+ All Categories
Home > Documents > Integrating National Research Agendas on Solar Heat for ...

Integrating National Research Agendas on Solar Heat for ...

Date post: 11-Nov-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
66
D 4.1 GA No. 731287 1 Co-funded by the Horizon 2020 Framework Programme of the European Union Integrating National Research Agendas on Solar Heat for Industrial Processes Project Deliverable D 4.1 Mass/energy balances and solar chemical reactor concepts D 4.1 MASS/ENERGY BALANCES AND SOLAR CHEMICAL REACTOR CONCEPTS WP 4 - Technology and applications to high temperature SHIP (400 °C to 1500 °C) Due date: M24 (December 2018) Submitted M29 (May 2019) Partner responsible D4.1 ETHZ Person responsible Vikas Patil, Aldo Steinfeld Reviewed/supervised by: Vikas Patil, Aldo Steinfeld GA Number 731287 Start of the project January 2017 Duration of the project 48 months (January 2017-December 2020) DISSEMINATION LEVEL PU (Public) NATURE OF THE DELIVERABLE R (Report) HISTORY Author Date Comments Vikas Patil 08.02.2019 First draft submitted Vikas Patil 17.05.2019 Revisions incorporated Vikas Patil 21.05.2019 Draft accepted & finalized Ref. Ares(2019)3349886 - 22/05/2019
Transcript

D 4.1 GA No. 731287 1 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Integrating National Research Agendas on Solar Heat

for Industrial Processes

Project Deliverable D 4.1

Mass/energy balances and solar chemical

reactor concepts

D 4.1 – MASS/ENERGY BALANCES AND SOLAR CHEMICAL REACTOR CONCEPTS

WP 4 - Technology and applications to high temperature SHIP (400 °C to

1500 °C)

Due date: M24 (December 2018)

Submitted M29 (May 2019)

Partner responsible D4.1 ETHZ

Person responsible Vikas Patil, Aldo Steinfeld

Reviewed/supervised by: Vikas Patil, Aldo Steinfeld

GA Number 731287

Start of the project January 2017

Duration of the project 48 months (January 2017-December 2020)

DISSEMINATION LEVEL

PU (Public)

NATURE OF THE DELIVERABLE

R (Report)

HISTORY

Author Date Comments

Vikas Patil 08.02.2019 First draft submitted

Vikas Patil 17.05.2019 Revisions incorporated

Vikas Patil 21.05.2019 Draft accepted & finalized

Ref. Ares(2019)3349886 - 22/05/2019

Deliverable Report

D 4.1 GA No. 731287 2 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Contents

1. Content of this Deliverable .......................................................................................................................... 3

2. Solar metals production for the metallurgical industry .......................................................................... 3

2.1. Aluminum ................................................................................................................................................ 3

2.2. Magnesium ........................................................................................................................................... 12

2.3. Copper .................................................................................................................................................. 15

2.4. Tin ............................................................................................................................................................ 18

3. Lime processing from limestone ................................................................................................................ 26

3.1. Solar hybridization for cement production .................................................................................... 26

3.2. Particle based gravity driven solar receiver development for solar lime ................................. 33

3.3. Thermal performance and transient operation of a solar receiver .......................................... 36

4. Solar fuels production from H2O and CO2 via redox cycles ................................................................ 39

4.1. Solar receiver–reactor for thermochemical redox cycling using ceria .................................... 41

4.2. Thermal performance of a lab-scale ceramic volumetric air receiver .................................... 44

4.3. Mass and energy balances expected for a metallic volumetric receiver .............................. 47

5. Fuels production from agricultural waste via gasification ................................................................... 51

5.1. Solar reactor concepts for gasification of carbonaceous materials ........................................ 51

5.2. Co-production of syngas and potassium-based fertilizer by solar-driven thermochemical

conversion of crop residues ........................................................................................................................... 53

6. Degree of progress ...................................................................................................................................... 57

7. Dissemination ................................................................................................................................................ 58

8. References .................................................................................................................................................... 59

Deliverable Report

D 4.1 GA No. 731287 3 Co-funded by the Horizon 2020 Framework

Programme of the European Union

1. Content of this Deliverable

Solar thermochemical processes make use of concentrated solar radiation as the energy source of

process heat to drive endothermic reactions. Three categories of high-temperature energy-

intensive processes are considered: i) solar metallurgy, ii) solar lime, and iii) solar fuels. To provide the

high-temperature solar reactors with highly concentrated sunlight, specific solar optics

technologies are required.

Work Package 4 of the INSHIP project (WP4: Technology and applications to high temperature SHIP

(400 °C to 1500 °C)) aims at advancing the development of high temperature/high-flux solar

chemical reactor technologies and concepts. One of the objectives is feasibility assessment of solar

chemical reactor technologies for metallurgical processes, lime production and solar fuels

production.

In line with this objective, this Deliverable report consolidates analyses carried out by various

institutes under the INSHIP consortium (until the end of year 2018) – both theoretical as well as

experimental – related to the mass and energy balances of the targeted high-temperature

industrial processes and related solar chemical reactor concepts.

2. Solar metals production for the metallurgical industry

The extractive metallurgical industry is a major consumer of high-temperature process heat. It is,

consequently, a major contributor of CO2 emissions derived from the combustion of fossil fuels for

heat and electricity generation. The amount of fuel needed to reduce metal oxides to metals

could be substantially reduced if process heat were supplied by an alternative clean energy

source. Concentrated solar thermal energy can provide the high-temperature process heat

required to drive these highly endothermic reactions. The candidate processes are the productions

of aluminum, magnesium, copper and tin from their ores.

The following sections provide an overview of the work carried out by INSHIP members under Task

4.1 of WP4 (Solar metals production for the metallurgical industry).

2.1. Aluminum

Aluminum production can be considered as a two-stage process. The first stage is from ore to

alumina purification and smelting. The second stage is from alumina to metallic aluminum. The most

common process for the first stage is the Bayer process while for the second stage the most

common one is the Hall-Heroult process by electrochemical reduction. The Bayer process produces

high-purity alumina which is essential for the downstream Hall-Heroult electrolytic process. However,

an important fall-out of this process is the toxic “red mud”, which is one of the most challenging

disposal problems.

Alternatively, the Bayer process can be substituted by a calcination process, and the Hall-Heroult

electrolytic process can be substituted by a carbothermic reduction process. Both thermochemical

processes are energy intensive and proceed at high temperatures, and thus can be driven by solar

process heat, as described in the following sub-sections.

Deliverable Report

D 4.1 GA No. 731287 4 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2.1.1. Solar hybridization paths for the Bayer process

This section has been prepared by the Middle East Technical University (METU), Turkey.

We targeted Seydisehir Aluminum plant in Konya Province of Turkey for mass and energy balances.

We could not get any response to our direct data collection attempts from the plant. However, we

obtained enough data for modelling the plant from a METU thesis dealing with Alumina production

in Seydisehir plant by Kurucak [1] and from a conference paper written by the plant personnel,

Avcu et al. [2]. Seydisehir plant produces Alumina and various Aluminum products but the thesis [1]

only deals with the Bayer process for alumina production.

Solar hybridization of both the Bayer and Hall-Heroult processes may be possible targeting their high

temperature heat input requirements. However, at METU, we plan to concentrate on identifying

solar hybridization paths for the Bayer process which is the most heat intensive one between the

two processes.

We plan to model Seydisehir Aluminum plant Bayer process using the obtained mass and energy

flows among the components of the plant. In the thesis [1], there is a review of process modelling

studies that can be useful in development of our own models. In addition to several studies dealing

with mass and energy balances of alumina production, there are also many other studies

concentrating on reduction of alumina (e.g. Obaidat et al. [3], Gusberti et al. [4] and Balomenos et

al. [5][6]) and these will also be useful in our modelling studies.

The model of the plant is expected to be useful in identification of solar hybridization options for the

plant. After confirming the accuracy of simulation results, solar input at the plant location will be

obtained for decided time periods and time dependent simulations with and without solar

hybridization will be done. From the comparison of the results of simulations, we expect to see how

useful solar hybridization will be in reducing fossil fuel use of the plant.

Table 1 Flow rates of major streams in Figure 1(from [2])

Major streams Stream number in Figure Flow rates (kg/h)

Bauxite S01 48826

Strong liquor S02 7396

Wash water S15 120000

Red mud S27 60626

Hydrate (Al(OH)3) S40 36686

Table 2 Composition of Bauxite (data from [2])

Al2O3 55.63%

SiO2 6.86%

CO2 0.83%

CaO 1.20%

Fe2O3 17.39%

Amorph 3.27%

H2O(s) 12.36%

H2O(l) 2.47%

Deliverable Report

D 4.1 GA No. 731287 5 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Fig

ure

1:

Flo

w s

he

et

of

Se

yd

ise

hir A

lum

inu

m P

lan

t (f

rom

[2])

Deliverable Report

D 4.1 GA No. 731287 6 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 2 Heating and evaporation flow sheet [2]

Table 3 Descriptions of streams and units in heating and evaporation flow sheet [1]

STREAMS Description UNITS Description

S41 Spent liquor from precipitation U-23 Heat exchanger

S43 Spent liquor to digestion U-24 Flash drum

S44 Steam U-25 Mixer

S46 Water

S47 Steam

S48 Steam

Table 4 Mass balance for heating and evaporation section [1]

Inlet streams (kg/h) Outlet streams (kg/h)

S41 S47 S48 TOTAL S43 S44 S46 TOTAL

Al2O3 0 0 0 0 0 0 0 0

SiO2 0 0 0 0 0 0 0 0

Na2O 0 0 0 0 0 0 0 0

CO2 0 0 0 0 0 0 0 0

CaO 0 0 0 0 0 0 0 0

Fe2O3 0 0 0 0 0 0 0 0

Amorph 0 0 0 0 0 0 0 0

H2O(s) 0 0 0 0 0 0 0 0

NaAlO2 25626 0 0 25626 25626 0 0 25626

NaOH 25812 0 0 25812 25812 0 0 25812

Na2CO3 105574 0 0 105574 105574 0 0 105574

H2O 185654 10116 21854 217624 108793 76861 31970 217624

TOTAL 342666 10116 21854 374636 265805 76861 31970 374636

Deliverable Report

D 4.1 GA No. 731287 7 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2.1.2. Solar calcination of alumina

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [7].

Alumina is an intermediate product in the production of aluminum and is also a product in its own

right. Alumina refining accounts for approximately 27% of the primary energy used in aluminum

production, over 90% of which is provided by the combustion of fossil fuels [8]. A total of 115.2 million

tons of alumina was produced globally in 2015 at an average energy intensity of 11.9 GJ/ton-

alumina, representing a significant contribution to GHG emissions [9]. Calcination is the last step of

the Bayer process, which heats the hydrated alumina (aluminum hydroxide or gibbsite, Al(OH)3)

from the precipitation stage, to drive off the water of hydration and form anhydrous alumina

(aluminum oxide, Al2O3). It is an energy-intensive process, which has historically been conducted

with the use of fossil fuels and a rotary-kiln [10]. However, since the oil crisis of the 1970s, with the

resulting increase in the cost of fuels, stationary calciners (circulating fluidised bed, gas suspension

and fluid-flash calciners) have been favored over rotary-kilns due to their significantly reduced fuel

consumption [11,12]. The process heat requirements of modern industrial calciners and estimated

resulting CO2 emissions (calculated from the combustion of methane) are now approximately

3 GJ/ton-alumina and 165 kg-CO2/ton-alumina product, respectively, for which the predominant

fuel is natural gas [13,14]. Today, the increasing cost of fuels continues to be a driver for

technological development of more fuel efficient alumina calciner technology [15]. With the price

of natural gas forecast to increase in some locations such as Australia [13], and with the global

need to reduce GHG emissions, there is strong incentive to seek alternative energy sources such as

CST for process heat [16]. This provides the motivation for ongoing investigations.

The calcination of alumina is a thermal decomposition chemical reaction, which proceeds

endothermically according to the following reaction:

( 1 )

The enthalpy change of this reaction is 185 kJ/mol. Modern plants typically process alumina

(generally in the gibbsite form) within flash calciners using particles of ~100 µm in diameter

transported in a gas suspension through the reactor with residence times on the order of a few

seconds. A solar reactor that processes powders of similar size to a flash calciner is the solar vortex

transport reactor. This employs a vortex flow to transport particles in suspension through an

irradiation zone where they undergo high heating rates to drive a thermochemical reaction. Direct

solar irradiation of such a suspension of reacting particles is an efficient means of heat transfer to

the reactants [17–20]. Furthermore, the temperature at which alumina calcination takes place in

the Bayer process, 1273 – 1373 K, is readily achievable in current commercially available solar tower

concentrators [16] and is significantly less than the 1750 K that has been achieved previously in the

solar vortex transport reactor used for solar gasification [21]. The potential for alumina to be

calcined with CST has been assessed with a packed bed of boehmite (an aluminum oxyhydroxide)

in a crucible positioned at the focal plane of a Fresnel concentrator by Padilla et al. [22]. They

reported 75% conversion after 10 minutes of exposure to solar radiation concentrated to 2644 suns

and full conversion with exposure time of 90 minutes. However, to date, no assessments have been

reported for the calcination of gibbsite in a practical reactor under conditions relevant to flash

calcination. The industrial gibbsite calcination process shares similarities with the calcination of

limestone, which takes place at ~1150 K and has previously been demonstrated with the

application of CST [23–26]. This demonstration has been performed with direct irradiation in a

cavity-receiver transporting a flow of entrained particles [23], which suggests that it may also be

Deliverable Report

D 4.1 GA No. 731287 8 Co-funded by the Horizon 2020 Framework

Programme of the European Union

suitable for alumina calcination. It is also noteworthy that the alumina calcination process is

sensitive to the gas-phase composition, which is one reason that natural gas is a preferred fuel for

alumina production [27,28]. The application of CST to the calcination process has the potential to

further improve product quality by eliminating the possibility of contamination with combustion

products and by reducing the significance of back-reactions from the presence of H2O, which is a

product of combustion. While these reasons suggest that the alumina calcination process is

potentially well suited to solar processing, its technical feasibility cannot be determined by

implication. Due to the complex, coupled nature of the process, it is necessary to assess this

experimentally with a prototype solar reactor.

In summary, the calcination stage of alumina refining with the Bayer process is potentially well

suited to the application of CST because it is a highly endothermic process that proceeds at

~1273 K, which is readily achievable with solar tower technology. In addition, the calcination

process is not sensitive to some ingress of air, which raises the possibility that direct irradiation of

particles may be achievable without the need for a window to completely seal a solar reactor.

2.1.3. Solar carbothermic reduction of alumina

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [29].

The production of aluminum via the Hall-Héroult electrolytic process is characterized by its high

energy consumption, approximately 45 GJ/ton Al, and its associated high specific greenhouse gas

emissions, 4.9-7.4 ton CO2-equiv/ton Al [30–32]. The alternative path via carbothermal reduction of

Al2O3 can be represented by the overall reaction:

Al2O3 + 3C = 2Al + 3CO ( 2 )

The enthalpy change of this reaction is 1345 kJ/mol. However, reaction (2) is complicated by the

formation of undesired byproducts such as aluminum carbide Al4C3, oxy-carbides Al2OC and

Al4O4C, and the lower valence oxide Al2O [33,34]. One approach to overcome this difficulty was to

reduce the Al2O-Al-CO gas mixture by carbon at about 2273 K (2000oC) [35]. According to the Le

Chatelier principle, the extent of the chemical reaction should be favored by a decrease in the

total gas pressure. Thus, under vacuum conditions, the equilibrium of reaction (2) should be shifted

to the right and the onset temperature for the metal production should be significantly lowered.

This principle had been successfully applied in the production of magnesium and silicon. The

carbothermal reduction of MgO in calcined dolomite yielded Mg using a moderate vacuum of

1.33 millibar at 1673 K (1400°C) [36]. The silicothermic reduction of CaO-MgO in the industrial

Pidgeon process for magnesium was achieved by reacting CaO-MgO with ferrosilicon in

evacuated retorts at 1450-1800 K (1177-1527°C) and less than 1.33·10-4 bar [37–39]. The production

of Si by vacuum carbothermal reduction of SiO2 in the range 1997–2263 K (1724-1990°C) at ~3×10-3

bar yielded Si/SiO2 mixtures condensed from the gas phase, with 79 wt.% Si content [40].

As it will be shown in the analysis that follows, reaction (2) is highly endothermic (ΔH°298K = 1344

kJ/mol) and proceeds at above about 1500 K (1227°C) when operating at 10-4 bar. The use of

concentrated solar energy as the source of high-temperature process heat reduces significantly

the discharge of greenhouse gases and other pollutants derived from the combustion of fossil fuels

[41,42]. Previous relevant metallurgical processes performed in solar furnaces include the

carbothermal and methanothermal reductions of Fe3O4, MgO, ZnO, and SiO2 to produce Fe, Mg,

Zn, and Si, respectively [40,43–46], the carbothermal reductions of Al2O3, CaO, SiO2, and TiO2 in an

Ar flow to produce Al4C3, CaC2, SiC, and TiC, respectively [46], and the carbothermal reductions of

Deliverable Report

D 4.1 GA No. 731287 9 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Al2O3, SiO2, TiO2, and ZrO2 in a N2 flow to produce AlN, Si3N4, TiN, and ZrN, respectively [46–48]. The

present study examines thermodynamically the vacuum carbothermal reduction of Al2O3, and

demonstrates experimentally the production of Al using a biomass-based reducing agent and

simulated concentrated solar energy.

