers based on Ni, mechanical durability was only release and to check if the material could be reoxidized by 5 vol.% O2
assessed by measuring the crushing strength as usage of the attrition after having been exposed to reducing gases such as CH4 and syngas.
rig is restricted to non-toxic material. In a similar manner as for the The phases determined from X-ray diffractograms are listed in Table 5.
Cu-Fe-based materials, the CO2 yield is plotted versus the mechanical The predominant phases detected in all Cu–Fe materials were Fe2O3
durability which in this case is represented by the crushing strength in- and the copper–iron spinel phase (CuxFe1 − x)3O4. This spinel phase is a
stead of the attrition index as can be seen in Fig. 9. solid solution of CuFe2O4 and Fe3O4. In some samples, also single Fe3O4
Interestingly, the particles with different Ni content followed deviat- of spinel type has been detected but this result is quite uncertain as peak
ing trends for CO2 yield and crushing strength when calcined at differ- overlapping occurs with the other spinel phase. Comparison with the
ent temperatures. Those produced from 34 wt.% NiO showed Cu–Fe–O phase diagram generated by Khvan et al. [36] shows that the
significant changes in crushing strength but had an almost constant detected crystalline phases in the used samples agree well with chemi-
gas conversion. In contrast, the materials produced from 5 wt.% NiO ex- cal equilibrium at 950 °C and 5 vol.% O2 indicating that reoxidation was
hibited an almost inversely proportional relationship between CO2 yield possible under these conditions. From the phase diagram and the XRD
and crushing strength. data, it can be concluded that CLOU oxygen release most likely occurred
It is apparent that the best candidate with respect to methane con- due to the phase change expressed by Eq. (9).
version is the material with the highest Ni content sintered at 1200 °C,
Ni33Mn_1200. This oxygen carrier had a fairly high crushing strength 
ðCux Fe1−x Þ3 O4 þ Fe2 O3 ↔ Cux Fe5=3−x 3 O4 þ 3=2O2 ð9Þ
and almost full gas conversion. Also, the other two materials with the
same Ni content might be interesting candidates as mechanical durabil-
ity was just tested for fresh materials and solidification might have oc- From the phase diagram [36], it can be seen that oxygen release
curred due to repeating reduction and oxidation cycles. Cu
should improve with increasing copper content in the range 0≤ CuþFe ≤0

Fig. 7. Average CO2 yield from methane conversion as a function of temperature. Left side: Cu–Fe materials, right side: Mn–Ni materials.
V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40 37

Fig. 8. Left side: attrition (used particles) and crushing strength (fresh particles) data for investigated materials. Right side: average CO2 yield for methane conversion at 950 °C as a function
of attrition index for Cu–Fe particles.

