Sie sind auf Seite 1von 8

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

Solar Chemical Reactor Technology for the


Industrial Solar Production of Lime
Anton Meier a,*, Enrico Bonaldi b, Gian Mario Cella b, Wojciech Lipinski a,c, Daniel Wuillemin a
a

Solar Technology Laboratory, Paul Scherrer Institute, CH-5232 Villigen, Switzerland.


b
QualiCal Srl, Via Verdi 3, I-24121 Bergamo, Italy.
c
Department of Mechanical and Process Engineering, ETH-Swiss Federal Institute of Technology,
CH-8092 Zurich, Switzerland.
Abstract - We developed the solar chemical reactor technology to effect the endothermic
calcination reaction CaCO3 CaO + CO2 at 1200-1400 K. The indirect heating 10 kW
multi-tube rotary kiln prototype processed 1-5 mm limestone particles, producing high purity
lime that is not contaminated with combustion by-products. The quality of the solar produced
quicklime meets highest industrial standards in terms of reactivity (low, medium, and high)
and degree of calcination (exceeding 98%). The reactors chemical efficiency, defined as the
enthalpy of the calcination reaction at ambient temperature (3184 kJ kg-1) divided by the solar
energy input, reached 30-35% for quicklime production rates up to 4 kg h-1. The solar lime
reactor prototype operated reliably for more than 100 hours at solar flux inputs of about 2000
kW m-2, withstanding the thermal shocks that occur in solar high temperature applications.
By substituting concentrated solar energy for fossil fuels as the source of process heat, one
can reduce by 20% the CO2 emissions in a state-of-the-art lime plant and by 40% in a
conventional cement plant. The cost of solar lime produced in a 20 MWth industrial solar
calcination plant is estimated in the range 131-158 $/t, i.e. about 2-3 times the current selling
price of conventional lime.

1. Introduction
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 ( s ) CaO( s ) + CO2 ( g )

(1)

at the decomposition temperature near 1173 K is about 3029 kJ kg-1 of CaO, while the heat of dissociation
of calcite relative to 298 K is 3184 kJ kg-1 of CaO [1]. Total energy usage of modern lime kilns ranges from
3600 kJ kg-1 of CaO for vertical double shaft kilns to 7500 kJ kg-1 of CaO for non-preheated long rotary
kilns [1,2]. Because the process heat is traditionally supplied by the combustion of carbon-based fossil fuels
such as oil, coal, or natural gas, a lime or cement plant releases CO2 both as a calcination reaction product
and from the combustion process that supplies the energy for the reaction. According to the World Business
Council for Sustainable Development [3], the cement industry produces 5% of global man-made CO2
emissions, of which 50% is from the chemical process, and 40% from burning fuel. The remainder is split
between electricity and transport use. Estimates of the global anthropogenic CO2 emissions from the lime
industry are near 1% [1].
By substituting concentrated solar energy for the fossil fuels used to drive the calcination reaction, one can
reduce by 20% the CO2 emissions in a state-of-the-art lime plant and by 40% in a conventional cement
plant. The cleanliness of the solar process also leads to very pure lime or quicklime (CaO) that is not
contaminated by combustion by-products. It seems conceivable that a pure product may be advantageous
for special niche markets in industries that use high-value quicklime for the production of chemicals.
In this paper, we present the development of the industrial chemical reactor technology for the solar
production of lime. Based on previous experience with a direct heating rotary reactor prototype [4,5], we
*

Corresponding author. Tel.: +41-56-310-27-88; fax: +41-56-310-31-60. E-mail address: anton.meier@psi.ch

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

have designed, built, and tested a scaleable indirect heating 10 kW multi-tube rotary reactor prototype.
Furthermore, we have developed a conceptual design of a megawatt-scale solar lime plant comprising both
the solar lime reactor and the solar concentrating system.
2. Solar Reactor Technology
The novel indirect heating multi-tube rotary kiln, for which a patent application is pending [6], comprises a
multi-tube reaction chamber made from high-temperature resistant SiC (Fig. 1). The concentrated solar
radiation enters through the circular aperture of a water-cooled aluminum front shield and heats the
absorber tubes that are regularly arranged along the cylindrical cavity wall. At the rear of the reactor, a
solar irradiated SiC plate separates the cavity from the preheating chamber. Small-grained limestone
particles are continuously fed into this preheating chamber, where they are heated almost to the reaction
temperature. Due to the rotational movement, the limestone is then transported and calcined within the
absorber tubes that deliver thermal energy indirectly to the reactants.
absorber tubes

reactants inlet

pre-heating chamber

ceramic insulation
rotating cavity

products outlet

Fig. 1: Left: Schematic representation of the multi-tube rotary kiln prototype. Right: 10 kW solar lime kiln prototype
mounted on the experimental platform at PSIs solar furnace. The experimental set-up comprises the rotary kiln, which
is protected with a water-cooled shield; a white target for solar flux measurements; a reactants charging system (hopper,
dozer); and a products discharging system (collector, outlet tube). (Source: PSI)