Thermodynamic Analysis

Thermochemical equilibrium computations were carried out using the FactSage code [49].

Reaction enthalpies were calculated using NIST webbook data [50] and HSC Outokumpu code

[51]. The equilibrium compositions as a function of temperature for the system Al2O3 + 3C at 1, 10-5,

and 10-6 bar are shown in Figures Figure 3-Figure 5. As observed in Figure 3, at atmospheric pressure,

the production of elementary Al(g) essentially free of Al2O and Al4C3 is thermodynamically

favorable above 2800 K (2527°C), i.e. at a practically prohibitive high temperature. Figures Figure 4

and Figure 5 show that lowering the gas pressure to 10-5 or 10-6 bar not only shifts the formation of

Al(g) to much lower temperatures, but also frees the Al(g) production from the undesired by-

products Al2O and Al4C3 already at more accessible temperatures. The temperatures for the onset

of Al formation at about 0.001% conversion of Al2O3 to Al for gas pressures of 10-6 to 1 bar are listed

in Table 5. At 1800 K (1527°C) and 10-4 bar, the system in equilibrium is mainly composed of a gas

mixture containing 40% Al(g) - 60% CO(g), with small traces of Al2O(g).

In a preliminary environmental evaluation of the potential of such a vacuum thermochemical

process it is assumed that either concentrated solar radiation or coal combustion are applied as

the energy source of process heat. Further, it is proposed that the CO released would be water-gas

shifted to syngas, followed by methanol synthesis, as shown in Table 6. The total enthalpy change of

the reaction was calculated in two steps: 1) for the stoichiometric reaction Al2O3(s) + 3C(gr) = 2Al(g)

+ 3CO(g) at 1 bar and 1800oK (1527°C), ΔH = 1941.1 kJ/mol Al2O3; 2) for the theoretical work of

isothermal expansion of the product gases at 1800 K (1527°C) from 1 to 10-4 bar, ΔG = nRTlnp = 698

kJ/mol Al2O3. Thus, the total theoretical heat and work input would amount to 49 GJ/ton Al, and

comparable to that in the Hall-Héroult electrolytic process. In case the reducing agent is derived

from a biomass source, the solar-driven carbothermal reduction would be CO2 neutral.

Table 5 Equilibrium temperatures for the onset of Al formation at 10-5 mole fraction for the system Al2O3 + 3C.

Extracted from Ref. [29].

Pressure [bar] Temperature [K]

10-6 1370 10-5 1465

10-4 1580

0.001 1703

0.01 1855

0.1 2037

0.5 2190

1.0 2258

Deliverable Report

D 4.1 GA No. 731287 10 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Table 6 Preliminary evaluation of the annual co-production Al and CH3OH by vacuum carbothermic reduction

of Al2O3 from a mixture of Al2O3 and coke, initially at 300 K (27°C) and 1 bar, transformed at 1800 K and 10-4 bar

according to Equation 2. Process heat supplied either by concentrated solar energy or by coal combustion.

Extracted from Ref. [29].

Design Al2O3 feed (kmol/yr) 1.00x106

Annual process heat (GJ/yr) 1.96x106

Annual coke feed (ton/yr) 1.18x106

Annual coal feed (ton/yr) 1 0.048x106

Annual pumping energy (GJ/yr) 2 0.84x106

Total annual fuel input (GJ/yr) 3 2.02x106

Total annual fuel input (GJ/yr) 1 3.98x106

Annual Al production (ton/yr) 4 0.040x106

Annual methanol production (ton/yr) 5 0.0288x106

Total annual CO2 release (ton/yr) 3 0.088x106

Total annual CO2 release (ton/yr) 1 0.26x106

1 Assuming coal combustion for process heat. 2 Assuming 50% efficiency in electricity production. 3 Assuming solar energy for process heat; Ton = metric ton. 4 Assuming 75% overall yield in production of Al. 5 Assuming 90% overall yield in the water-gas shift of 2/3 of CO to syngas, converted to methanol.

0

1

2

3

2200 2400 2600 2800 3000

Eq

uilib

riu

m c

om

po

sit

ion

[m

ol]

Temperature [K]

CO(g)

Al(g)

Al2O(g)

C(gr)

Al4C3(s)

Al2O3(s)

Figure 3: Variation of the equilibrium composition as a function of temperature for the system Al2O3 + 3C at 1

bar. Extracted from Ref. [29].

Deliverable Report

D 4.1 GA No. 731287 11 Co-funded by the Horizon 2020 Framework

Programme of the European Union

0

1

2

3

1400 1500 1600 1700

Eq

uilib

riu

m c

om

po

sit

ion

[m

ol]

Temperature [K]

CO (g)

Al (g)

Al2O (g)

C(gr)

Al4C3 (s)

Al2O3 (s)

Figure 4: Variation of the equilibrium composition as a function of temperature for the system Al2O3

+ 3C at 10-5 bar. Extracted from Ref. [29].

0

1

2

3

1300 1350 1400 1450 1500

Eq

uilib

riu

m c

om

po

sit

ion

[m

ol]

Temperature [K]

CO(g)

Al(g)

Al2O(g)

C(gr)

Al4C3(s)

Al2O3(s)

Figure 5: Variation of the equilibrium composition as a function of temperature for the system Al2O3 + 3C at

10-6 bar. Extracted from Ref. [29].

Deliverable Report

D 4.1 GA No. 731287 12 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2.2. Magnesium

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [52].

Mg produced by electrolytic reduction of magnesium chloride is said to consume 40% more energy

per unit weight of metal as compared to Al production, ranging from 10-20 kWh per kg Mg, which

translates to at least 9 kg CO2 emitted per kg Mg produced when using coal-fired electricity. During

the last 20 years the Pidgeon process achieved a strong revival, particularly in China as the major

producer of Mg, reaching 72.5% of the world primary production of this metal [39]. However, this

process is plagued by severe environmental pollution along with heavy CO2 emissions [53]. Its

global warming impact has been estimated by a detailed life-cycle study to be 42 kg CO2-eq/kg

Mg [39]. Replacing fossil fuels by solar process heat for bringing the reactants to the operating

reactor temperature and for providing the enthalpy change of the reaction would significantly

reduce greenhouse gases emissions. CO2 emissions of a solar-driven process are those derived only

from the use of a carbonaceous reducing agent and should not exceed - assuming stoichiometric

carbothermic reduction - 1.83 kg CO2 per kg Mg, resulting in a 5-fold reduction vis-à-vis the

conventional fossil-driven electrolytic production. Furthermore, since the product gases evolved are

not contaminated by combustion byproducts, they are sufficiently valuable commodities to justify

their collection, eliminating inherent gas emissions to the environment.

The alternative path via carbothermal reduction of MgO can be represented by the overall

reaction:

MgO + C = Mg + CO ( 3 )

The enthalpy change of this reaction is 491 kJ/mol. Direct carbothermic reduction has been

developed at pilot scale by CSIRO in Australia [54,55] and by researchers in the USA [56]. The

availability of cheap carbon sources such as petroleum coke, charcoal, or char and the potential

capability of continuous operation are the main advantages of this process. Main challenge

concerns the quenching of the gaseous product necessary to prevent the formation of MgO

through the reverse reaction between magnesium vapor and carbon monoxide.

Thermodynamics Analysis

Thermochemical equilibrium computations were performed using the HSC Outokumpu software

code. Species whose mole fractions were less than 10-5 have been omitted from the graphs. (a &

b) shows the equilibrium compositions of the system MgO+C and MgO+CH4, respectively, at 1 bar

and over a wide range of temperatures of interest. In both cases, the formation of stable carbides

or carbonates of Mg is not thermodynamically favored. When using C as reducing agent, MgO

undergoes carbothermic reduction at above 1750ºC, reaching completion at 1765ºC and

producing an equimolar gas mixture of Mg vapor and CO. When using CH4 as reducing agent, CH4

thermally decomposes into its elements at above 200ºC, which in turn reduce MgO at above

1500ºC, reaching completion at 1800ºC and producing a gas mixture of 25% Mg(g), 50% H2, 25%

CO, with traces (< 130 ppm) of CH4. MgH, and C2H2.

MgO + CH4 = Mg + CO + 2H2 ( 4 )

In principle, the separation of Mg(g) from CO or syngas can be accomplished by condensing the

product gases below its saturation point (b.p. 1107ºC). Once free of Mg, the syngas mixture may be

Deliverable Report

D 4.1 GA No. 731287 13 Co-funded by the Horizon 2020 Framework

Programme of the European Union

cooled to ambient temperatures - by using heat exchangers to pre-heat reactants - and

recuperate part of the sensible heat of the products without fear of unwanted recombination

between H2 and CO since the reverse, methanation reaction is unlikely to occur in the absence of

a catalyst. Fast quenching rates may be required to avoid recombination of Mg and CO.

Adiabatic expansion through a Lavalle nozzle has been proposed for fast quenching [54].

Figure 7 shows the Mg yield of reactions reactions 3 and 4 as a function of temperature. The

parameter is the total pressure: 0.1, 1, and 10 bars. When higher, industrially preferred pressures are

used, the thermodynamic equilibrium is displaced in such a way so as to relieve the pressure

according to le Chatelier’s principle, i.e. to the left of Eqs. 3 and 4. For example, at 0.1, 1, and 10

bars, reaction 3 goes to completion at 1545, 1765 and 2070ºC, respectively, and reaction 4 goes to

completion at 1500, 1700 and 2000ºC, respectively. As far as the steam-hydrolysis of Mg is

concerned, this reaction is thermodynamically favorable at any temperature, but, as shown in a

thermogravimetric analysis [52], kinetic considerations indicate the need of effecting the process at

above 350ºC. The production of undesirable byproducts in equilibrium is not forecasted.

1500 1600 1700 1800 1900 20000.0

0.2

0.4

0.6

0.8

1.0

1.2a)

CO

Mg

C

MgO

Equil

ibri

um

com

posi

tion (

mol)

Temperature (ºC)

Deliverable Report

D 4.1 GA No. 731287 14 Co-funded by the Horizon 2020 Framework

Programme of the European Union

200 400 600 800 1000 1200 1400 1600 1800 20000.0

0.5

1.0

1.5

2.0

2.5

CO

Mg

CH4

MgO

C

H2

b)

Equil

ibri

um

com

posi

tion (

mol)

Temperature (ºC)

Figure 6: Equilibrium compositions of the system (a) MgO + C, and (b) MgO + CH4, as a function

of temperature at 1 bar. Extracted from Ref. [52].

Figure 7: Mg yield for reactions 3 (“1a”) and 4 (“1b”) as a function of temperature and total pressure.

Extracted from Ref. [52].

Deliverable Report

D 4.1 GA No. 731287 15 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2.3. Copper

This section has been prepared by the Middle East Technical University (METU), Turkey.

Copper is produced from copper sulfide concentrates by heating them to 1200-1250 °C in a

furnace and then oxidizing them with air or oxygen-enriched air. Oxidized particles melt into two

phases, a copper iron sulfide solution which is called matte and an iron silicate slag.

There are two types of copper smelting processes: reverberatory smelting and flash smelting. In

reverberatory smelting, the concentrate is reacted in a turbulent bath of matte, slag and gas. In

flash smelting, the exothermic reactions occur between air, or oxygen enriched air, and dispersed

concentrate particles similar to smelting in Alumina production.

Samsun Copper Plant in Turkey is targeted for obtaining process mass and energy balances. The

plant uses flash smelting and the plant data was collected and presented in a previous METU thesis

by Aidar Argyn [57]. At the site, there is an Outokumpu type flash smelter in operation since 1973.

Outokumpu flash smelting consists of injecting dry concentrate and preheated air or oxygen-

enriched air into a furnace. Feed enters through the top of a reaction shaft where it combusts with

the oxygen blast. Off-gases leave via a vertical uptake shaft at the other end of the furnace. The

concentrate particles should pass through dryers and preheaters before entering the furnace.

In the Samsun plant, most of the gases produced during smelting are converted to H2SO4. Also,

there is a flotation plant for processing and enrichment of the slag coming out from the flash

furnace.

In the thesis [57], starting with the flow sheet of the Samsun plant (shown in Figure 8) and flash

furnace parameters (Table 7), compositions of input and output flows for the flash furnace were

given as shown in Table 8. The material flow compositions among the components of the plant are

presented in Figure 9. The final results of the energy audit presented in the thesis are mass and

energy balances of flash smelting process for 1 ton of concentrate input as shown in Table 9 and

Table 10, respectively.

We plan to generate our own model of the plant streams using the information provided in [57] and

investigate the simulation results for a possible high temperature solar heat input. Solar hybridization

of a flash smelting plant may be kept limited to drying and preheating of the concentrated before

the flash furnace or may even involve a solar reactor that can replace part of the furnace. Time

dependent simulations we plan to perform will help us to evaluate various solar hybridization

options we have already identified.

Deliverable Report

D 4.1 GA No. 731287 16 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 8 Flowsheet of the Samsun plant (from [57])

Table 7 Parameters of flash furnace of the Samsun plant (data from [57])

Reaction air temperature 350-400°C

Amount of reaction air for 1 ton of feeding 650-800 m3/h

Feeding 35-40 tons/h

Concentration of SO2 in off-gases of flash

furnace

8-12%

Density of matte 4.7 g/cm3

Density of slag 3.7 g/cm3

Matte grade 45-50%

Copper in furnace slag 0.8-2%

Deliverable Report

D 4.1 GA No. 731287 17 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Table 8 Compositions of input and output flows, %, Symbols in parenthesis overrides the column symbol (from

[57])

Materials Cu Pb Zn Fe S SiO2 Fe3O4 others

Input:

Cu concentrate 19.62 0.26 2.26 29.95 33.31 3.22 3.62 7.76

Flux 87 13

Input dust 20 5 10 25 7 3 - 30

Lignite 53 (C) 6 (H) 23.4

(O)

- 1 - - -

Natural gas (mol %) 85

(CH4)

8

(C2H6)

3

(C3H8)

1

(C4H10)

1

(CO2)

0.5

(O2)

1

(N2)

0.5

(S)

Blast (vol. %) - - 21 (O) 79 (N)

Output:

Matte 45 0.6 1.8 22 23 - 7.4 0.2

Furnace slag 1.5 0.1 3 41 2 30 6.5 15.9

Dust (collected

from Off-gas)

20 5 10 25 7 3 - 30

Off-gas Remainder of calculation

Figure 9 Material flow compositions of Samsun plant (from [57])

Deliverable Report

D 4.1 GA No. 731287 18 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Table 9 Heat balance of flash smelting for 1 ton of concentrate (data from [57])

Inlet heat kJ

Heats evolved in exothermic reactions 2356747

Heat of process air (preheated air T=400°C) 511343

Heat of Lignite 302783

Heat of Natural Gas 833821

Total: 4004694

Outlet heat kJ

Heat of matte (T=1162°C) 315566

Heat of furnace slag (T=1190°C) 439370

Heat of off-gases (T=1263°C) 2437301

Heat of dust (T=1263°C) 82073

Heat loss of reaction shaft 730384

Total: 4004694

Table 10 Material balance of flash smelting for 1 ton of copper sulfide concentrate (from [57])

Materials kg m3

Inlet

Copper concentrate 1000

Flux 112.47

Lignite 28.52

Dust 72.83

Natural Gas 36.2

Air:

for process

for lignite

for natural gas

1012

151.38

378.15

Total: 1213.82 1541.53

Outlet

Matte 423

Furnace slag 510.73

Output dust 83.75

Gas:

from process

from lignite combustion

from natural gas combustion

from leakage air

976.15

164.03

417.22

300

1017.48 1857.4

2.4. Tin

Modeling of tin smelting processes

This section has been prepared by the Middle East Technical University (METU), Turkey.

The principal mineral for tin is cassiterite, SnO2, containing 87.8% of tin [58]. Extracting tin from tin ore

is done differently according to the source of the ore deposit and its quality. Tin is extracted by

reduction with carbon, to produce tin and carbon dioxide. Starting with the mining process,

followed by the smelting and refining processes, tin is extracted.

Deliverable Report

D 4.1 GA No. 731287 19 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Following the mining process, smelting is started. The low-grade concentrate is first roasted in a

furnace. The furnace could be any of 3 types: the reverberatory furnace which is the most

commonly used, the blast furnace, and the electric furnace [58]. Roasting is done to take out the

sulphur, converting them to oxide, where temperatures between 550 to 660 °C are reached [59].

Following that, leaching with water or acid solutions is done to remove impurities made soluble by

roasting. Next, smelting starts. Heating the cassiterite with carbon from anthracite coal and coke

[58] with limestone is done in the furnace to temperatures between 1,300–1,400 °C for 15 hours [59].