:45 as the phase boundary approaches lower temperatures. This agrees be suitable for CLC with respect to methane conversion and mechanical
well with the results for oxygen uncoupling in Section 3.2. durability.
For the materials based on Mn–Ni, spinel-type (MnxNi1 − x)3O4 has In the overview images, it can be noted that the particles mostly
been found in all samples. Furthermore, Mn3O4 (Hausmannite) and have a spherical shape and it seems to be quite common that they
Mn2O3 were frequently found in the used samples. For materials pro- have one or more small satellite particles attached to them. Some parti-
duced from 34 wt.% NiO, NiO was found in all fresh as well as used sam- cles even exhibit a donut shape and are sometimes filled with smaller
ples. By comparison with a phase diagram for Mn–Ni in air, published by particles. When comparing pre and post-reaction particles, no agglom-
Golikov and Balakirev [37], it becomes obvious that the predominant erations can be seen at this scale but some debris can be found in the
phases occurred according to chemical equilibrium. For the fresh sam- used samples. The satellites on the particles are likely to have been
ples of Ni5Mn and Ni10Mn, the existence of the phases Mn–Ni-spinel, worn off during the attrition test and with an optimized production
Mn2O3 and Mn3O4 can be explained either by the production conditions method it is likely that more uniform particles can be produced that
or by the fact that they were cooled down in air after calcination. Only have less satellites on them and hence lower attrition.
for the Ni33Mn materials, the presence of the phases Mn3O4 and NiO At 10,000× magnification, it can be noted that for the copper–iron
can't be explained by chemical equilibrium as they are not stable in material the grain size increased and the pore number decreased as a re-
the relevant temperature interval. A possible explanation for this is sult of activity testing which indicates a decreased particle porosity. For
that the raw materials of these two compounds might not have been the Mn–Ni material, no major change in surface morphology can be ob-
milled and mixed sufficiently to form the common Mn–Ni-spinel served. The grain size and shape seem to be similar but there might be a
phase. In the used samples, it can be seen that the same phases occurred slight increase in pore number. These observations agree well with the
as in the fresh samples. With respect to reoxidation of the particles, this changes in physical properties described in Section 3.1 as a significant
is a good sign because all phases formed during reduction seem to have increase in bulk density and a notable decrease in surface area have
been oxidized back to the original phases. been detected for Cu30Fe_950. In contrast, Ni33Mn_1200 showed no
Possible mechanisms for CLOU oxygen release for Ni5Mn and change in bulk density and just a slightly higher specific surface area.
Ni10Mn can't be formulated from the phase diagram [37] because
phase transformation from the two phases, stable at 900 °C, Mn–Ni-spi-
nel and Hausmannite (Mn3O4) to Mn–Ni-spinel does not yield gaseous
oxygen. This has also been observed during CLOU experiments (s.
Section 3.2). Oxygen release is only possible at strongly reducing condi-
tions. In contrast, for Ni33Mn, oxygen release is possible by converting
the Mn–Ni-spinel to NixMn1 − xO which enables formation of gaseous
oxygen according Eq. (10).

1
ðNix Mn1−x Þ3 O4 ↔3 Nix Mn1−x O þ O2 ð10Þ
2

At strongly reducing conditions e.g. very high temperatures,


NixMn1 − xO might even decompose to elemental nickel and depleted
NixMn1 − xO [37].

3.6. Electron imaging (SEM)

Electron images were taken with low (260 ×), intermediate


(~ 2000 ×) and high magnification (10,000 ×) to identify changes in
the particle collective, the morphology of a selected particle and the mi-
crostructure on the particle's surface when comparing fresh and used
samples. From the Cu–Fe and Mn–Ni-based materials, Cu30Fe_950 Fig. 9. Average CO2 yield from methane conversion at 950 °C as a function of crushing
(Table 6) and Ni33Mn_1200 (Table 7) are presented as they proved to strength.
38 V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40