This produces molten pool tin, called as crude tin. Above that, slag forms, which contains impurities

and tin. The slag is reprocessed to extract the residual tin over and over again until hardhead forms

(tin recovered from iron-based compounds) and the remaining slag is discarded [59].

The crude tin from the smelting process is refined. The refining process can be of two types: fire

refining and electrolytic refining [58]. Fire refining can be done using boiling, liquation or vacuum

distillation. In boiling, the same furnace used for smelting is used to heat the impure tin in vessels

agitated with compressed air. This oxidizes the impurities that is discarded. In liquation, the impure

tin and the dross from the smelting process are treated. It is fortunate that tin has a low melting

point; hence the material is heated in a reverberatory furnace that melts the tin which is then

collected in a vessel. The unmelted residue is obtained and is treated again. Vacuum distillation is

similar to liquation but is done in a graphite vessel to temperatures between 1100 – 1300 °C [58].

In the reverberatory furnace, hot gases pass over the surface of the metal in a brick box. The brick

box is equipped with end wall burners to provide the heat for smelting [59]. Canadian practice uses

electric furnaces [59], which are the same as the reverberatory except that the heating is done by

a row of graphite electrodes projecting through the roof into the slag and the tin layer. They are

used to achieve very high temperatures, especially in secondary smelting processes. The heating is

then done by means of electrical resistance. Electric currents of 6–20 kA at 50–150 V are used [60].

Depending on the raw material concentration, energy consumption is 750–1400 kW h/t

concentrate (1300–1860 kW h/t tin). The depth of the electrodes can be controlled for process

control. Blast furnaces are sometimes used but are being replaced by reverberatory furnaces [59].

Smelting processes of temperatures up to 1500 °C in electrical furnaces consume 500 – 1000 kW h/t

and 1 – 10kg of electrode per ton of slag [60].

Modelling the various smelting processes with different combinations of fossil fuel and electricity use

and trying to identify solar energy use in the form of heat or PV electricity will be the next step.

In addition to these modelling efforts at METU, there is an ongoing project at AEE-INTEC which

constitutes an in-kind contribution to this INSHIP task. The following sub sections summarize the

project and its outcomes.

Energy audit of a pewter manufacturer

The following section has been prepared by the AEE – Institute for Sustainable Technologies (AEE-

INTEC), Austria.

2.4.1. The company

The data is based on an audit performed in a Malaysian company, operating worldwide (Kuala

Lumpur, London, Hong Kong, Singapore, Beijing, Melbourne and Sydney). The company is

manufacturing giftware and accessories in pewter, wood, gold & precious stones, silver, stannium,

bonded porcelain & glass.

Deliverable Report

D 4.1 GA No. 731287 20 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2.4.2. Process description

The pewter manufacturing process includes: selection of raw materials, casting, handwork, lathe,

cut, burnishing, soldering, hand polishing and packaging. The raw materials (tin, copper and

antimony) are cut up and mixed in the right proportion. The percentage of tin varies from 93-97%

depending on the application. The mixture is then heated to a temperature of about 350°C in a

melting pot above a furnace, and a molten liquid is formed, which using a casting ladle, is poured

into a mold with extreme care. This process is called casting. Then the product is further finished by

hammering, turning in a lathe, polishing, and sometimes engraving. Some product items, such as

snuffboxes, are built from separate pewter pieces and then soldered together.

Since pewter is an expensive material, the waste from the handwork and the lathe cut and other

processes such as polishing, soldering and hand finishing are collected and smelted to be used as

recycled pewter to create new product items. Figure 10 shows the generalized standard

production processes flow sheet of pewter products from casting to finishing.

Figure 10: Flow sheet of pewter manufacturing process

The thermal energy consuming processes are the melting and the die casting. However, most of

the die casting is supplied electrically. Figure 11 shows the flow scheme for the melting process.

Deliverable Report

D 4.1 GA No. 731287 21 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 11: Diesel supply, heat utilization and temperature levels in the melting process

a. Energy sources used

The energy sources typically used are electricity and diesel. Electricity is used to provide lighting

and air conditioning services for the show room, office and production areas. The production

processes use electricity to run the motors, machines and to provide heating for casting process.

Diesel fuel is fired to generate heat supplied to the melting process and parts of casting process.

Electricity is supplied by TNB. Diesel is purchased from few external suppliers.

b. Process evaluation

Melting of swarf (waste material after lathe cut) as shown in Figure 12a was selected for study due

to the low thermal conductivity from its shape and form. Loss of heat to the ambient was minimal

due to good rock wool insulation. Potential loss is from the sensible heat loss in the combustion

product. 76 kg of (swarf + moisture) is heated in the cast iron cauldron. The heat is supplied using

diesel fired furnace as shown in Figure 12b.

The heat generated from the burning of diesel is used to produce a flame. The furnace receives

heat from the flame and heat by convection and conduction from the combustion gas which swirls

around the pot before leaving the chamber and going to the stack. The burner specification states

that the efficiency is in the range of 75-85%. The burner nozzle specification is 2.5 gal/hr and the

burner output specification is 10.7 kg/hr.

Deliverable Report

D 4.1 GA No. 731287 22 Co-funded by the Horizon 2020 Framework

Programme of the European Union

(a) (b)

Figure 12: (a) Photo of swarf material in the cauldron and (b) drawing of the furnace and cauldron

Several physical processes run simultaneously during the melting step: the melting of the material,

the evaporation of the coolant and the heat generation by combustion process. The melting

process requires heat to raise the temperature of the material to the melting point. From the

literature review the pewter is fully melted at a temperature of 290-295⁰C. In this case, because the

swarf material does contain water therefore additional energy is needed as shown in Figure 13 and

in the actual energy demand in Table 11.

Figure 13: Melting process input and output diagram

Table 11 Heat demand of the melting process

Stream Phase Duration

[h]

Mass

[Kg]

T

initial

[⁰C]

T final

[⁰C]

ΔT

[⁰C

]

Cp

[kJ/kgK]

Latent

Heat

[KJ/KgK]

Q

[kJ]

H

change

[kJ/h]

Pewter Solid/

liquid

0.98 67.45 39 421 382 0.211 64.54 9,790 9,990

Water

(coolant)

Liquid 0.98 8.55 39 100 61 4.186 2260 21,506 21,945

Deliverable Report

D 4.1 GA No. 731287 23 Co-funded by the Horizon 2020 Framework

Programme of the European Union

The supply system (diesel) was evaluated based on burnt diesel and the air used. By this the energy

input was calculated.

Sankey diagram

The energy flow for the melting process is presented in the Sankey diagram as shown in Figure 14.

Figure 14: The Sankey diagram for swarf melting

The thermal energy audit and relevant energy data were collected and analyzed. The analysis

revealed that only 2.4% of heat supplied is used for melting, 2.5% for heating the pot, 5.3% for

removing the water and 89.2 % is loss as shown in in Sankey diagram. This is mainly based on the

outlet of flue gas and hot air as well as radiation and convection losses. To a certain extent this will

also include the heat stored in the hot product, but as this reflects only 2.5% of the total energy

demand, this is seen limited.

For system optimization a Pinch analysis was performed, but is not part of the project

documentation within INSHIP.

2.4.3. Identified solar thermal integration concepts

Solar and electricity for melting

The heating process for melting will be carried out through the introduction of Scheffler solar

collector/ electric heaters and the press. Scheffler collector would reduce power requirement of

electrical heaters to melt pewter swarf. Most of the trapped coolant would be removed and safely

disposed through the press without burning process.

Description of the measures

Overall system improvements: The proposed improvement to pewter swarf melting is by combining

electrical heaters and solar thermal heating. Electrical power requirement is reduced by using a

smaller pot and melting in smaller quantity. Solar thermal heating through Scheffler collector is the

second heat source. The Scheffler reflector focuses heat directly on the melting pot. The reflector

surface area of the test rig is 8 m2 and the average aperture area is 6.4 m2. At peak period with

estimated 33% efficiency, the reflector produces 2.3 kW of heating power. The combined heating

sources are to melt minimum of 100 kg of pewter swarf at 350 ⁰C.

Deliverable Report

D 4.1 GA No. 731287 24 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Efficiency of the Scheffler collector system was estimated by setting up 36 mirrors as reflector and

fabricating an aluminum mini cauldron as collector. System efficiency was initially recorded at 28%

with convection heat loss from the exposed or non-insulated aluminum surface of the collector.

System efficiency improved to 33% by fixing a glass plate covering the focal area at the collector.

A prototype Scheffler collector with 2.7 m2 reflector area is being fabricated. The prototype enables

a better study on reducing convection heat loss from focal point area. The proposed solutions are

to install heat resistant glass partition and sun ray sensitive automated trap door at the focal area

of the collector.

Solar heating using Ritter XL

Another option is to preheat the materials and to evaporate the water coolant. A solar thermal

system is designed to achieve a process temperature of 200°C using Ritter XL solar plasma

collectors. The auxiliary heating will be electric. For this system, a SHIP Tool (developed by AEE

INTEC) was used for the evaluation.

The heat required for the process includes the heat required to raise the water coolant

temperature from ambient to 100⁰C and heat required for evaporation and the heat required to

raise the pewter temperature to melting point and provide the heat of fusion. Table 12 shows the

heat requirements and Figure 15 shows the load profile of the process.

Table 12 Daily load and temperature levels of the process

Deliverable Report

D 4.1 GA No. 731287 25 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 15: Annual load profile of the process

Deliverable Report

D 4.1 GA No. 731287 26 Co-funded by the Horizon 2020 Framework

Programme of the European Union

3. Lime processing from limestone

The decomposition of limestone (mainly CaCO3) is the main endothermic step in the production of

lime (CaO) and cement. Lime and cement manufacturing are high-temperature energy intensive

processes. The minimum amount of energy in the form of process heat that is required to drive the

calcination reaction, CaCO3 → CaO + CO2 at the decomposition temperature near 1173 K is about

3029 kJ per kg of CaO, while the heat of dissociation of calcite relative to 298 K is 3184 kJ per kg of

CaO. Total heat consumption of modern lime kilns ranges from 3600 kJ/kg of CaO for vertical

double shaft kilns to 7500 kJ/kg of CaO for non-preheated long rotary kilns.

Because the process heat is traditionally supplied by the combustion of fossil fuels (e.g., oil, coal, or

natural gas) a lime or cement plant releases CO2 derived both from the calcination reaction and

from the combustion process that supplies the energy for the reaction. The use of concentrated

solar energy as the source of process heat can mitigate these emissions [42-46]. The cleanliness of

the solar process leads to very pure lime. It seems conceivable that a pure product may be

advantageous for special niche markets in the chemical and pharmaceutical industries. The

technical feasibility of the solar calcination process has already been demonstrated for 10 kW solar

reactor prototypes [61].

The following sections provide an overview of the work carried out by INSHIP members under Task

4.2 of WP4 (Solar lime production for the cement industry). High-temperature process heat can be

obtained in a solar receiver heating a heat transfer medium with concentrated solar energy. The

medium can be a fluid (e.g. air, steam) or can be a particle flow, as highlighted in the following

sections.

3.1. Solar hybridization for cement production

This section has been prepared by the Middle East Technical University (METU), Turkey.

Within the scope of this research activity, we aim to obtain mass and energy balances for cement

production processes which will be useful in identification of solar hybridization paths.

The most common cement product in the market, Ordinary Portland Cement is produced by

heating limestone with small amounts of other materials such as clay to 1450°C to produce “clinker”

and then by grinding clinker together with a small amount of gypsum and even smaller amounts of

other minerals.

Clinker production process has a large carbon footprint. According to the report prepared by

International Energy Agency (IEA) and the Cement Sustainability Initiative (CSI) of the World

Business Council for Sustainable Development (WBCSD) [62], the cement industry is responsible for

7% of the global carbon dioxide (CO2) emissions. In 2014, this corresponds to 2.2 gigatons of CO2 in

a year which was 27% of total direct industrial CO2 emissions. To be able to understand how we can

reduce CO2 emissions of cement industry, we need to take a look at the sub processes and their

mass and energy balances.

Cement production starts with extraction of limestone, clay and other minerals from quarries. This is

done by excavation machinery and the extracted material is carried by trucks to the usually

Deliverable Report

D 4.1 GA No. 731287 27 Co-funded by the Horizon 2020 Framework

Programme of the European Union

nearby cement plant. The machinery and trucks use fossil fuel and emit CO2 but hybridization with

high temperature solar heat is not possible at this initial step of production.

The next step is crushing the materials together to obtain the raw meal that will go into the cement

plant. The crushing machinery uses electricity and electricity can be supplied from a photovoltaic

(PV) field but we are not interested in hybridization with solar PV either. Because of our limited

scope to solar heat, we only consider the cement production steps after the raw meal enters the

cement plant.

In a modern dry cement plant, the raw meal is heated initially in the preheater then in the pre-

calciner up to 900°C at which the calcination reaction starts to occur. The rest of the calcination

reaction occurs at the early stages of the rotary kiln. The products of the calcination reaction are

lime (CaO) and CO2. In fact, 60-70% of CO2 emissions from cement production occurs as the

product of calcination. About 65% of the fuel use in a cement plant occurs for the calcination

process.

Next, the pre-calcined meal goes into the rotary kiln. The rotary kiln is a fire brick lined long and

heavy steel cylinder. The kiln rotates with a rotation speed in the range of 1-5 rotations per minute

and it is usually inclined to advance the meal slowly towards the hot flame that is at the bottom

lower part of the kiln. The meal gets hotter and hotter as it moves towards the end of the kiln. The

kiln reaches to temperatures as high as 1450°C. The rest of limestone that is not calcined in pre-

calciner gets into reaction and calcination is completed here and all of the raw materials are

combined together in a partially molten form. The product at the exit of the kiln is called clinker.

At the next step, hot clinker coming from the kiln is cooled by fresh air blown onto it at a grate

cooler. Here, clinker temperature reduces to 100°C very rapidly. The resulting clinker goes to

storage.

To produce cement, the stored clinker is mixed with gypsum and other minerals and ground into a

powder. The resulting grey powder is called Ordinary Portland Cement. Clinker can also be

blended with ash or some other materials coming from various industries to produce blended

cement mixtures. The mixing, blending and grinding processes require electricity.

Only the part of cement production process starting with crushed raw meal and ending with

cooled clinker involves heat processes. We call that, clinker production and target it for solar

hybridization.

In modern cement plants, heat processes are optimized for efficiency. Fresh air required for

combustion is preheated by the cooling clinker and exhaust of the kiln goes to the pre-calciner

and after that to the pre-heater. Even the exhaust of the pre-heater can be utilized for site heating

or for district heating. If the exhaust is hot enough, it can be used in Heat Recovery Steam

Generators to produce steam for producing electricity with a steam turbine.

There are still some old cement plants operating without a pre-heater or even without a pre-

calciner but with a long kiln. Thus, their heat recovery options are limited, leading to low

efficiencies. Some of the old plants operate on a wet process instead of a dry one, requiring

evaporation of moisture leading to low thermal efficiencies. We consider a modern cement plant

with pre-heater steps and with a pre-calciner and determine the mass and energy balances for

such a plant starting with dry materials. Hence, our aim is to identify the paths for improving modern

cement plants by solar hybridization.

Deliverable Report

D 4.1 GA No. 731287 28 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Our initial attempt for obtaining mass and energy balances was to try to obtain operational data

from a modernized cement plant in Ankara. That attempt turned out to be unsuccessful because

of two reasons: i) the company operating the plant was not responsive to provide detailed data

most probably due to being protective of their trade information; ii) placement of a student

working on INSHIP at the site for a long period of time to collect data was not foreseen and

budgeted in INSHIP. Considering the operational data even in International Energy Agency (IEA)

reports are always 2 years old due to the same trade secret concerns, it became obvious that

either the cement company or some of its employees should be partners in INSHIP to be able to do

our own energy auditing. Thus, we turned our attention to earlier energy audits of cement plants in

the literature. We found two such audits for cement plants, one of them in Van (eastern Turkey) and

the other in Gaziantep (southeastern Turkey). Especially, Gaziantep is a promising location for solar

hybridization due to its very high year-round Direct Normal Irradiation (DNI) values. We will present

the mass and energy balances of both of the cement plants here.

3.1.1. Van Cement Plant Mass and Energy Balances

Engin and Ari [63] performed mass and energy balances for a dry processing cement plant located

in the province of Van in eastern Turkey. The data have been collected over a long period of time

when the first author was a process engineer in the plant. The plant is a relatively modern cement

plant. As in the case of all energy auditing studies performed for cement plants in the literature,

authors first present the raw material and clinker compositions given in Table 13 as it was presented

in [63]. The mass flows in and out of the kiln system consisting of pre-heater, rotary kiln and grate

cooler are shown in Figure 16. The plant operates by burning coal. Mass inputs and outputs of the

kiln system including the coal and combustion air are shown in Figure 17. The complete energy

balance of the kiln system is also presented in [63]. We extracted only the results of the balance

and formed Table 14 based on that data.