Table 5 min (45 wt.% CuO) according to the definition in Section 2.2. In our
Crystalline phases detected by X-ray diffraction. work, the highest rate measured was 3.3 wt.%/min for Cu30Fe_950
Oxygen carrier fresh sample used sample which is in line with the results from Clayton and Whitty.
Cu5Fe_950 Fe2O3, (CuxFe1 − x)3O4, (Fe3O4) Fe2O3, (CuxFe1 − x)3O4
Moreover, all Cu–Fe materials showed nearly full syngas conversion.
Cu5Fe_1025 Fe2O3, (CuxFe1 − x)3O4, Fe3O4 Fe2O3, (CuxFe1 − x)3O4, Fe3O4 With respect to methane conversion, still 4 materials achieved a CO2
Cu5Fe_1100 Fe2O3, (CuxFe1 − x)3O4 Fe2O3, (CuxFe1 − x)3O4, Fe3O4 yield above 90% for a process temperature as low as 950 °C. If related
Cu15Fe_950 Fe2O3, (CuxFe1 − x)3O4, Fe3O4 Fe2O3, (CuxFe1 − x)3O4, Fe3O4 to the Cu-based oxygen carrier CuAl2O4 investigated previously by
Cu15Fe_1025 Fe2O3, (CuxFe1 − x)3O4, (Fe3O4) Fe2O3, (CuxFe1 − x)3O4, Fe3O4
Arjmand et al. [24], it can be said that the methane conversion is
Cu15Fe_1100 Fe2O3, (CuxFe1 − x)3O4, (Fe3O4) Fe2O3, (CuxFe1 − x)3O4, Fe3O4
Cu30Fe_950 (CuxFe1-x)3O4, Fe2O3 (CuxFe1-x)3O4, Fe2O3, Fe3O4 lower for the materials investigated here. The fact that the oxygen car-
Cu30Fe_1025 (CuxFe1-x)3O4, Fe2O3 (CuxFe1-x)3O4, Fe2O3, Fe3O4 riers with 30 wt.% CuO, Cu30Fe_950 and Cu30Fe_1025, showed lower
Ni5Mn_1000 Mn2O3, (MnxNi1 − x)3O4 Mn3O4, (MnxNi1 − x)3O4, methane conversion than the carriers with 15 wt.% CuO is somewhat
Mn2O3 unexpected especially if taking the oxygen transfer capacities (s.
Ni5Mn_1100 Mn2O3, Mn3O4, (MnxNi1 − x)3O4 Mn3O4, (MnxNi1 − x)3O4,
Mn2O3
Table 4) into account. A possible explanation might be the 60 s inert
Ni5Mn_1200 Mn3O4, (MnxNi1 − x)3O4 Mn3O4, (MnxNi1 − x)3O4, purging period before introducing the fuel as a significant amount of ox-
Mn2O3 ygen could have already been released due to CLOU, especially for ma-
Ni10Mn_1000 Mn2O3, (MnxNi1 − x)3O4 (MnxNi1-x)3O4, Mn3O4, Mn2O3 terials with high oxygen release rate.
Ni10Mn_1100 Mn2O3, (MnxNi1 − x)3O4 (MnxNi1-x)3O4, Mn3O4, Mn2O3
The weaknesses of copper-based materials become obvious when
Ni10Mn_1200 (MnxNi1 − x)3O4, Mn3O4, Mn2O3 (MnxNi1-x)3O4, Mn3O4, Mn2O3
Ni33Mn_1000 (MnxNi1 − x)3O4, Mn3O4, NiO (MnxNi1-x)3O4, NiO turning to the mechanical properties. Low values for crushing strength
Ni33Mn_1100 (MnxNi1 − x)3O4, Mn3O4, NiO (MnxNi1-x)3O4, Mn3O4, NiO were measured before reactivity testing especially for oxygen carriers
Ni33Mn_1200 (MnxNi1 − x)3O4, Mn3O4, NiO (MnxNi1-x)3O4, Mn3O4, NiO with higher copper content. Interestingly, solidification occurred for
the weakest materials with highest copper content when exposing
them to high temperatures and reducing atmosphere so that they
4. Discussion reached the best values for the attrition index (used state). In the exten-
sive attrition study published by Rydén et al. [35], spray-dried CuO/ZrO2
The combined copper iron oxygen carriers, assessed within this particles showed attrition indices between 6.78 and 27 wt.%/h. Com-
work have shown very interesting chemical properties. They exhibited pared to these values, the particles with 5 and 30 wt.% CuO showed sig-
significant oxygen release even at temperatures as low as 850 °C. Both nificantly better attrition behavior.
materials produced from 30 wt.% CuO, Cu30Fe_950 and Cu30Fe_1025, When considering chemical and mechanical performance, it
showed oxygen release superior to the other materials with O2 concen- can be seen that the two materials produced from 30 wt.% CuO
trations above 2 vol.% for a process temperature of 1050 °C. Likewise, showed both a high methane conversion and a substantial me-
they had the highest oxygen transfer capacity and the fastest oxygen re- chanical durability after having been exposed to cyclic oxidation
lease rate which can be explained by the course of the phase boundary and reduction which makes them interesting candidates for fur-
relevant for CLOU oxygen release. In literature, some data on CLOU reac- ther research.
tion kinetics determined by TGA is available. Cu-based oxygen carriers The most striking result for the combined Mn–Ni oxygen carriers is
prepared by wet impregnation and freeze granulation were assessed the CLOU oxygen release: materials with high Ni content released sig-
by Clayton and Whitty [38]. They determined a decomposition rate of nificant amounts of oxygen whereas materials produced from 5 or
about 0.003 gO2/(gCu*s) at 950 °C for the carrier with 45 wt.% CuO on 10 wt.% NiO released only traces. The materials produced from
ZrO2 and a slightly lower rate for 16 wt.% CuO on SiO2 which corre- 34 wt.% NiO also proved to have a high oxygen transfer capacity and
sponds to a rate ranging from 2.3 wt.%/min (16 wt.% CuO) to 6.5 wt.%/ O2 release rate.