Table 13 Raw material and clinker components and their percentages (from [63])

Component Raw material (%) Clinker (%)

SiO2 13.55 20.55

Al2O2 4.10 5.98

Fe2O3 2.60 4.08

CaO 40.74 64.58

MgO 2.07 2.81

SO3 0.56 1.30

K2O 0.30 0.50

Na2O 0.08 0.20

H2O 0.5 -

Organics 0.9 -

Ignition loss 34.6 -

Total 100 100

Deliverable Report

D 4.1 GA No. 731287 29 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 16: Control volume and mass flows of the kiln system together with the temperatures (as it was

presented in [63])

Figure 17: Mass inputs and outputs of the kiln system including the coal and combustion air [63]

Deliverable Report

D 4.1 GA No. 731287 30 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Table 14 Complete energy balance results (data from [63])

Description Result (kJ/kg-clinker)

Heat Inputs

Combustion of coal 3519 (95.47%)

Sensible heat by coal 7 (0.19%)

Heat by raw material 72 (1.95%)

Organics in the kiln feed 19 (0.52%)

Heat by cooling air 69 (1.87%)

Total heat input 3686 (100%)

Heat outputs

Formation of clinker 1795 (48.70%)

Kiln exhaust gas 706 (19.15%)

Moisture in raw material and coal 26 (0.71%)

Hot air from cooler 207 (5.61%)

Heat loss by dust 11 (0.30%)

Clinker discharge 86 (2.33%)

Radiation from kiln surface 386 (10.47%)

Convection from kiln surface 171 (4.64%)

Radiation from pre-heater surface 7 (0.19%)

Natural convection from pre-heater surface 5 (0.14%)

Radiation from cooler surface 4 (0.11%)

Natural convection from cooler surface 9 (0.24%)

Unaccounted losses 273 (7.41%)

Total heat output 3686 (100%)

3.1.2. Gaziantep Cement Plant Mass and Energy Balances

Atmaca and Yumrutas [64][65][66] performed energy auditing of a cement plant in Gaziantep

province of southeastern Turkey. They performed their first and second law analyses for the cases

before and after some improvements for reducing heat losses. The clinker composition of the plant

is given in Table 15 as it is shown in [64]. Mass inputs and outputs of the kiln system together with

their temperatures are given in Figure 18. Energy flows are presented as a Sankey diagram in Figure

19. The mass and energy balances are shown in Table 16.

Table 15 Clinker composition (as presented in [64])

Chemical name Chemical structure Chemical form Percentage (%)

Calcium ferrite 4CaO.Al2O3.Fe2O3 C4AF 10.4

Di-calcium silicate 2CaO.SiO2 C2S 13.2

Calcium aluminate 3CaO.Al2O3 C3A 9.1

Tri-calcium silicate 3CaO.SiO2 C3S 60.2

Potassium oxide K2O - 2.5

Sulfur trioxide SO3 - 2.1

Magnesium oxide MgO - 1.2

Sodium oxide Na2O - 1.3

Total 100

Deliverable Report

D 4.1 GA No. 731287 31 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 18 Mass flow diagram of the kiln system (from [64])

Figure 19 The energy Sankey diagram of the kiln (from [64])

Deliverable Report

D 4.1 GA No. 731287 32 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Table 16 Mass and energy balance of the unit (data from [64])

Input materials Mass flow rate (kg/h) Energy input rate (kW) Percentage (%)

Farine 105000 20089.47 18.1

Pulverized coal 7200 112.6 0.1

Primary air 9866 87.48 0.1

Secondary air 89511 23240.13 20.9

Electrical work - 4341.5 3.9

Combustion of coal - 63311.65 56.9

Total 211577 111182.83 100

Output materials Mass flow rate (kg/h) Energy output rate

(kW)

Percentage (%)

Formation of clinker - 36537.63 32.9

Clinker 65200 24293.94 21.8

Hot gas 133492 36277.74 32.6

Dust and ash 12885 1581.00 1.4

Heat transfer from kiln - 12542.51 11.3

Total 211577 111182.83 100

3.1.3. CSI European Plant Data and General Discussions

There are large variations in mass and energy balances of individual plants due to variations in

initial raw materials and heat sources. Unless the audited cement plant is the actually targeted

plant for solar hybridization, obtaining mass and energy balances for a specific plant cannot lead

to generalized solar hybridization solutions for the cement industry. Thus, to be able to represent an

average plant, we decided to rely on CSI database which is available through getting the

numbers right (GNR) program [67]. When we take a look at the 2016 data that is published very

recently in 2018, we realized that most of the data about cement production in Europe is

independently verified which makes the European data even more valuable to us. It might be

possible to obtain energy use estimates for an average European cement plant and to create

rough mass and energy balances based on the European data from GNR. That will be our strategy

to create a third European location for our system level simulations.

Essentially, in a dry clinker production process, the production rate of clinker is equal to initial raw

meal inlet rate. The clinker production process runs 24 hours a day all year long for being able to

recover heat from the flame exhaust of the kiln and heated air during the cooling of clinker in the

cooler. Obviously, if the processes need to start cold e.g. after a maintenance period, initial

heating of the thermal mass and cold raw meal will require extra energy. Since the clinker

production under normal circumstances runs continuously, in energy balances, only the steady

state operation of the plant is considered. We will also consider the steady state operation of the

plants in our system level simulations with and without solar hybridization. During the modeling of

the reference plant, while selecting model parameters and determining the methodology, we plan

to use Kabir et al. [68], Ustaoglu et al. [69], and Cao et al. [70]. For comparison of our simulation

results, we plan to use two review articles from the literature [71][72].

CSI has identified the clinker ratio reduction in cement mixtures as a way to reduce CO2 emissions.

Unfortunately, looking at the change of clinker ratio in the CSI data over the years, no significant

reduction trend can be observed. If the clinker is going to be produced with the existing processes

in the future, then we need to reduce fossil fuel use which will lead to direct reduction in CO2

Deliverable Report

D 4.1 GA No. 731287 33 Co-funded by the Horizon 2020 Framework

Programme of the European Union

emissions. High temperature solar heat is a good candidate for replacing some of fossil fuel burning.

Targeting the kiln and the pre-calciner for high temperature solar hybridization looks promising.

3.2. Particle based gravity driven solar receiver development for solar lime

This section has been prepared by the Middle East Technical University (METU), Turkey.

3.2.1. Modeling Solid Particles

Concentrating Solar Power utilizing solid particles as a heat transfer and storage medium has been

gaining attention among researchers [73][74][75], with potential benefits including higher

temperatures, higher efficiencies, and lower cost [76] [77] compared to state-of-the-art molten salt

based systems. The high working range of these systems, projected to be around 700 to 900 ⁰C,

make them well-suited for pre-heating and high temperature industrial applications. Unlike fluids,

which have a long history of research and countless experimental studies validating the

mathematical models, modeling a group of flowing solid particles surrounded by a fluid is much

more difficult to model, given the various ways the particles, fluid, and walls can interact in terms of

both heat transfer and momentum transfer. Therefore, in the pursuit of designing gravity-driven solar

receivers which will be useful for lime calcination, we are improving upon the state-of-the-art

modeling techniques for dense particulate flows. Furthermore, heat transfer experiments will be run

to validate the new modeling techniques. After the models have been built and experimentally

validated, solar receiver designs for lime calcination can be designed and evaluated. The methods

and results are given in more detail in the proceedings from the SolarPaces 2018 conference [78].

Unlike fluids, a flowing collection of particles does not follow the continuum assumptions which are

common in fluid mechanics. By coupling Computational Fluid Dynamics (CFD) with the Discrete

Element Method (DEM), a coupled CFD-DEM simulation can account for both frictional effects of

the particle phase as well as behavior of the interstitial gas [79][80]. With a coupled model, both

frictional effects of the particles as well as the hydrodynamic and thermal effects of the fluid are

simulated, which is not possible with a DEM or CFD model alone. A generic domain consisting of

particles and an interstitial fluid bounded by a heated, opaque wall consists of six heat transfer

modes. Four modes are already incorporated into the LIGGGHTS-CFDEM-OpenFoam, the leading

open-source coupled CFD-DEM software: particle-particle conduction, particle-wall conduction,

particle-fluid convection, wall-fluid convection. Two heat transfer modes, particle-particle radiation

and particle-wall radiation, are not included in this or other available software. However, since

radiation plays a large role in heat transfer at high temperatures, it is crucial to add the radiative

heat transfer modes before a CFD-DEM code can accurately model a solid particle solar receiver.

3.2.2. View Factor Correlation Concept

To calculate the radiative transfer between two bodies in a CFD-DEM simulation, the view factor is

the key unknown parameter. View factors can be calculated analytically or with Monte Carlo ray

tracing [81], but implementing these directly into a DEM code would add a prohibitive

computational burden. Instead, a computationally faster method is pursued here, with a similar

approach previously suggested by Malinouski [82]. First, a randomly packed, static particle bed is

generated using DEM. Second, a Monte Carlo simulation is run to find the view factor between

each particle and its neighbors. Third, each particle-particle view factor is plotted against the

center-to-center distance between the particles, and a polynomial correlation is found. Finally, this

view factor correlation is implemented into DEM. The benefit of this approach is that after the

Monte Carlo simulation has been run once to obtain the correlation for a certain set of conditions,

Deliverable Report

D 4.1 GA No. 731287 34 Co-funded by the Horizon 2020 Framework

Programme of the European Union

view factors in DEM can be quickly calculated based only on the distance between particles, and

radiative heat transfer is readily calculated. Given the random packing and shading of particles,

any isolated pair of particles may have a true view factor more or less than the correlation, but the

correlation should accurately predict radiative heat transfer across the particle bed as a whole. In

this analysis, the particles are assumed to be densely packed with a volume fraction of ~0.64, all

particles and walls are assumed to be black bodies, and all particles are assumed to have the

same diameter.

3.2.3. View Factor Correlation Results

The Monte Carlo method was used to find the view factor from one particle to its neighbors. In this

simulation, many “photons” or rays were emitted in random directions, from a random location on

the particle surface, and the trajectory of each photon was followed until it hit a neighboring

particle. After emitting many photons, the view factors were found by dividing the number of

photons hitting a neighbor by the total number sent. Results are shown in Figure 20, where the view

factor is plotted against distance for both the particle-particle and particle-wall cases. Though

some scatter is inevitable due to the random nature of the particle positions, there is a clear trend,

and the two-part correlations are presented below. The correlations are valid for particles of any

size because distance is expressed in units of particle radii.

Figure 20: Particle-particle view factors (left) and particle-wall view factors (right) found with Monte-Carlo

simulation

Particle-particle view factor:

Particle-wall view factor:

R ≤ 4:

VFP-P = a0 + a1R + a2R2 + a3R3 + a4R4 + a5R5 + a6R6

a0 = 7.406948, a1 = -13.706197, a2 = 10.628515, a3 = -4.3657168

a4 = 0.99654538, a5 = -0.11966892, a6 = 0.0059067682

4 < R < 6:

VFP-P = a0 + a1R + a2R2 + a3R3

a0 =0.012264422, a1 = -0.0045066324, a2 =0.00053308638, a3 = -

0.000020380795

Deliverable Report

D 4.1 GA No. 731287 35 Co-funded by the Horizon 2020 Framework

Programme of the European Union

R ≤ 2.75:

VFP-P = a0 + a1R + a2R2 + a3R3

a0 =0.51909859, a1 = -0.18425542, a2 = -0.045247181, a3 = 0.016712257

2.75 < R < 6:

VFP-P = a0 + a1R + a2R2 + a3R3

a0 =0.13749899, a1 = -0.076620909, a2 =0.014266074, a3 = -0.00088424010

3.2.4. Correlation Validation

For validation of the particle-particle view factor correlation, the equation was put into LIGGGHTS,

and simulations were run to find the effective thermal conductivity. Results were compared with

those predicted by the Zehner, Bauer, and Schlünder (ZBS) model, and a close agreement was

found as shown in Figure 21 (a). For validation of the particle-wall view factor correlation, the ZBS

model is not applicable, results were compared with a fully Monte-Carlo simulation, and a close

match can be seen in Figure 21 (b).

Figure 21: Verification of correlations for particle-particle (a) and particle-wall view factors (b).

3.2.5. Full Simulation

To demonstrate the new code and show the importance of including radiation, a coupled

simulation was run with the original code and with the addition of the new radiation models. Results

from the two simulations are shown in Figure 22 after 0.39 seconds of modeled time. Total heat

added to the particles increased 220% when radiation was included, clearly demonstrating the

need to include radiation in the heat transfer modeling of granular flows at high temperatures.

(a) (b)

Figure 22 Results showing the original model (a) and with radiation heat transfer included (b)

Deliverable Report

D 4.1 GA No. 731287 36 Co-funded by the Horizon 2020 Framework

Programme of the European Union

3.2.6. Conclusions

A method for including radiation heat transfer within a granular flow has been presented and initial

validation performed. The importance of including radiation has been demonstrated with the first

simulation where radiation heat transfer was implemented into a coupled CFD-DEM code. With the

aim of designing a solar receiver capable of efficiently transferring heat to the particles flowing

inside for solar lime production, the CFD-DEM modeling capabilities currently being built will help to

design and optimize such a receiver, by comparing the effects of fin spacing, other geometric

parameters, and material properties. Additionally, particles flowing through the receiver may be an

intermediate material such as sintered bauxite, which is widely available from the hydraulic

fracturing industry, or silica sand, available worldwide. Alternatively, it may be possible to heat the

limestone particles directly in the solar receiver. Given the various options in terms of particle

choice and receiver design, CFD-DEM will be used to compare and different options and evaluate

the potential designs in terms of working temperature range and thermal efficiency.

3.3. Thermal performance and transient operation of a solar receiver

This section has been prepared by the Centro de Investigaciones Energéticas, Medioambientales y

Tecnológicas (CIEMAT), Spain

The manufacture of lime and cement are energy-intensive processes at high temperatures. The

energy in the form of process heat required to carry out the calcination reaction is at

approximately 1173 K.

23 COCaOCaCO

Because process heat is traditionally supplied by the combustion of a fossil fuel (e.g. oil, coal or

natural gas), CO2 is released in a lime or cement plant. Several strategies have been proposed to

mitigate these emissions. Examples include CO2 capture and sequestration, fuel switching, the use

of waste as a fuel source, and the use of concentrated solar energy as a process heat source.

Several approaches to solar reactors to integrate this chemical process have been studied in the

scientific literature. The main objective is to provide the best heat transfer between concentrated

solar energy and the reaction.

The use of chemical reactors with cavity geometries allows solar energy to be captured and

facilitates the reduction of thermal emission losses. These receptors usually showed a low response

to changes in operating conditions, which avoids abrupt temperature changes in the presence of

small transients, which is particularly useful when solar heat is applied in a chemical plant. In this

regard, a thermodynamic characterization of a solar chemical reactor under real conditions

emerges as a relevant matter to be considered.

The research activity of CIEMAT under Task 4.2 of INSHIP focuses on investigation of thermal

performance of a solar receiver on non-steady state conditions. A 100 kW receiver prototype has

been thermally tested at the CRS tower at Plataforma Solar d’Almeria, Spain in real conditions.

Reactor dimensions are the following: cavity reactor with semi-cylindrical shape and a square

aperture of 46 cm x 46 cm.

Mass and Energy Balances

The chemical process defined in the following equation:

Deliverable Report

D 4.1 GA No. 731287 37 Co-funded by the Horizon 2020 Framework

Programme of the European Union

23 COCaOCaCO ΔH298=178 kJ/mol (Endothermic)

produces both volatile carbon dioxide at a rate mCO2 out and a mixture of CaO and CaCO3, which

contains calcium oxide and some un-calcined calcium carbonate. A simplified mass balance is

given in the following equation:

323 CaCOCaOCOCaCO mmm

A complete energy balance scheme for the calcination is provided by Meier et al. [83]. The heat

balance model accounts for non-steady-state conditions by modifying the solar power Qsolar

entering the reactor. The effective power input includes the enthalpy change Qstorage of the reactor

wall materials caused by energy stored and released:

othercondirreradheatdissstoragesolarinput QQQQQQQQ

The authors stated that, for example, 13.6% of the effective power input is conducted through the

walls of the reactor(Qcond), 20.5% is lost by reradiation (Qrerad), 20.2% (Qheat) is carried away in the hot

products, and 14.9% (Qother) of the effective power input leaves the reactor as “other heat losses”

mainly convection heat losses and un-recovered calcined powder. The authors conclude that, the

reactor design changes can reduce any of these terms and it will lead to better reactor efficiency.

A brief heat balance is presented for calculating the efficiency of CIEMAT’s cavity reactor under

Task 4.2.

heatproductlossesCaCOCaCOinput QQHmQ ,33

where Q product heat represents the sensible heat of the flowing gas.

othercondreradlosses QQQQ

Q losses account for thermal energy losses by radiation (Qrad), re-radiation (Qrerad) and convection

(Qother).

The efficiency η is defined as the enthalpy change of the calcination reaction at ambient

temperature, divided by the power input.

input

heatproductCaCOCaCO

Q

QHm ,33

A preliminary evaluation of our cavity receiver has been performed during some experimental

campaigns and an empirical value of Q losses has been obtained from experimental data. This value

could represent a measure of how quickly or slowly the cavity reactor gains or loses heat to its

environment and it is associated with the geometry and working conditions of the receiver.

The Qlosses value was estimated to be about 114 kJ/K. It would represent the power needed for this

cavity receiver for steady state operation at 1100 ºC, that is, the temperature to carry out the

calcination reaction.

Deliverable Report

D 4.1 GA No. 731287 38 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 23: Colored curves correspond to thermocouples placed at different positions across the reactor. Black

points are the incident power in the aperture. Black lines are the number of heliostats focused on the receiver.

By a simplified energy balance across the cavity reactor, considering the Q input the value of power

obtained by the flux measurement system (black dots in Figure 23)

lossesnetinput QQQ

and the heat flows leaving the reactor, our design is able to achieve a thermal efficiency value of

ηreceiver = 0.35 to 0.40, which is a characteristic value for a cavity receiver.

Deliverable Report

D 4.1 GA No. 731287 39 Co-funded by the Horizon 2020 Framework

Programme of the European Union

4. Solar fuels production from H2O and CO2 via redox cycles

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [84].

Storage of the vast, yet intermittent, diffuse and unevenly distributed solar energy resource is

essential for the transition away from fossil fuels [85]. Liquid hydrocarbon fuels offer exceptionally

high energy densities and are convenient for the transportation sectors without changes in the

current global infrastructure. However, their production from H2O and CO2 using solar energy input

has remained a grand challenge.

Concentrated solar energy provides a virtually unlimited source of clean, non-polluting, high-

temperature heat. Solar flux concentration ratios exceeding 2000 suns (1 sun = 1 kW/m2) are

attainable with large-scale solar tower and dish systems, and higher concentrations are possible

with the incorporation of secondary optics in tandem with the primary concentrating systems, e.g.

compound parabolic concentrators. Dish systems use sun-tracking paraboloidal mirrors to focus

sunlight onto a solar receiver positioned at a focus. These solar concentrating modular systems

have been proven to be technically feasible for power generation in traditional Stirling cycles. Solar

thermochemical applications, although not as far developed as solar thermal electricity

generation, can employ the same solar concentrating infrastructure with the solar reactor

positioned at the focus of the solar dish in decentralized (approx. 100 kW each) modular systems

[84].

Solar thermochemical approaches to splitting CO2 and H2O inherently operate at high

temperatures and utilize the entire solar spectrum, and as such provide an attractive path to solar

fuels production with high energy conversion efficiencies in the absence of precious metal catalysts

[84]. In contrast to direct thermolysis of CO2 and H2O, two-step thermochemical cycles using metal

oxide redox reactions further bypass the CO/O2 or H2/O2 separation problem. Figure 24 shows

schematically the proposed two-step H2O/CO2 splitting thermochemical cycle based on metal

oxide redox reactions, encompassing:

1) an endothermic reduction of the metal oxide using concentrated solar process heat; and

2) an exothermic reaction of the reduced metal oxide with H2O/CO2 to produce H2/CO.

The net reactions are H2O= H2+½O2 and/or CO2 = CO+½O2. Since O2 and H2/CO are released in

separate steps, the need for high-temperature gas separation is thereby eliminated. The syngas

mixture H2/CO can be further processed to liquid hydrocarbon fuels (via Fischer-Tropsch and other

catalytic processes), such as diesel, kerosene, methanol, and gasoline using existing conventional

technologies.

Deliverable Report

D 4.1 GA No. 731287 40 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 24: Schematic of a two-step solar thermochemical cycle for H2O/CO2 splitting based on metal oxide

redox reactions. MOox denotes a metal oxide, and MOred the corresponding reduced metal or lower-valence

metal oxide. In the first, endothermic, solar step, MOox is thermally dissociated into MOred and oxygen.

Concentrated solar radiation is the energy source for the required high-temperature process heat. In the

second step, MOred reacts with H2O/CO2 to produce H2/CO (syngas). The resulting MOox is then recycled back

to the first step, while syngas is further processed to liquid hydrocarbon fuels [84].

Amongst candidate redox materials, ferrite-based oxides exhibit relatively slow reaction rates,

whereas ZnO, SnO2, and analogous volatile oxides that sublime during decomposition require rapid

quenching of gaseous products to avoid recombination [86–92]. Cerium oxide (ceria) has emerged

as a highly attractive redox active material choice because it displays rapid fuel production

kinetics and high selectivity [93–98]. The non-stoichiometric ceria fluorite-type structure and phase

are maintained as oxygen vacancies are created, making cyclability possible. The two-step

H2O/CO2 splitting solar thermochemical cycle based on oxygen-deficient ceria is represented by:

High-T reduction:

H x

2-δ 2-δ-x 22CeO CeO O

( 5 )

Low-T oxidation with H2O: 2-δ-x 2 2-δ 2CeO xH O CeO xH

( 6 )

Low-T oxidation with CO2: 2-δ-x 2 2-δCeO xCO CeO xCO

( 7 )

Figure 25 shows a) the equilibrium properties for CeO2-δ reduction [99]; and b) the equilibrium mole

fractions of H2 and CO for various reduction temperatures of ceria reacting with a mixture of H2O

and CO2. These graphs indicate that the reduction is thermodynamically favorable at above 1400

°C and 1 bar while the oxidation is favorable at below 1200 °C.

Deliverable Report

D 4.1 GA No. 731287 41 Co-funded by the Horizon 2020 Framework

Programme of the European Union

(a) (b)

Figure 25: (a) Equilibrium O2 pressures for CeO2-δ as a function of nonstoichiometry at 1400-1750 °C (b)

Equilibrium H2 and CO mole fractions for various reduction temperatures, as a function of oxidation

temperature H2O/CO2 molar ratio = 2 [99]

The first realization of a solar-driven ceria-based cycle focused on CeO2/Ce2O3 redox reactions [93].

The thermodynamics and kinetics of the cycles were studied for CeO2-δ [95]. Exploratory studies

have examined the effect of doping ceria with samarium [94], chromium [96], manganese, iron,

nickel, copper [97], and zirconium [98] to improve performance. These studies have largely been

confined to bench-top demonstrations of components or individual steps of the cycle with no

meaningful experimental evaluation of the energy conversion efficiency and limited assessment of

cyclability.

The following sections provide an overview of the work carried out by INSHIP members under Task

4.3 of WP4 (Solar fuel production for the transportation sector).

4.1. Solar receiver–reactor for thermochemical redox cycling using ceria

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [100]. This work has been carried out by ETH under the H2020 project SUN-

to-LIQUID (year 2016 to 2019) and is being reported as an in-kind contribution in the INSHIP project.

This work [100] reports on the development of a transient heat transfer model of a solar receiver–

reactor designed for thermochemical redox cycling by temperature and pressure swing of pure

cerium dioxide in the form of a reticulated porous ceramic (RPC).

To analyze the performance of the solar reactor and to gain insight into improved design and

operational conditions, a transient heat transfer model of the solar reactor for a solar radiative input

power of 50 kW during the reduction step was developed and implemented in ANSYS CFX. The

numerical model couples the incoming concentrated solar radiation using Monte Carlo ray tracing,

incorporates the reduction chemistry by assuming thermodynamic equilibrium, and accounts for

internal radiation heat transfer inside the porous ceria by applying effective heat transfer

properties.

The model was experimentally validated using data acquired in a high-flux solar simulator (HFSS),

where temperature evolution and oxygen production results from model and experiment agreed

Deliverable Report

D 4.1 GA No. 731287 42 Co-funded by the Horizon 2020 Framework

Programme of the European Union

well. Figure 26 shows a schematic of the experimental setup used to test the receiver. The reactor

configuration and experimental setup are described in detail in Zoller et al. [100].

Figure 26: Schematic of the experimental setup. During the endothermic reduction step (Eq. (5)) Ar flow is used

to protect the quartz window and is pumped out together with the released O2, while the reactor is under

vacuum. During the exothermic oxidation step (Eq. (7)), the ceria is fully reoxidized with CO2, thereby

producing CO. Details of the reactor configuration and experimental setup are available in Ref. [100].

Energy Flows

The instantaneous energy balance for the reduction step is illustrated as a function of time in Figure

27. Note that heat recovery was not applied. Losses by reradiation from the hot cavity, the change

in sensible heat content of the RPC, the remaining reactor components (Al2O3–SiO2 insulation,

aluminum front, reactor shell, and insulating jacket), the energy consumed by the endothermic

reduction reaction, the conductive heat loss to the water-cooled reactor front, and other heat

losses are indicated. Other heat losses include reflection of incoming solar radiation inside the

reactor cavity and at the quartz window, absorption of incoming radiation at the window, and

convection and radiation at the outer reactor surfaces.

Initially, sensible heating of the RPC dominates energy consumption, consuming 87% of the solar

radiative power input (Psolar ), while on average it consumes 33%. By the end of the reduction step,

reradiation dominates heat loss, accounting for 31% of Psolar on average and 45% at the peak.

Reradiation losses could be lowered by decreasing the size of the aperture, provided that solar

radiation can be delivered with higher concentration. A selective coating with high transmissivity in

the visible region of the solar spectrum, but high reflectivity in the IR region of the radiation emitted

by the hot cavity, could be considered for the quartz window, provided that the coating can

withstand very high temperatures (>500 °C). Sensible heating of the bulk materials consumes 21%

of Psolar on average, but levels off early in the reduction cycle, with the Al2O3–SiO2 insulation being

the dominant consumer, while the aluminum front, reactor shell, and insulating jacket consume

1.2% or less each. Energy loss through sensible heating of the bulk materials could be lowered if

insulation materials with lower specific heat capacity were used. The energy fraction driving the

endothermic reduction reaction of ceria quickly increases with time, and on average, accounts for

5.6% of Psolar.

The conduction heat losses to the water-cooled reactor front are significant, with an average

consumption of 2.7% of Psolar. The losses by convection and radiation at the outer reactor surfaces,

as well as the energy lost by reflection of the incoming solar radiation inside the reactor cavity,

Deliverable Report

D 4.1 GA No. 731287 43 Co-funded by the Horizon 2020 Framework

Programme of the European Union

account for less than 0.3% of Psolar each. The remaining 7.1% of Psolar is lost by absorption and

reflection at the quartz window (transmissivity = 0.929). Although not considered in the simulation,

convective losses associated with gases exiting the solar reactor during the reduction step are also

negligible (less than 0.3% of the input power). The share of energy used to drive the reduction

reaction, and therefore also representative of the solar-to-fuel energy conversion efficiency, could

potentially be increased by using doped ceria to increase the reduction extent [101,102], or by

minimizing the temperature swing with near isothermal operation [103–105], although this does not

necessarily increase the efficiency due to other limitations introduced with a lower temperature

swing.

Figure 27: Instantaneous energy balance for the duration of a reduction step. Other heat losses include

convection and radiation at the outer reactor surfaces, reflection of incoming solar radiation inside the

reactor cavity, and absorption and reflection at the quartz window. Extracted from Ref. [100].

Solar-to-fuel energy conversion efficiency

The solar-to-fuel energy conversion efficiency is defined as

Qsolar is the total solar energy input integrated over the reduction step. Note that Qsolar is only

delivered during the endothermic reduction step. Assuming complete re-oxidation using CO2, the

energy content of the fuel (CO) produced is calculated as , where ΔHfuel is

the heating value of CO (ΔHfuel = 283 kJ mol−1) and is the rate of released O2 integrated over

the reduction step. Qpump and Qinert are the energy penalties associated with vacuum pumping

and the consumption of the inert gas Ar during the reduction step, respectively, and are

Deliverable Report

D 4.1 GA No. 731287 44 Co-funded by the Horizon 2020 Framework

Programme of the European Union

calculated as described in Ref. [106]. Note that ηsolar-to-fuel is weakly dependent on the assumptions

used for the calculation of these two energy penalties, because Qsolar is roughly 2 orders of

magnitude larger than Qpump and Qinert.

An efficiency of ηsolar-to-fuel = 3.4% was predicted by the simulation, which is comparable to the

experimentally determined efficiency ηsolar-to-fuel = 3.2% (for Psolar = 30.5 kW). Heat recovery was not

applied. The slight overestimation is correlated directly to the slight overestimation in total O2 yield

from the simulation. The maximum ηsolar-to-fuel reported to date is 5.25% [107]. The numerical results

indicate the prominent influence of solar radiative input power, where increasing it substantially

reduces reduction time of the cerium dioxide structure. Consequently, the model predicts a solar-

to-fuel energy conversion efficiency of >6% at a solar radiative power input of 50 kW; efficiency

>10% can be obtained provided the RPC macroporosity is substantially increased, and better

volumetric absorption and uniform heating is achieved. Managing the ceria surface temperature

during reduction to avoid sublimation is a critical design consideration for direct absorption solar

receiver–reactors.

4.2. Thermal performance of a lab-scale ceramic volumetric air receiver

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [108].

Under the framework of the project INSHIP, a window-less ceramic cavity receiver is being

developed and tested at ETH Zurich to demonstrate supply of air as the heat transfer fluid for

industrial processes at high temperatures (>1000°C). An experimental setup has been developed to

test a lab-scale receiver. The cavity was made up of ceramic foam, namely reticulate porous

ceramic (RPC). Two different RPC materials were tested: ceria (~7 ppi (pores per inch)) and silicon-

infiltrated silicon carbide (SiSiC, 10 ppi). The goal was to demonstrate the proof-of-concept of the

open cavity air receiver and deliver air at high temperatures with reasonable receiver thermal

efficiencies.

A schematic of the receiver is depicted in Figure 28. The receiver consists of a cavity made up of

RPC material. In this work, pure ceria and silicon-infiltrated silicon carbide (SiSiC) are the materials

used as RPC materials. The receiver has a -diameter aperture to allow the entry of solar

radiation as well as ambient air. The cavity is composed of eight trapezoidal bricks, assembled to

form an octagonal cross-section, closed on the rear end with an octagonal cap, all made up of

the same material. The cavity is lined by fibrous ceria and Al2O3-SiO2 insulation layers. Between the

volumetric absorber and the fibrous ceria insulation, an air gap of is maintained to allow fluid

flow around the RPC. The outer shell of the receiver is made of Inconel 600 and the front is closed

by an Al2O3-SiO2 insulation disk and a water-cooled aluminum front plate.

Deliverable Report

D 4.1 GA No. 731287 45 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Outer Shell, Inconel 600

Al2O3-SiO2

insulation

Water-cooledaluminumfront plate

CeO2 fibrousinsulation

Aperture

Air Outlet

Volumetricabsorber

Figure 28: Receiver Schematic. Extracted from Ref. [108].

Schematic of the experimental setup is shown in Figure 29. The receiver was tested at the High-Flux

Solar Simulator of ETH Zurich, with four solar power input levels between 2.5-5.0 kW at the receiver

aperture and for a range of air mass flow rates between 2-10 kg/h. A vacuum pump downstream

of the receiver suctions ambient air into the receiver. The “solar” power input is provided by a set of

xenon arc lamps.

Deliverable Report

D 4.1 GA No. 731287 46 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Steel T-joints

Alumina Tube

Ambient Air in

Solar Flux in

RPC thermocouples

Open Volumetric Receiver

Cold Air Supply (blocked)

Inconel Outlet

Hot Air Outlet

Thermocouple for Air Outlet

T T T

T T

T

Air Mass Flow

Meter

Pu

mp

Co

rro

gate

dP

ipes

To Exhaust

WaterIn

WaterOut

WaterIn

WaterOut

Infrared Camera

T

T

Quartz-glass HX

Shell-and-Tube HX

Figure 29: Schematic of the final experimental setup. Extracted from Ref. [108].

Thermal efficiency of the receiver

Figure 30 shows the thermal efficiency of the receiver as a function of air outlet temperature, for the

four power input levels and two RPC materials, for a range of air mass flow rates. Air outlet

temperatures between 500°C to 1150°C were achieved, with receiver thermal efficiencies

between 0.58 and 0.25. Higher air mass flow rates lead to lower air outlet temperatures but also to

higher thermal receiver efficiencies. It is important for such a receiver to extract as much incoming

heat as possible and deliver it at the receiver outlet at as high air temperatures as possible. From

Figure 30, it can be seen that for a high input power level ( ), air outlet temperatures of

with corresponding thermal receiver efficiencies are possible. The receiver

efficiency is defined as the ratio of the enthalpy rise of the air (from ambient to outlet temperature)

to the solar radiative power coming in through the aperture. The receiver has therefore a viable

trade-off between the air outlet temperature and the thermal receiver efficiency at such a power

input level. This hints towards the level of input power at which this receiver can be operated to

extract high air outlet temperatures efficiently.

Deliverable Report

D 4.1 GA No. 731287 47 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 30: Thermal receiver efficiency over the air outlet temperature for different materials and power input

levels. Extracted from Ref. [108].

The difference between the two volumetric absorber materials is not clearly apparent from these

experiments. Only at low input power levels, SiSiC produces slightly higher air outlet temperatures at

air mass flow rates above than ceria RPC. For high input power levels, the two materials show

similar performances. The subtle differences in the performance of these two materials could be

explained when the overall heat transfer and fluid flow model of the receiver would be completed

and validated, which are also part of ongoing work under INSHIP. The validated model in turn

would assist in further improvements to the receiver design.