Table 6
SEM images of Cu30Fe_950 in fresh and used state: overview image, individual particle and microstructure at 10,000× magnification.

Overview Individual particle /10,000× magn.

Cu30Fe_950 (fresh)

Cu30Fe_950 (used)
V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40 39

Table 7
SEM images of Ni33Mn_1200 in fresh and used state: overview image, individual particle and microstructure at 10,000× magnification.

Overview Individual particle /10,000× magn.

Ni33Mn_1200 (fresh)

Ni33Mn_1200 (used)

Regarding conversion of fuel, all Mn–Ni-based materials showed sig- A drawback of all materials assessed in this study is that they formed
nificant gas conversion as additional oxygen release is possible at agglomerations during testing in the batch fluidized bed reactor at tem-
strongly reducing conditions. Even though all materials showed almost peratures as high as 1050 °C. Whether this turns out to be relevant for
full and stable syngas conversion, huge differences for methane conver- real application remains to be shown in a continuous unit. This is mainly
sion could be seen: Ni5Mn and Ni10Mn showed similar CO2 yields be- because the fluidization conditions present in the batch fluidized bed
tween 40 and 90% but Ni33Mn exhibited a CO2 yield between 90 and reactor are different from those in continuous dual fluidized bed reac-
100%. Interestingly, the materials with high Ni content showed almost tors. Especially the reactor temperatures used for investigation of the
full gas conversion (99.6–99.9%) for methane at 1050 °C and a similar materials are higher than those commonly used in continuous CLC
conversion for syngas at 950 °C. In literature, it is reported that the sys- units which are mostly limited to 950 °C. Furthermore, it is likely that
tem NiO/Ni is not capable of fully converting methane or CO at relevant the degree of reduction of the oxygen carrier will be different in a con-
temperatures [39]. By combining Ni with Mn, this limitation seems to tinuous reactor system which can significantly change the melting point
have been overcome. A reason for this is possibly the formation of of the material and thereby its agglomeration tendency.
solid solutions such as Ni–Mn spinel which have significant impact on
oxygen release and possibly also cause changes in chemical equilibrium 5. Conclusions
for combustion of methane and CO. An intensively tested Ni-based oxy-
gen carrier is NiO on NiAl2O4 produced by spray drying. Jerndal et al. The fuel conversion and physical properties of 17 spray-dried com-
[40] have investigated different manufacturing parameters for this car- bined Cu–Fe and Mn–Ni oxygen carriers were investigated. Parameters
rier and have determined a crushing strength between 1.2 and 1.9 N varied during material production were the material composition and
(same NiO quality as used in this manuscript: Novamet Green grade the calcination temperature.
F) and a CO2 yield for combustion of methane between 0.8 and 0.99. Regarding conversion of syngas all materials showed an almost full
Ni33Mn_1200, which can be best used for comparison (similar NiO con- gas conversion. With respect to conversion of methane, several material
tent and calcination temperature), has an analog crushing strength and combinations were identified with a CO2 yield of N90% at 950 °C. Excep-
seems to be superior with respect to methane conversion. tionally high CO2 yields were measured for the Ni–Mn-based materials
With respect to physical properties, it was established that the calci- produced from 34 wt.% NiO which is equal or even better than compa-
nation temperature is decisive for bulk density, specific surface area and rable materials presented in literature.
crushing strength i.e. the mechanical durability. Attrition tests on used Significant CLOU oxygen release was detected for all Cu-Fe materials
Mn–Ni oxygen carriers were not performed because of their toxicity. but just for one material combination of Mn–Ni which could be ex-
Anyhow, it is likely that the Mn–Ni particles retained their mechanical plained by the corresponding phase diagrams.
strength throughout activity testing as their physical properties By testing mechanical durability and investigation of surface mor-
remained fairly constant. The same can be concluded when looking at phology before and after fluidized-bed testing, promising candidates
the electron microscope images as they show a similar particle mor- for application in chemical-looping combustion were identified. Com-
phology in fresh and used state. pared to similar materials investigated by other researchers, an im-
From a mechanical and reactivity point of view, it can be concluded proved mechanical strength has been acquired for some of the
that among Mn–Ni-based materials, Ni33Mn_1200 is most suitable for copper- and nickel-based materials.
conversion of gaseous fuels as it fully converted both methane and syn- Especially noteworthy with respect to reactivity and mechanical du-
gas (CO + H2). Nevertheless, also Ni5Mn_1200 and Ni10Mn_1200 are rability are the two oxygen carriers produced from 30 wt.% CuO and
interesting candidates as they had a high crushing strength and high sintered at 950/1025 °C as well as the carrier produced from 34 wt.%
methane as well as full syngas conversion. NiO and calcined at 1200 °C.
40 V. Frick et al. / Fuel Processing Technology 150 (2016) 30–40