4.3. Mass and energy balances expected for a metallic volumetric

receiver

This section has been prepared by the Fondazione Bruno Kessler (FBK), Italy.

Among scientific literature, several works have presented different arrangements using volumetric

receivers for solar thermochemical fuel production[109][84][110]. In the first possible arrangement,

the reactor is directly impinged by the concentrated radiation; in this case, the reactor acts as a

volumetric receiver and is made by a reacting material. The second arrangement uses a heat

transfer fluid (HTF) to provide heat for the reaction, in this case, the volumetric receiver is directly

irradiated and transfers the heat to the HTF. For the second arrangement, depending on the

volumetric receiver type, the HTF could be used in an open or closed cycle. The volumetric receiver

is a crucial device to obtain a good efficiency of the system. FBK has developed a new

hierarchically layered volumetric receiver that could suit as receiver of the solar fuel production

cycle. In Figure 31, a lateral view of the receiver is presented [111].

Deliverable Report

D 4.1 GA No. 731287 48 Co-funded by the Horizon 2020 Framework

Programme of the European Union

2 0 0

725 600

550

261

145

Irradiance from homogenizer kW/m2

150 200

260 550 600

725

Receiver

Iso-irradiance

Homogeneizer

Figure 31: (left) Lateral view of FBK’s absorber, (right) Irradiance distribution at the absorber aperture plane

[109].

The development of variable geometry receivers built from the fractal repetition of elementary cells

has been made possible by recent advances in additive manufacturing techniques. These allow

for the design of creative highly-customized components based on complex geometries with

improved functionality. Selective laser melting (SLM), which utilizes a laser beam to melt successive

layers of powder into a finished part, has been the manufacturing technique employed here.

Receivers are manufactured in AISI 316L stainless steel, but Inconel 625 components can also be

produced. Porosity can be varied over a wide range of scales, the smallest pores achievable being

approximately 0.1mm in diameter. Convective heat transfer can be thus enhanced by employing if

required, augmentation features in this submicron scale. Additionally, variable porosity geometries

allow thermal conduction to be geometry-driven (by adjusting solid cross-sectional areas and

creating bottlenecks for the outward flow of heat) as well as material-driven.

A simplified system of a solar chemical reactor, based on the metallic volumetric receiver is

composed of i) solar dish as the dish developed during the CONTEST project, ii) the FBK hierarchical

receiver, iii) a blower that provides the air circulation to a iv) reactor, see Figure 32. Usually, the

system is designed to have some air return channels in order to maximize the hot air temperature at

the outlet of the receiver. The power input to receiver comes from the solar dish, as CONTEST dish,

that has a theoretical mirroring surface of 59m2. The efficiency of the concentrator (reflectivity,

tracking, and fouling) could be assumed to 0.85. The current focal area is designed to be 0.045 m2

with a geometrical concentration factor of 1311. The total power that hits the focal area is 47,2kW

with a DNI of 800W/m2. Now, the concentration energy is too high to guarantee the thermal and

mechanical stability of a metallic receiver (AISI 316L melting point is 1375 °C). Therefore, a full

concentration of the sun on the receiver should be avoided. Experiment conducted at the IMDEA

Energy Institute using calorimetric facility reveals that an irradiance of 700 kW/m2 can be tolerated

by the volumetric receiver based on AISI 316L [109].

Deliverable Report

D 4.1 GA No. 731287 49 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 32: Simplified System Chart of solar chemical reaction equipment

Figure 31 shows the total irradiance distribution on the solar volumetric receiver with a flux peak of

702 kW/m2. The typical temperature reached by air in this apparatus was in the range between

800-1200K. In this range of temperature, the efficiency can be calculated as the rate of enthalpy

rise of the working fluid divided by the total incident power. It thus gives an indication of the

capability of solar absorbers to transform incident solar power into enthalpy of the working fluid

through a series of radiative-to-convective heat transfer processes.

As is shown in Figure 33, the efficiency generally decreases with operating temperature, mainly

because of thermal re-radiation losses, proportional to the fourth power of temperature. The

maximum efficiency is 0.732. Taking into account the maximum incident irradiation power (650W)

the corresponding value of Q /m ̇0 at the best efficiency is 405.6 kJ kg−1, with corresponding working

fluid temperature rise of 285.8 K. Maximum thermal conversion efficiencies are observed in general

around 400 kJ kg−1 with temperature rises of the working fluid below 300 K in all cases, which further

proves the pronounced trade-off between thermal conversion efficiency and air outlet

temperature.

Using the calculated efficiencies, we can estimate in first approximation the energy and the mass

balances on a real dish (assuming an enough thermomechanical stability of the receiver), Figure

34.

Deliverable Report

D 4.1 GA No. 731287 50 Co-funded by the Horizon 2020 Framework

Programme of the European Union

0.4 0.6 0.8 1.0 1.2

0.40

0.45

0.50

0.55

0.60

0.65

0.70

0.75

Incident radiative power

350 W

500 W

650 WE

ffic

ien

cy

Mass Flow Rate [g/s]

Figure 33: Thermal conversion efficiency of the volumetric receiver as a function of incident radiative power

and mass flow rate

Figure 34: Sketch of the transformation of the solar irradiance solar chemical reactor with a thermal conversion

efficiency

Considering solar irradiance on the dish of 800 W/m2 and a mirror reflectance of 0.85, a 25%

energy loss due to fill factor induced by geometry of volumetric receiver, and 73,2% receiver

efficiency around 22kW can be transferred to the heated fluid in the cycle, which results to an air

mass flow rate of 52,8g/s.

The authors underline that the experiment is limited to a single receiver with a focused area of 0.045

m2 and for a limited number of the conditions. Further investigation using computational

experiments is necessary to estimate in accurate mode the total energy balance and mass

balance of a real solar dish, such as the CONTEST installation.

Deliverable Report

D 4.1 GA No. 731287 51 Co-funded by the Horizon 2020 Framework

Programme of the European Union

5. Fuels production from agricultural waste via gasification

The following sections provide an overview of the work carried out by INSHIP members under Task

4.3 of WP4 (Solar fuel production for the transportation sector).

Among existing biomass conversion technologies, biomass gasification offers high conversion

efficiency and the possibility to handle different bio-feedstock for a wide variety of applications

(heat, electricity, chemicals and transport fuels). Biomass gasification is a thermochemical process

that converts, through partial oxidation at high temperatures, solid biomass into fuel gas, small

quantities of char, ash and several condensable compounds (tars and oils). A promising alternative

to the conventional gasification process is represented by solar-assisted steam-only gasification,

which can convert carbonaceous materials into a quality syngas, high in H2 content and CO2 free.

Concentrated solar energy is used to drive the highly endothermic gasification reactions. Various

solar reactor concepts have been investigated as described in the following sections.

5.1. Solar reactor concepts for gasification of carbonaceous materials

This section has been prepared by the Centro de Investigaciones Energéticas, Medioambientales y

Tecnológicas (CIEMAT), Spain

Solar reactors for biomass gasification

There are many varying ideas for optimizing the use of concentrated solar energy to gasify biomass.

Designs that have been proposed are packed bed, fluidized bed, entrained flow, and drop-tube.

Some advantages and limitations of different reactor configurations are shown in Table 17.

Table 17 Advantages and limitations of different gasifier types

Gasifier Type Advantages Limitations

Packed Bed Simple design At low temperature, more

tar produced

Long solid residence time Limitations of mass

and heat transfer

Ash buildup

Fluidized Bed Compact construction Gas stream contains fine

particles of dust

Uniform temperature profile Complex system due to low

biomass hold up in the fuel

bed

Accept fuel size variation Variety of biomass can be

used but fuel flexibility is

applicable for biomass of

0.1 cm to 1 cm size

High ash melting point of biomass dos

not lead to clinker formation

Entrained flow Applicable to large systems High investment

Short residence time Strict fuel requirements

Deliverable Report

D 4.1 GA No. 731287 52 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Mass and Energy Balances

The global reaction of steam gasification process of a biomass fuel can be expressed as:

)(42222 sCCHOHCOCOHOHOCH ba

where, CHaOb represents the biomass fuel. A simplified mass balance is given in the following

equation:

CCHOHCOCOHOHOCH nnnnnnnnba

42222

where n represents the molar flow of the biomass, steam, residual carbon and the gas products. A

heat balance is also provided considering that there are several inputs and outputs associated with

a solar gasifier that can be considered to examine the energetic and efficiencies of the gasifier,

allowing for comparison with other solar and traditional gasifier systems: inputs are solar heat,

steam, and biomass, while the output is a product gas and char.

A brief heat balance is presented for calculating the efficiency of the cavity reactor:

lossesnetinput QQQ

The Q input includes the heat transferred to the reactants Qnet as well as the major heat losses. Q losses

account for thermal energy losses by radiation (Qrad), re-radiation (Qrerad) and convection (Qother).

The net heat transfer to the reactants Qnet was determined by an energy balance across the

reactor, considering heat flows entering the reactor and leaving it.

heatproductccii isteamsteamOCHOCHnet QHnHnHnHnQbaba ,

The unreacted char is labelled as c and it was assumed to be pure carbon. Gas species i are H2,

CO, CO2, H2O and CH4 and Q product heat represents the sensible heat of the flowing gas.

The enthalpies of the biomass ΔHCHaOb and the char ΔHc are based on their LHV and composition.

OCHOCHOHCOOCH aabaLHVmHHH

22 21

charcharCOc LHVmHH 2

The energy conversion efficiency is defined by:

baba OCHOCHsolar

heatproductccii i

LHVmQ

QHnHn

,

Our preliminary studies conclude that the best energy conversion efficiency was found for packed-

bed reactor configuration with a value of 30%. However, this overview indicates that their capacity

ranges are only a few kilowatts (2–25 kW) or even smaller.

Some works for process integration in the existing solar technologies have been given by Segal and

Epstein [112] and Piatkowski et al. [113] with a beam-down solar concept as the preferred option to

packed bed, etc.

The beam-down concept has some advantages over standard tower-top reactor arrangements. It

allows a large scale reactor to be built on the ground and the solar radiation to enter the reactor

through a window. Another advantage of this arrangement is that it allows for a large heliostat field

Deliverable Report

D 4.1 GA No. 731287 53 Co-funded by the Horizon 2020 Framework

Programme of the European Union

compared with that in conventional tower system, thus much higher solar energy can be obtained

into a one receiver with high concentration. According to Siegel et al. [114], The beam-down

secondary optics perform with an optical efficiency of 93% resulting in an overall collection

efficiency.

Several measures of efficiency can be used in examining a solar gasifier plant performance. A solar

efficiency can be defined based on the equation as follows, indicating that the efficiency of the

gasifier is dependent on the efficiency of the solar technology.

beamdownsolar

The resulting value for the solar for a packed bed coupled with a beam-down concept would be

about 28%.

5.2. Co-production of syngas and potassium-based fertilizer by solar-driven

thermochemical conversion of crop residues

This section has been prepared by the Swiss Federal Institute of Technology (ETH), Zurich,

Switzerland based on Ref. [115].

This work [115], conducted under the INSHIP project, reports on the thermochemical conversion of

inedible crop residues using concentrated solar energy as the source of high-temperature process

heat. Experiments were performed using a 5 kWth solar packed-bed reactor exposed to radiative

fluxes up to 1788 suns (Figure 35). The waste biomass feedstock consisted of unprocessed batches

of cotton boll, soybean husk, and black mustard husk and straw, which were pyrolysed and steam-

based gasified at nominal temperatures in the range 879–1266 °C, yielding high-quality syngas with

molar ratios in the range H2:CO = 1.43–3.25, CO2:CO = 0.28–1.40, and CH4:CO = 0.03–0.28.

Deliverable Report

D 4.1 GA No. 731287 54 Co-funded by the Horizon 2020 Framework

Programme of the European Union

Figure 35: (left) Schematic of the solar packed-bed reactor, (right) Solar experimental setup: (T)

thermocouples, (1) peristaltic pump, (2) steam generator, (3) electrically heated tubes, (4) mass flow

controllers, (5) spring over-pressure valve and pressure transmitter, (6) high-flux solar simulator, (7) mechanical

shutter, (8) deflection mirror, (9) tar cracker, (10) condensation tube, (11) scrubber, (12) filter. Extracted from

Ref. [115].

The carbon conversion, XC, is defined as the portion of the initial mass of carbon present in the

feedstock that has been converted:

C,residue

C

C,feedstock

1m

Xm

The carbon-gas yield, Ygas, is defined as the portion of the initial molar amount of carbon present in

the feedstock that has been converted to CO, CO2, and CH4:

2 4CO CO CH

gas

C,feedstock

n n nY

n

The solar-to-fuel energy conversion efficiency, ηsolar-to-fuel, is defined as the ratio of the heating value

of the syngas produced to the solar radiative energy input and the heating value of the feedstock

[116]:

syngas syngas

solar-to-fuel

solar feedstock feedstock

m LHV

Q m LHV

,

where Qsolar = ʃ Q̇solar dt, integrated over the duration of the entire experimental run. The sensible

heat required for steam generation is assumed to be generated using the waste heat from the

exhaust gases. The energetic upgrade factor, U, is defined as the ratio of the heating value of the

syngas produced to that of the feedstock [116]:

Deliverable Report

D 4.1 GA No. 731287 55 Co-funded by the Horizon 2020 Framework

Programme of the European Union

syngas syngas

feedstock feedstock

m LHVU

m LHV

Table 18 – Properties of the produced soybean husk charcoal and the operational conditions and

results of the solar pyrolysis of soybean husk and of the solar gasification of soybean husk charcoal.

Extracted from Ref. [115].

Characterization of the SBH charcoal

Ultimate (daf)

C 75.31 wt%

H 2.60 wt%

N 2.64 wt%

O 19.29 wt%

S 0.16 wt%

Proximate

Fixed

carbon 58.50 wt%

Volatiles 21.90 wt%

Moisture* 3.96 wt%

Ash 15.62 wt%

LHV 20.32 MJ/kg

SSA 5.15 ± 0.15 m2/g

Operational conditions and results

Pyrolysis: Gasification:

Tabsorber 1252 °C

Ttop,peak 1035 °C Ttop 1209 °C

mfeedstock 356 g mfeedstock 102 g

nH2 4.13 mol nH2 5.91 mol

nCO 3.21 mol nCO 3.57 mol

nCO2 1.88 mol nCO2 0.91 mol

nCH4 1.11 mol nCH4 0.07 mol

mresidue 99 g mresidue 10.2 g

∆t 36 min ∆t 98 min

XC 55.8 % XC 99.9 %

Ygas 54.6 % Ygas 87.9 %

ηsolar-to-fuel 30.7 % ηsolar-to-fuel 9.9 %

U 0.80 U 1.2

*Moisture content is not considered in the pyrolysis run.

The solar-to-fuel energy conversion efficiency, defined as the ratio of the heating value of the

syngas produced to the solar radiative energy input and the heating value of the feedstock,

reached 18%. The heating value of the feedstock was solar-upgraded by 7%, thus outperforming

autothermal gasification that typically downgrades by at least 15%. The ash contained 23%

Deliverable Report

D 4.1 GA No. 731287 56 Co-funded by the Horizon 2020 Framework

Programme of the European Union

potassium. The solar-driven thermochemical process offers a sustainable and efficient path for the

conversion of agricultural wastes into valuable fuels and soil fertilizers.

Summary

The technical feasibility of the solar packed-bed reactor was demonstrated for the pyrolysis and

gasification of unprocessed cotton boll, soybean husk, and black mustard husk and straw. The

quality of the syngas produced was substantially higher than the one resulting from autothermal

gasification [117–119], and was not significantly influenced by the characteristics of the

heterogeneous feedstocks. Energetic upgrade factors above 1 further supported the advantages

of solar-driven gasification vis-à-vis autothermal gasification. The relatively high values obtained for

the solar-to-fuel energy conversion efficiency at the 5 kWth laboratory scale pointed out to the

economic potential of the technology; higher values are expected for a scale-up because of the

larger volume-to-area ratio of the reaction chamber [120]. The ash was shown to be suitable for K-

based fertilizers. In sum, the solar-driven valorization of inedible crop residues is beneficial for the

local and global environment, as it co-produces a carbon-neutral synthetic fuel and a natural

fertilizer using renewable energy and without competing with the food chain.

Deliverable Report

D 4.1 GA No. 731287 57 Co-funded by the Horizon 2020 Framework

Programme of the European Union

6. Degree of progress

This document constitutes Deliverable D4.1 under Work Package 4 of the INSHIP project. This

Deliverable reports results of ongoing work related to mass and energy balances of selected high-

temperature industrial processes driven by concentrated solar process heat and related solar

chemical reactor concepts thereof. Further progress is expected from these research activities until

the end of the project (end of the year 2020).