Acknowledgement [22] M. Rydén, D. Jing, M. Källén, H. Leion, A. Lyngfelt, T. Mattisson, CuO-based oxygen-
carrier particles for chemical-looping with oxygen uncoupling — experiments in
batch reactor and in continuous operation, Ind. Eng. Chem. Res. 53 (2014)
This study has been funded by the Swedish Energy Agency and 6255–6267, http://dx.doi.org/10.1021/ie4039983.
Preem AB. The authors gratefully acknowledge the financial support. [23] P. Cho, T. Mattisson, A. Lyngfelt, Comparison of iron-, nickel-, copper- and
manganese-based oxygen carriers for chemical-looping combustion, Fuel 83
(2004) 1215–1225, http://dx.doi.org/10.1016/j.fuel.2003.11.013.
References [24] M. Arjmand, A.-M. Azad, H. Leion, T. Mattisson, A. Lyngfelt, Evaluation of CuAl2O4 as
an oxygen carrier in chemical-looping combustion, Ind. Eng. Chem. Res. 51 (2012)
[1] M. Ishida, H. Jin, A novel chemical-looping combustor without NOx formation, Ind. 13924–13934, http://dx.doi.org/10.1021/ie300427w.
Eng. Chem. Res. 35 (1996) 2469–2472, http://dx.doi.org/10.1021/ie950680s. [25] L.F. de Diego, P. Gayán, F. García-Labiano, J. Celaya, A. Abad, J. Adánez, Impregnated
[2] A. Lyngfelt, B. Leckner, T. Mattisson, A fluidized-bed combustion process with inher- CuO/Al2O3 oxygen carriers for chemical-looping combustion: avoiding fluidized bed
ent CO2 separation; application of chemical-looping combustion, Chem. Eng. Sci. 56 agglomeration, Energy Fuel 19 (2005) 1850–1856, http://dx.doi.org/10.1021/
(2001) 3101–3113. ef050052f.
[3] IEA, CO2 Emissions Overview, International Energy Agency, Paris, 2012 2014. [26] I. Adánez-Rubio, P. Gayán, A. Abad, L.F. de Diego, F. García-Labiano, J. Adánez, Eval-
[4] A. Lyngfelt, Chemical-looping combustion of solid fuels — status of development, uation of a spray-dried CuO/MgAl2O4 oxygen carrier for the chemical looping with
Appl. Energy 113 (2014) 1869–1873, http://dx.doi.org/10.1016/j.apenergy.2013. oxygen uncoupling process, Energy Fuel 26 (2012) 3069–3081, http://dx.doi.org/
05.043. 10.1021/ef3002229.
[5] A. Lyngfelt, B. Leckner, A 1000 MWth boiler for chemical-looping combustion of [27] P. Gayán, C.R. Forero, A. Abad, L.F. de Diego, F. García-Labiano, J. Adánez, Effect of
solid fuels — discussion of design and costs, Appl. Energy 157 (2015) 475–487, support on the behavior of Cu-based oxygen carriers during long-term CLC opera-
http://dx.doi.org/10.1016/j.apenergy.2015.04.057. tion at temperatures above 1073 K, Energy Fuel 25 (2011) 1316–1326, http://dx.
[6] I. Abdulally, C. Beal, H. Andrus, B. Epple, A. Lyngfelt, B. Lani, Alstom's chemical doi.org/10.1021/ef101583w.
looping prototypes program update, Proc. 37th Int. Tech. Conf. Clean Coal Fuel [28] B. Wang, W. Wang, Q. Ma, J. Lu, H. Zhao, C. Zheng, In-depth investigation of chemical
Syst., 2012. looping combustion of a Chinese bituminous coal with CuFe2O4 combined oxygen
[7] L.-S. Fan, L. Zeng, W. Wang, S. Luo, Chemical looping processes for CO2 capture and carrier, Energy Fuel (2016).