Deliverable Report

D 4.1 GA No. 731287 58 Co-funded by the Horizon 2020 Framework

Programme of the European Union

7. Dissemination

Parts of this document that are directly related to work carried out under the INSHIP project have

been disseminated through the following publications:

Journal articles

Davis D., Müller F., Saw W., Steinfeld A., Nathan G.J., “Solar-driven alumina calcination for CO2

mitigation and improved product quality”, Green Chemistry, Vol. 19, pp. 2992-3005, 2017.

F. Müller, H. Patel, D. Blumenthal, P. Poživil, P. Das, C. Wieckert, P. Maiti, S. Maiti, A. Steinfeld, Co-

production of syngas and potassium-based fertilizer by solar-driven thermochemical conversion of

crop residues, Fuel Process. Technol. 171 (2018) 89–99. doi:10.1016/J.FUPROC.2017.08.006.

Muhich C., Blaser S., Hoes M., Steinfeld A., “Comparing the solar-to-fuel energy conversion

efficiency of ceria and perovskite based thermochemical redox cycles for splitting H2O and CO2”,

International Journal of Hydrogen Energy, Vol. 43, pp. 18814-18831, 2018.

S. Luque, G. Menéndez, M. Roccabruna, J. González-Aguilar, L. Crema, M. Romero, Exploiting

volumetric effects in novel additively manufactured open solar receivers, Sol. Energy. 174 (2018)

342–351. doi:10.1016/j.solener.2018.09.030.

Conference proceedings/presentations

E. Johnson, D. Baker, I. Tari, Development of View Factor Correlations for Modeling Thermal

Radiation in Solid Particle Solar Receivers Using CFD-DEM, AIP Conf. Proc. (2018).

E. Johnson, D. Baker, I. Tari, Proposal of a novel gravity-fed, particle-filled solar receiver, in: AIP Conf.

Proc., 2017. doi:10.1063/1.4984371.

V. Patil and A. Steinfeld, Heat Transfer and Fluid Flow Analysis of a Volumetric Ceramic Solar

Receiver for High-Temperature Processing, ASME Energy Sustainability Conference, Orlando, USA

(2018)

V. Patil and A. Steinfeld, “Ceramic volumetric receivers for high-temperature solar process heat:

investigating the potential of anisotropic morphology”, 13th SOLLAB Doctoral Colloquium on Solar

Concentrating Technologies, Berlin, Germany, May 15-17, 2017.

Deliverable Report

D 4.1 GA No. 731287 59 Co-funded by the Horizon 2020 Framework

Programme of the European Union

8. References

[1] A. Kurucak, Flow Sheet Optimization By the Concept of Sustainable, METU, 2010.

[2] S. Avcu, G.K. Demir, M.E. Martı, M. Özgören, AA25 - Investigation of energy efficiency

improvements in alumina refinery of Seydiehir ETI Aluminium Plant.pdf, in: n.d.: pp. 1–8.

[3] M. Obaidat, A. Al-Ghandoor, P. Phelan, R. Villalobos, A. Alkhalidi, Energy and exergy analyses

of different aluminum reduction technologies, Sustain. 10 (2018) 1–21.

doi:10.3390/su10041216.

[4] V. Gusberti, D.S. Severo, B.J. Welch, M. Skyllas-Kazacos, Modelling the Aluminium Smelting

Cell Mass and Energy Balance–a Tool Based on the 1st Law of Thermodynamics, Proc. 10th

Aust. Alum. Smelt. Technol. Conf. Launceston, TAS. (2011) 1–23.

[5] E. Balomenos, D. Panias, I. Paspaliaris, Energy and exergy analysis of the primary aluminum

production processes: A review on current and future sustainability, Miner. Process. Extr.

Metall. Rev. 32 (2011) 69–89. doi:10.1080/08827508.2010.530721.

[6] E. Balomenos, D. Panias, Carbothermic Reduction of Alumina: A Review of Developed

Processes and Novel Concepts, Proc. EMC. (2011) 729–744.

[7] D. Davis, F. Müller, W.L. Saw, A. Steinfeld, G.J. Nathan, Solar-driven alumina calcination for

CO 2 mitigation and improved product quality, Green Chem. 19 (2017) 2992–3005.

doi:10.1039/C7GC00585G.

[8] K. Kermeli, P.-H. ter Weer, W. Crijns-Graus, E. Worrell, Energy efficiency improvement and GHG

abatement in the global production of primary aluminium, Energy Effic. 8 (2015) 629–666.

doi:10.1007/s12053-014-9301-7.

[9] The International Aluminium Institute, Statistics, (2016). http://www.world-

aluminium.org/statistics/ (accessed December 4, 2016).

[10] F. Williams, H.-W. Schmidt, Flash- and CFB Calciners, History and Difficulties of Development of

Two Calcination Technologies, in: Light Met. 2012, Springer International Publishing, Cham,

2012: pp. 135–140. doi:10.1007/978-3-319-48179-1_24.

[11] B.E. Raahauge, Smelter Grade Alumina Quality in 40+ Year Perspective: Where to From

Here?, in: Light Met. 2015, John Wiley & Sons, Inc., Hoboken, NJ, USA, 2015: pp. 73–78.

doi:10.1002/9781119093435.ch14.

[12] L.M. Perander, Z.D. Zujovic, T.F. Kemp, M.E. Smith, J.B. Metson, The nature and impacts of fines

in smelter-grade alumina, JOM. 61 (2009) 33–39. doi:10.1007/s11837-009-0164-x.

[13] K. Lovegrove, S. Edwards, N. Jacobson, J. Jordan, J. Petersheim, J. Rutowitz, H. Saddler, M.

Watt, J. Wyder, Renewable energy options for Australian industrial gas users, 2015.

[14] J. Barrie, B. Christine, Improvements in the Design and Operation of Alumina Flash Calciners,

IFRF Combust. J. (2001).

http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.599.1095&rep=rep1&type=pdf

(accessed December 14, 2018).

[15] C. Klett, L. Perander, Alumina Calcination: A Mature Technology Under Review from Supplier

Perspective, in: Light Met. 2015, Springer International Publishing, Cham, 2015: pp. 79–84.

doi:10.1007/978-3-319-48248-4_15.

Deliverable Report

D 4.1 GA No. 731287 60 Co-funded by the Horizon 2020 Framework

Programme of the European Union

[16] T. Eglinton, J. Hinkley, A. Beath, M. Dell’Amico, Potential Applications of Concentrated Solar

Thermal Technologies in the Australian Minerals Processing and Extractive Metallurgical

Industry, JOM. 65 (2013) 1710–1720. doi:10.1007/s11837-013-0707-z.

[17] D. Hirsch, A. Steinfeld, Radiative transfer in a solar chemical reactor for the co-production of

hydrogen and carbon by thermal decomposition of methane, Chem. Eng. Sci. 59 (2004)

5771–5778. doi:10.1016/J.CES.2004.06.022.

[18] R. Bertocchi, Carbon Particle Cloud Generation for a Solar Particle Receiver, J. Sol. Energy

Eng. 124 (2002) 230. doi:10.1115/1.1488666.

[19] L.A. Dombrovsky, W. Lipiński, A. Steinfeld, A diffusion-based approximate model for radiation

heat transfer in a solar thermochemical reactor, J. Quant. Spectrosc. Radiat. Transf. 103

(2007) 601–610. doi:10.1016/J.JQSRT.2006.08.003.

[20] D. Mischler, A. Steinfeld, Nonisothermal Nongray Absorbing-Emitting-Scattering Suspension of

Fe3O4 Particles Under Concentrated Solar Irradiation, J. Heat Transfer. 117 (1995) 346.

doi:10.1115/1.2822528.

[21] A. Z’Graggen, P. Haueter, D. Trommer, M. Romero, J.C. de Jesus, A. Steinfeld, Hydrogen

production by steam-gasification of petroleum coke using concentrated solar power—II

Reactor design, testing, and modeling, Int. J. Hydrogen Energy. 31 (2006) 797–811.

doi:10.1016/J.IJHYDENE.2005.06.011.

[22] I. Padilla, A. López-Delgado, S. López-Andrés, M. Álvarez, R. Galindo, A.J. Vazquez-

Vaamonde, The application of thermal solar energy to high temperature processes: case

study of the synthesis of alumina from boehmite., Sci. World J. 2014 (2014) 825745.

doi:10.1155/2014/825745.

[23] A. Steinfeld, A. Imhof, D. Mischler, Experimental Investigation of an Atmospheric-Open

Cyclone Solar Reactor for Solid-Gas Thermochemical Reactions, J. Sol. Energy Eng. 114 (1992)

171–174. doi:10.1115/1.2930001.

[24] A. Meier, E. Bonaldi, G.M. Cella, W. Lipinski, D. Wuillemin, Solar chemical reactor technology

for industrial production of lime, Sol. Energy. 80 (2006) 1355–1362.

doi:10.1016/J.SOLENER.2005.05.017.

[25] V. Nikulshina, M. Halmann, A. Steinfeld, Coproduction of Syngas and Lime by Combined

CaCO3-Calcination and CH4-Reforming Using a Particle-Flow Reactor Driven by

Concentrated Solar Radiation, Energy & Fuels. 23 (2009) 6207–6212. doi:10.1021/ef9007246.

[26] G. Flamant, D. Hernandez, C. Bonet, J.-P. Traverse, Experimental aspects of the

thermochemical conversion of solar energy; Decarbonation of CaCO3, Sol. Energy. 24 (1980)

385–395. doi:10.1016/0038-092X(80)90301-1.

[27] H. Wang, B. Xu, P. Smith, M. Davies, L. DeSilva, C. Wingate, Kinetic modelling of gibbsite

dehydration/amorphization in the temperature range 823–923 K, J. Phys. Chem. Solids. 67

(2006) 2567–2582. doi:10.1016/J.JPCS.2006.07.016.

[28] K. Wefers, C. Misra, Oxides and hydroxides of aluminum, Report 19, Alcoa, 1987.

http://epsc511.wustl.edu/Aluminum_Oxides_Alcoa1987.pdf (accessed December 14, 2018).

[29] M. Kruesi, M.E. Galvez, M. Halmann, A. Steinfeld, Solar Aluminum Production by Vacuum

Carbothermal Reduction of Alumina—Thermodynamic and Experimental Analyses, Metall.

Mater. Trans. B. 42 (2011) 254–260. doi:10.1007/s11663-010-9461-6.

[30] International Aluminum Institute, International Aluminum Institute, Aluminum for Future

Generations Sustainability Update 2007., (2007). http://www.world-aluminium.org.

Deliverable Report

D 4.1 GA No. 731287 61 Co-funded by the Horizon 2020 Framework

Programme of the European Union

[31] W.T. Choate, J.A.S. Green, Light Metals 2006 Volume 2: Aluminum Reduction Technology, T.J.

Galloway Ed, 2006.

[32] W.T. Choate, J.A.S. Green, U.S. Energy Requirements for Aluminum Production, Historical

Perspective, Theoretical Limits and New Opportunities, 2003, 2003.

http://www1.eere.energy.gov/industry/aluminum/pdfs/al_theoretical.pdf.

[33] J.H. Cox, L.M. Pidgeon, AN INVESTIGATION OF THE ALUMINUM–OXYGEN–CARBON SYSTEM,

Can. J. Chem. 41 (1963) 671–683. doi:10.1139/v63-095.

[34] M. Halmann, A. Frei, A. Steinfeld, Carbothermal reduction of alumina: Thermochemical

equilibrium calculations and experimental investigation, Energy. 32 (2007) 2420–2427.

doi:10.1016/j.energy.2007.06.002.

[35] R.J. Fruehan, Y. Li, G. Carkin, Mechanism and rate of reaction of Al2O, Al, and CO vapors

with carbon, Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 35B (2004) 617–623.

doi:10.1007/s11663-004-0002-z.

[36] R. Winand, M. Van Gysel, A. Fontana, L. Segers, J. Carlier, Production of Magnesium by

vacuum carbothermic reduction of calcined dolomite, Trans. Inst. Min. Metall. Sect. C -

Miner. Process. Extr. Metall. 99 (1990) C105–C112.

[37] J.M. Toguri, L.M. Pidgeon, HIGH-TEMPERATURE STUDIES OF METALLURGICAL PROCESSES: PART I.

THE THERMAL REDUCTION OF MAGNESIUM OXIDE WITH SILICON, Can. J. Chem. 39 (1961) 540–

547. doi:10.1139/v61-065.

[38] J.M. Toguri, L.M. Pidgeon, HIGH-TEMPERATURE STUDIES OF METALLURGICAL PROCESSES: PART II.

THE THERMAL REDUCTION OF CALCINED DOLOMITE WITH SILICON, Can. J. Chem. 40 (1962)

1769–1776. doi:10.1139/v62-271.

[39] S. Ramakrishnan, P. Koltun, Global warming impact of the magnesium produced in China

using the Pidgeon process, Resour. Conserv. Recycl. 42 (2004) 49–64.

doi:10.1016/J.RESCONREC.2004.02.003.

[40] P.G. Loutzenhiser, O. Tuerk, A. Steinfeld, Production of Si by vacuum carbothermal reduction

of SiO2 using concentrated solar energy, JOM. 62 (2010) 49–54. doi:10.1007/s11837-010-0137-

0.

[41] A. Steinfeld, High-temperature solar thermochemistry for CO2 mitigation in the extractive

metallurgical industry, Energy. (1997). doi:10.1016/S0360-5442(96)00103-X.

[42] J.P. Murray, Aluminum Production Using High-Temperature Solar Process Heat, Sol. Energy. 66

(1999) 133–142. doi:10.1016/S0038-092X(99)00011-0.

[43] A. Steinfeld, P. Kuhn, J. Karni, High-temperature solar thermochemistry: Production of iron and

synthesis gas by Fe3O4-reduction with methane, Energy. 18 (1993) 239–249. doi:10.1016/0360-

5442(93)90108-P.

[44] A. Steinfeld, M. Brack, A. Meier, A. Weidenkaff, D. Wuillemin, A solar chemical reactor for co-

production of zinc and synthesis gas, Energy. 23 (1998) 803–814. doi:10.1016/S0360-

5442(98)00026-7.

[45] T. Osinga, U. Frommherz, A. Steinfeld, C. Wieckert, Experimental Investigation of the Solar

Carbothermic Reduction of ZnO Using a Two-cavity Solar Reactor, J. Sol. Energy Eng. 126

(2004) 633–637. doi:10.1115/1.1639001.

[46] J.P. Murray, A. Steinfeld, E.A. Fletcher, Metals, nitrides, and carbides via solar carbothermal

reduction of metal oxides, Energy. 20 (1995) 695–704. doi:10.1016/0360-5442(95)00032-C.

Deliverable Report

D 4.1 GA No. 731287 62 Co-funded by the Horizon 2020 Framework

Programme of the European Union

[47] J.P. Murray, Solar Production of Aluminum by Direct Reduction: Preliminary Results for Two

Processes, J. Sol. Energy Eng. 123 (2001) 125. doi:10.1115/1.1351809.

[48] M.E. Gálvez, A. Frei, F. Meier, A. Steinfeld, Production of AlN by carbothermal and

methanothermal reduction of Al2O3in a N2flow using concentrated thermal radiation, Ind.

Eng. Chem. Res. 48 (2009) 528–533. doi:10.1021/ie8011193.

[49] Centre for Research in Computational Thermochemistry Ecole Polytechnique de Montreal

Canada, FactSage - Thermochemical Software & Database Package, (n.d.).

www.crct.polymtl.ca.

[50] NIST, Standard Reference Data Program, Chemistry Webbook, (n.d.).

http://webbook.nist.gov/chemistry.

[51] A. Roine, HSC Chemistry 5, Outokumpu Research, Finland, n.d.

[52] M.E. Gálvez, A. Frei, G. Albisetti, G. Lunardi, A. Steinfeld, Solar hydrogen production via a two-

step thermochemical process based on MgO/Mg redox reactions—Thermodynamic and

kinetic analyses, Int. J. Hydrogen Energy. 33 (2008) 2880–2890.

doi:10.1016/J.IJHYDENE.2008.04.007.

[53] H. Li, S. Xie, Research on development of dolomite-ferrosilicon thermal reduction process of

magnesium production, J. Rare Earths. 23 (2005) 606–610.

[54] G. Brooks, S. Trang, P. Witt, M.N.H. Khan, M. Nagle, The carbothermic route to magnesium,

JOM. 58 (2006) 51–55. doi:10.1007/s11837-006-0024-x.

[55] L.H. Prentice, M.W. Nagle, T.R.D. Barton, S. Tassios, B.T. Kuan, P.J. Witt, K.K. Constanti-Carey,

Carbothermal Production of Magnesium: Csiro’s MagsonicTM Process, in: Magnes. Technol.

2012, Springer International Publishing, Cham, 2012: pp. 31–35. doi:10.1007/978-3-319-48203-

3_6.

[56] J.W. Evans, The evolution of technology for light metals over the last 50 years: Al, Mg, and Li,

JOM. 59 (2007) 30–38. doi:10.1007/s11837-007-0020-9.

[57] A. Argyn, Material and Heat Balance Calculations of Eti-Bakir Plant By Computer a Thesis,

METU, 2008.

[58] B.T.K. Barry, C.J. Thwaites, Tin and its alloys and compounds, Ellis Horwood, 1983.

[59] C.L. Mantell, Tin: Its Mining, Production, Technology and Application, Hafner Publishing Co

Ltd, 1970.

[60] G.G. Graf, Tin, Tin Alloys, and Tin Compounds, in: Ullmann’s Encycl. Ind. Chem., Wiley-VCH

Verlag GmbH & Co. KGaA, Weinheim, Germany, 2000. doi:10.1002/14356007.a27_049.

[61] A. Meier, N. Gremaud, A. Steinfeld, Economic evaluation of the industrial solar production of

lime, Energy Convers. Manag. 46 (2005) 905–926. doi:10.1016/J.ENCONMAN.2004.06.005.

[62] IEA, Technology Roadmap: Low-Carbon Transition in the Cement Industry, (2018) 66.

doi:10.1007/SpringerReference_7300.

[63] T. Engin, V. Ari, Energy auditing and recovery for dry type cement rotary kiln systems - A case

study, Energy Convers. Manag. 46 (2005) 551–562. doi:10.1016/j.enconman.2004.04.007.

[64] A. Atmaca, R. Yumrutaş, Analysis of the parameters affecting energy consumption of a rotary

kiln in cement industry, Appl. Therm. Eng. 66 (2014) 435–444.

doi:10.1016/j.applthermaleng.2014.02.038.

Deliverable Report

D 4.1 GA No. 731287 63 Co-funded by the Horizon 2020 Framework

Programme of the European Union

[65] A. Atmaca, R. Yumrutaş, Thermodynamic and exergoeconomic analysis of a cement plant:

Part II - Application, Energy Convers. Manag. 79 (2014) 799–808.

doi:10.1016/j.enconman.2013.11.054.

[66] A. Atmaca, R. Yumrutaş, Thermodynamic and exergoeconomic analysis of a cement plant:

Part i - Methodology, Energy Convers. Manag. 79 (2014) 790–798.

doi:10.1016/j.enconman.2013.11.053.

[67] World Business Council for Sustainable Development (WBCSD), Getting the Numbers Right

(GNR) - Global Cement Database on CO₂ and Energy Information Geographic coverage by

country and region, GNR Proj. Report. CO2. (2016). https://www.wbcsdcement.org/GNR-

2016/ (accessed February 7, 2018).

[68] G. Kabir, A.I. Abubakar, U.A. El-Nafaty, Energy audit and conservation opportunities for

pyroprocessing unit of a typical dry process cement plant, Energy. 35 (2010) 1237–1243.

doi:10.1016/j.energy.2009.11.003.

[69] A. Ustaoglu, M. Alptekin, M.E. Akay, Thermal and exergetic approach to wet type rotary kiln

process and evaluation of waste heat powered ORC (Organic Rankine Cycle), Appl. Therm.

Eng. 112 (2017) 281–295. doi:10.1016/j.applthermaleng.2016.10.053.

[70] Z. Cao, L. Shen, J. Zhao, L. Liu, S. Zhong, Y. Yang, Modeling the dynamic mechanism between

cement CO2emissions and clinker quality to realize low-carbon cement, Resour. Conserv.

Recycl. 113 (2016) 116–126. doi:10.1016/j.resconrec.2016.06.011.

[71] M.B. Ali, R. Saidur, M.S. Hossain, A review on emission analysis in cement industries, Renew.

Sustain. Energy Rev. 15 (2011) 2252–2261. doi:10.1016/j.rser.2011.02.014.

[72] N.A. Madlool, R. Saidur, M.S. Hossain, N.A. Rahim, A critical review on energy use and savings

in the cement industries, Renew. Sustain. Energy Rev. 15 (2011) 2042–2060.

doi:10.1016/j.rser.2011.01.005.

[73] C.K. Ho, J.M. Christian, J. Yellowhair, N. Siegel, S. Jeter, M. Golob, S.I. Abdel-Khalik, C. Nguyen,

H. Al-Ansary, On-sun testing of an advanced falling particle receiver system, AIP Conf. Proc.

1734 (2016). doi:10.1063/1.4949074.

[74] E. Johnson, D. Baker, I. Tari, Proposal of a novel gravity-fed, particle-filled solar receiver, in: AIP

Conf. Proc., 2017. doi:10.1063/1.4984371.

[75] G. Flamant, D. Gauthier, H. Benoit, J.-L. Sans, B. Boissière, R. Ansart, M. Hemati, A New Heat

Transfer Fluid for Concentrating Solar Systems: Particle Flow in Tubes, Energy Procedia. 49

(2014) 617–626. doi:10.1016/j.egypro.2014.03.067.

[76] C.K. Ho, A new generation of solid particle and other high-performance receiver designs for

concentrating solar thermal (CST) central tower systems, Adv. Conc. Sol. Therm. Res. Technol.

(2017) 107–128. doi:10.1016/B978-0-08-100516-3.00006-X.

[77] A.B. Morris, Z. Ma, S. Pannala, C.M. Hrenya, Simulations of heat transfer to solid particles

flowing through an array of heated tubes, Sol. Energy. 130 (2016) 101–115.

doi:10.1016/j.solener.2016.01.033.

[78] E. Johnson, D. Baker, I. Tari, Development of View Factor Correlations for Modeling Thermal

Radiation in Solid Particle Solar Receivers Using CFD-DEM, AIP Conf. Proc. (2018).

[79] C. Goniva, C. Kloss, A. Hager, S. Pirker, An open source CFD-DEM perspective, Proc.

OpenFOAM Work. (2010) 1–10.

[80] C. Kloss, C. Goniva, A. Hager, S. Amberger, S. Pirker, Models , algorithms and validation for

Deliverable Report

D 4.1 GA No. 731287 64 Co-funded by the Horizon 2020 Framework

Programme of the European Union

opensource DEM and CFD-DEM, Pcfd. 12 (2012) 140–152. doi:10.1504/PCFD.2012.047457.

[81] F.P. Incropera, D.P. DeWitt, T.L. Bergman, A.S. Lavine, Fundamentals of Heat and Mass

Transfer, John Wiley & Sons, 2007. doi:10.1016/j.applthermaleng.2011.03.022.

[82] A. Malinouski, Evaluation of particle-scale radiative heat transfer in polydisperse beds, (2017).

[83] A. Meier, E. Bonaldi, G.M. Cella, W. Lipinski, Multitube Rotary Kiln for the Industrial Solar

Production of Lime, J. Sol. Energy Eng. 127 (2005) 386. doi:10.1115/1.1979517.

[84] M. Romero, A. Steinfeld, Concentrating solar thermal power and thermochemical fuels,

Energy Environ. Sci. 5 (2012) 9234–9245. doi:10.1039/c2ee21275g.

[85] N.S. Lewis, D.G. Nocera, Powering the planet: chemical challenges in solar energy utilization.,

Proc. Natl. Acad. Sci. U. S. A. 103 (2006) 15729–35. doi:10.1073/pnas.0603395103.

[86] J.E. Miller, M.D. Allendorf, R.B. Diver, L.R. Evans, N.P. Siegel, J.N. Stuecker, Metal oxide

composites and structures for ultra-high temperature solar thermochemical cycles, J. Mater.

Sci. 43 (2008) 4714–4728. doi:10.1007/s10853-007-2354-7.

[87] T. Kodama, N. Gokon, Thermochemical Cycles for High-Temperature Solar Hydrogen

Production, Chem. Rev. 107 (2007) 4048–4077. doi:10.1021/cr050188a.

[88] M. Chambon, S. Abanades, G. Flamant, Solar thermal reduction of ZnO and SnO2:

Characterization of the recombination reaction with O2, Chem. Eng. Sci. 65 (2010) 3671–

3680. doi:10.1016/J.CES.2010.03.005.

[89] C. Perkins, A.W. Weimer, Solar-thermal production of renewable hydrogen, AIChE J. 55 (2009)

286–293. doi:10.1002/aic.11810.

[90] L.O. Schunk, A. Steinfeld, Kinetics of the thermal dissociation of ZnO exposed to

concentrated solar irradiation using a solar-driven thermogravimeter in the 1800-2100 K

range, AIChE J. 55 (2009) 1497–1504. doi:10.1002/aic.11765.

[91] D. Gstoehl, A. Brambilla, L.O. Schunk, A. Steinfeld, A quenching apparatus for the gaseous

products of the solar thermal dissociation of ZnO, J. Mater. Sci. 43 (2008) 4729–4736.

doi:10.1007/s10853-007-2351-x.

[92] P.G. Loutzenhiser, A. Meier, A. Steinfeld, Review of the Two-Step H2O/CO2-Splitting Solar

Thermochemical Cycle Based on Zn/ZnO Redox Reactions, Materials (Basel). 3 (2010) 4922–

4938. doi:10.3390/ma3114922.

[93] S. Abanades, G. Flamant, Thermochemical hydrogen production from a two-step solar-

driven water-splitting cycle based on cerium oxides, Sol. Energy. 80 (2006) 1611–1623.

doi:10.1016/J.SOLENER.2005.12.005.

[94] W.C. Chueh, S.M. Haile, Ceria as a Thermochemical Reaction Medium for Selectively

Generating Syngas or Methane from H2O and CO2, ChemSusChem. 2 (2009) 735–739.

doi:10.1002/cssc.200900138.

[95] W.C. Chueh, S.M. Haile, A thermochemical study of ceria: exploiting an old material for new

modes of energy conversion and CO2 mitigation, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci.

368 (2010) 3269–3294. doi:10.1098/rsta.2010.0114.

[96] P. Singh, M.S. Hegde, Ce 0.67 Cr 0.33 O 2.11 : A New Low-Temperature O 2 Evolution Material

and H 2 Generation Catalyst by Thermochemical Splitting of Water, Chem. Mater. 22 (2010)

762–768. doi:10.1021/cm9013305.

[97] H. Kaneko, T. Miura, H. Ishihara, S. Taku, T. Yokoyama, H. Nakajima, Y. Tamaura, Reactive

Deliverable Report

D 4.1 GA No. 731287 65 Co-funded by the Horizon 2020 Framework

Programme of the European Union

ceramics of CeO2–MOx (M=Mn, Fe, Ni, Cu) for H2 generation by two-step water splitting

using concentrated solar thermal energy, Energy. 32 (2007) 656–663.

doi:10.1016/J.ENERGY.2006.05.002.

[98] S. Abanades, A. Legal, A. Cordier, G. Peraudeau, G. Flamant, A. Julbe, Investigation of

reactive cerium-based oxides for H2 production by thermochemical two-step water-splitting,

J. Mater. Sci. 45 (2010) 4163–4173. doi:10.1007/s10853-010-4506-4.

[99] M. Zinkevich, D. Djurovic, F. Aldinger, Thermodynamic modelling of the cerium–oxygen

system, Solid State Ionics. 177 (2006) 989–1001. doi:10.1016/J.SSI.2006.02.044.

[100] S. Zoller, E. Koepf, P. Roos, A. Steinfeld, Heat Transfer Model of a 50 kW Solar Receiver–Reactor

for Thermochemical Redox Cycling Using Cerium Dioxide, J. Sol. Energy Eng. 141 (2019)

021014. doi:10.1115/1.4042059.

[101] J.R. Scheffe, R. Jacot, G.R. Patzke, A. Steinfeld, Synthesis, Characterization, and

Thermochemical Redox Performance of Hf 4+ , Zr 4+ , and Sc 3+ Doped Ceria for Splitting CO 2,

J. Phys. Chem. C. 117 (2013) 24104–24114. doi:10.1021/jp4050572.

[102] M. Takacs, J.R. Scheffe, A. Steinfeld, Oxygen nonstoichiometry and thermodynamic

characterization of Zr doped ceria in the 1573–1773 K temperature range, Phys. Chem.

Chem. Phys. 17 (2015) 7813–7822. doi:10.1039/C4CP04916K.

[103] R. Bader, R. Bala Chandran, L.J. Venstrom, S.J. Sedler, P.T. Krenzke, R.M. De Smith, A. Banerjee,

T.R. Chase, J.H. Davidson, W. Lipinski, Design of a Solar Reactor to Split CO 2 Via Isothermal

Redox Cycling of Ceria, J. Sol. Energy Eng. 137 (2015) 031007. doi:10.1115/1.4028917.

[104] R. Bader, L.J. Venstrom, J.H. Davidson, W. Lipiński, Thermodynamic Analysis of Isothermal

Redox Cycling of Ceria for Solar Fuel Production, Energy & Fuels. 27 (2013) 5533–5544.

doi:10.1021/ef400132d.

[105] Y. Hao, C.-K. Yang, S.M. Haile, High-temperature isothermal chemical cycling for solar-driven

fuel production, Phys. Chem. Chem. Phys. 15 (2013) 17084. doi:10.1039/c3cp53270d.

[106] P. Furler, A. Steinfeld, Heat transfer and fluid flow analysis of a 4 kW solar thermochemical

reactor for ceria redox cycling, Chem. Eng. Sci. 137 (2015) 373–383.

doi:10.1016/J.CES.2015.05.056.

[107] D. Marxer, P. Furler, M. Takacs, A. Steinfeld, Solar thermochemical splitting of CO 2 into

separate streams of CO and O 2 with high selectivity, stability, conversion, and efficiency,

Energy Environ. Sci. 10 (2017) 1142–1149. doi:10.1039/C6EE03776C.

[108] F. Kiener, Master Thesis: Window-less Ceramic Cavity Receiver: Assembly, Testing and Heat

Loss Analysis, ETH Zurich, 2019.

[109] S. Luque, G. Menéndez, M. Roccabruna, J. González-Aguilar, L. Crema, M. Romero, Exploiting

volumetric effects in novel additively manufactured open solar receivers, Sol. Energy. 174

(2018) 342–351. doi:10.1016/j.solener.2018.09.030.

[110] H.I. Villafán-Vidales, S. Abanades, C. Caliot, H. Romero-Paredes, Heat transfer simulation in a

thermochemical solar reactor based on a volumetric porous receiver, Appl. Therm. Eng. 31

(2011) 3377–3386. doi:10.1016/j.applthermaleng.2011.06.022.

[111] F. Alberti, S. Santiago, M. Roccabruna, S. Luque, J. Gonzalez-Aguilar, L. Crema, M. Romero,

Numerical analysis of radiation propagation in innovative volumetric receivers based on

selective laser melting techniques, AIP Conf. Proc. 1734 (2016). doi:10.1063/1.4949053.

[112] A. Segal, M. Epstein, Solar ground reformer, Sol. Energy. 75 (2003) 479–490.

Deliverable Report

D 4.1 GA No. 731287 66 Co-funded by the Horizon 2020 Framework

Programme of the European Union

doi:10.1016/J.SOLENER.2003.09.005.

[113] N. Piatkowski, C. Wieckert, A. Steinfeld, Experimental investigation of a packed-bed solar

reactor for the steam-gasification of carbonaceous feedstocks, Fuel Process. Technol. 90

(2009) 360–366. doi:10.1016/J.FUPROC.2008.10.007.

[114] N.P. Siegel, J.E. Miller, I. Ermanoski, R.B. Diver, E.B. Stechel, Factors Affecting the Efficiency of

Solar Driven Metal Oxide Thermochemical Cycles, Ind. Eng. Chem. Res. 52 (2013) 3276–3286.

doi:10.1021/ie400193q.

[115] F. Müller, H. Patel, D. Blumenthal, P. Poživil, P. Das, C. Wieckert, P. Maiti, S. Maiti, A. Steinfeld,

Co-production of syngas and potassium-based fertilizer by solar-driven thermochemical

conversion of crop residues, Fuel Process. Technol. 171 (2018) 89–99.

doi:10.1016/J.FUPROC.2017.08.006.

[116] N. Piatkowski, C. Wieckert, A.W. Weimer, A. Steinfeld, Solar-driven gasification of

carbonaceous feedstock-a review, Energy Environ. Sci. 4 (2011) 73–82.

doi:10.1039/C0EE00312C.

[117] Y. Zhao, S. Sun, H. Tian, J. Qian, F. Su, F. Ling, Characteristics of rice husk gasification in an

entrained flow reactor, Bioresour. Technol. 100 (2009) 6040–6044.

doi:10.1016/j.biortech.2009.06.030.

[118] S. Maiti, P. Bapat, P. Das, P.K. Ghosh, Feasibility study of jatropha shell gasification for captive

power generation in biodiesel production process from whole dry fruits, Fuel. 121 (2014) 126–

132. doi:10.1016/j.fuel.2013.12.048.

[119] H. Karatas, H. Olgun, F. Akgun, Experimental results of gasification of cotton stalk and

hazelnut shell in a bubbling fluidized bed gasifier under air and steam atmospheres, Fuel. 112

(2013) 494–501. doi:10.1016/j.fuel.2013.04.025.

[120] C. Wieckert, A. Obrist, P. von Zedtwitz, G. Maag, A. Steinfeld, Syngas Production by

Thermochemical Gasification of Carbonaceous Waste Materials in a 150 kW th Packed-Bed

Solar Reactor, Energy & Fuels. 27 (2013) 47770–4776. doi:10.1021/ef4008399.


Recommended