carbonaceous fuel conversion — prospect and opportunity, Energy Environ. Sci. 5 [29] R.V. Siriwardane, E. Ksepko, H. Tian, J. Poston, T. Simonyi, M. Sciazko, Interaction of
(2012) 7254, http://dx.doi.org/10.1039/c2ee03198a. iron–copper mixed metal oxide oxygen carriers with simulated synthesis gas de-
[8] J. Adanez, A. Abad, F. Garcia-Labiano, P. Gayan, L.F. de Diego, Progress in chemical- rived from steam gasification of coal, Appl. Energy 107 (2013) 111–123, http://dx.
looping combustion and reforming technologies, Prog. Energy Combust. Sci. 38 doi.org/10.1016/j.apenergy.2013.01.063.
(2012) 215–282, http://dx.doi.org/10.1016/j.pecs.2011.09.001. [30] C. Linderholm, T. Mattisson, A. Lyngfelt, Long-term integrity testing of spray-dried
[9] A. Lyngfelt, Oxygen carriers for chemical-looping combustion, Calcium Chem. particles in a 10-kW chemical-looping combustor using natural gas as fuel, Fuel
Looping Technol. Power Gener. Carbon Dioxide CO2 Capture, Woodhead Publishing 88 (2009) 2083–2096, http://dx.doi.org/10.1016/j.fuel.2008.12.018.
Limited 2015, p. 390. [31] F. García-Labiano, L.F. de Diego, P. Gayán, J. Adánez, A. Abad, C. Dueso, Effect of fuel
[10] H. Leion, T. Mattisson, A. Lyngfelt, Solid fuels in chemical-looping combustion, Int. J. gas composition in chemical-looping combustion with Ni-based oxygen carriers. 1.
Greenhouse Gas Control 2 (2008) 180–193. Fate of sulfur, Ind. Eng. Chem. Res. 48 (2009) 2499–2508, http://dx.doi.org/10.
[11] A. Lyngfelt, T. Mattisson, Materials for chemical-looping combustion, in: D. Stolten, 1021/ie801332z.
V. Scherer (Eds.), Effic. Carbon Capture Coal Power Plants, WILEY-VCH Verlag [32] I. Díaz-Castro, K. Mayer, T. Pröll, H. Hofbauer, Effect of sulfur on chemical-looping
GmbH & Co. KGaA, Weinheim 2011, pp. 475–504. combustion of natural gas using a nickel-based oxygen carrier, 21st Int. Conf.
[12] H. Leion, T. Mattisson, A. Lyngfelt, The use of petroleum coke as fuel in chemical- Fluid. Bed Combust., Naples, 2012.
looping combustion, Fuel 86 (2007) 1947–1958. [33] D. Jing, T. Mattisson, H. Leion, M. Rydén, A. Lyngfelt, Examination of perovskite
[13] T. Mattisson, Materials for chemical-looping with oxygen uncoupling, ISRN Chem. structure CaMnO3-d with MgO addition as oxygen carrier for chemical looping
Eng. 2013 (2013), e526375, http://dx.doi.org/10.1155/2013/526375. with oxygen uncoupling using methane and syngas, Int. J. Chem. Eng. 2013
[14] E.M. Eyring, G. Konya, J.S. Lighty, A.H. Sahir, A.F. Sarofim, K. Whitty, Chemical looping (2013), e679560, http://dx.doi.org/10.1155/2013/679560.
with copper oxide as carrier and coal as fuel, Oil Gas Sci. Technol. – Rev. IFP Energ. [34] D. Kunii, O. Levenspiel, Fluidization Engineering, vol. 2, Butterworth-Heinemann
Nouv. 66 (2011) 209–221, http://dx.doi.org/10.2516/ogst/2010028. Boston, 1991.
[15] T. Mattisson, H. Leion, A. Lyngfelt, Chemical-looping with oxygen uncoupling using [35] M. Rydén, P. Moldenhauer, S. Lindqvist, T. Mattisson, A. Lyngfelt, Measuring attrition
CuO/ZrO2 with petroleum coke, Fuel 88 (2009) 683–690, http://dx.doi.org/10.1016/ resistance of oxygen carrier particles for chemical looping combustion with a cus-
j.fuel.2008.09.016. tomized jet cup, Powder Technol. 256 (2014) 75–86, http://dx.doi.org/10.1016/j.
[16] A. Abad, I. Adánez-Rubio, P. Gayán, F. García-Labiano, L.F. de Diego, J. Adánez, Dem- powtec.2014.01.085.
onstration of chemical-looping with oxygen uncoupling (CLOU) process in a [36] A.V. Khvan, O.B. Fabrichnaya, G. Savinykh, R. Adam, H.J. Seifert, Thermodynamic as-
1.5 kWth continuously operating unit using a Cu-based oxygen-carrier, Int. J. Green- sessment of the Cu–Fe–O system, J. Phase Equilib. Diffus. 32 (2011) 498–511, http://
house Gas Control 6 (2012) 189–200, http://dx.doi.org/10.1016/j.ijggc.2011.10.016. dx.doi.org/10.1007/s11669-011-9951-5.
[17] Q. Zafar, A. Abad, T. Mattisson, B. Gevert, M. Strand, Reduction and oxidation kinetics [37] Y.V. Golikov, V.F. Balakirev, Phase equilibrium diagram of the system—Ni–Mn–O, J.
of Mn3O4/Mg–ZrO2 oxygen carrier particles for chemical-looping combustion, Phys. Chem. Solids 49 (1988) 329–332, http://dx.doi.org/10.1016/0022-
Chem. Eng. Sci. 62 (2007) 6556–6567, http://dx.doi.org/10.1016/j.ces.2007.07.011. 3697(88)90087-X.
[18] A. Abad, T. Mattisson, A. Lyngfelt, M. Rydén, Chemical-looping combustion in a [38] C.K. Clayton, K.J. Whitty, Measurement and modeling of decomposition kinetics for
300 W continuously operating reactor system using a manganese-based oxygen copper oxide-based chemical looping with oxygen uncoupling, Appl. Energy 116
carrier, Fuel 85 (2006) 1174–1185. (2014) 416–423, http://dx.doi.org/10.1016/j.apenergy.2013.10.032.
[19] M. Rydén, H. Leion, T. Mattisson, A. Lyngfelt, Combined oxides as oxygen-carrier ma- [39] E. Jerndal, T. Mattisson, A. Lyngfelt, Thermal analysis of chemical-looping combus-
terial for chemical-looping with oxygen uncoupling, Appl. Energy 113 (2014) tion, Chem. Eng. Res. Des. 84 (2006) 795–806.
1924–1932, http://dx.doi.org/10.1016/j.apenergy.2013.06.016. [40] E. Jerndal, T. Mattisson, I. Thijs, F. Snijkers, A. Lyngfelt, Investigation of NiO/NiAl2O4
[20] A. Shulman, E. Cleverstam, T. Mattisson, A. Lyngfelt, Manganese/iron, manganese/ oxygen carriers for chemical-looping combustion produced by spray-drying, Int. J.
nickel, and manganese/silicon oxides used in chemical-looping with oxygen Greenhouse Gas Control 4 (2010) 23–35, http://dx.doi.org/10.1016/j.ijggc.2009.09.
uncoupling (CLOU) for combustion of methane, Energy Fuel 23 (2009) 5269–5275. 007.
[21] A. Shulman, E. Cleverstam, T. Mattisson, A. Lyngfelt, Chemical-looping with oxygen
uncoupling using Mn/Mg-based oxygen carriers — oxygen release and reactivity
with methane, Fuel 90 (2011) 941–950, http://dx.doi.org/10.1016/j.fuel.2010.11.
044.

You might also like