The tilted rotary kiln is sitting on rubber wheels that are driven by a small electric motor. The kiln consists
of a 2 mm thick cylindrical steel drum of 418 mm length and 400 mm external diameter. The entire inner
part of the reactor was manufactured by INSULTECH AG, Switzerland. The absorber cavity of 225 mm
length and 252 mm inner diameter as well as the absorber tubes of 250 mm length and 17 mm internal
diameter are made from RSiC (Re-crystallized Silicon Carbide, density 2600-2800 kg m-3, maximum
operating temperature 1600 C). The reactor is lined with ceramic insulation material consisting of a 26
mm thick inner layer (INSULBOARD 1600: hard, density 170 kg m-3, maximum operating temperature
1600 C) and a 50 mm thick outer layer (INSULTHERM 1000: soft, porous, density 200 kg m-3, maximum
operating temperature 1000 C).
Figure 1 shows the experimental set-up for the solar calcination experiments in PSIs solar furnace
delivering solar power close to 15 kW with a peak solar flux concentration of about 3000 suns (1 sun = 1
kW m-2) on a focal spot of 8 cm diameter. The solar power input into the reactor and, therefore, the
limestone burning temperature inside the reaction chamber was controlled using a shutter between the suntracking heliostat and the stationary parabolic dish. The solar flux distribution on a white target was
measured with a CCD camera before and after each experiment, and the solar power entering the circular
aperture was computed using OPTIMAS from MediaCybernetics, USA, and calibration data obtained with a
Kendall radiometer. Temperatures within the rotary reactor were measured with type K thermocouples

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

(TC) and transmitted by wireless telemetry to a stationary receiver-decoder connected to a data acquisition
system. The MT32 Mini Telemetry system from KMT GmbH, Germany consisted of eight miniature TC
modules, an encoder, and a transmitter that were mounted on an air-cooled plate at the rear of the rotary
reactor. The thermocouple temperatures, the feeder speed of rotation and the reactor drum speed of rotation,
the direct normal solar irradiation, and the shutter position (opening angle) were simultaneously recorded
with the data acquisition system MessHaus from Delphin Technology AG, Germany.
Figure 2 is a typical representation of a solar experimental day showing solar irradiation and temperature
profiles within the rotary kiln. One can observe that it takes about 1.5-2 hours until steady state conditions
are reached at the desired temperature and that the highest temperatures are measured in the middle of the
reactor. Typically, a calcination experiment was running for 30 min with a constant reactant particle flow
rate and fixed experimental parameters. During well-defined time intervals, representative product samples
were taken and analyzed.
The indirect heating multi-tube rotary reactor also may be operated in hybrid mode using any conventional
external heat source like oil or gas burners, which may be introduced through the aperture into the cavity in
place of concentrated solar radiation. In fact, the type of the heat source does not influence the conditions
inside the indirect heating absorber tubes, i.e. the quality of the end product always remains the same. For
convenience, we tested the indirect heating multi-tube rotary reactor with an electric heating system
consisting of a heating element from KANTHAL International, Sweden. The cavity temperature (maximum
1200 C) could be controlled by adjusting the electric power (maximum 7 kW) supplied to the SiC heating
element.
1200

1200

1000

800

800

600

600

400

400

200

200

0
09:00:00

10:00:00

11:00:00

12:00:00

13:00:00

14:00:00

15:00:00

16:00:00

Irradiation [W/m2]

Temperature [C]

Day 2002-09-13
1000

Int_Front
Int_Middle
Int_Back
Preheat
Ext_Front
Ext_Middle
Ext_Back
Irradiation

0
17:00:00

Time [hh:mm:ss]

Fig. 2: Temperature profiles within the solar lime kiln prototype: 1) inside the absorber tubes (Int_Front, Int_Middle,
and Int_Back); 2) in the preheating chamber; 3) between the steel mantle and the outer insulation (Ext_Front,
Ext_Middle, and Ext_Back); solar irradiation during a typical solar experimental day at PSIs solar furnace.

3. Results and Discussion


3.1 Solar Reactor Performance
The indirect heating 10 kW solar rotary kiln prototype processed 1-5 mm limestone particles, producing
high purity lime with a degree of calcination exceeding 98% and any desired t60 reactivity ranging from
15 s (high reactivity) to 21 min (low reactivity). For details on product quality see [7].
A complete mass and heat balance of the 10 kW solar multi-tube rotary kiln prototype is found in [8]. Here,
we briefly outline the procedure for calculating the efficiency of our solar reactor:

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

out

m& CaO
Q& diss (Ta )
=
=
Q&

H CaCO3 ,diss (Ta )


M CaO
&
Qinput

input

(2)

The efficiency is defined as the enthalpy change of the calcination reaction at ambient temperature,
Q& diss (Ta ) , divided by the power input Q& input . The heat balance model accounts for non-steady-state
conditions by modifying the solar power Q& solar entering the reactor. The effective power input includes the
enthalpy change of the reactor wall materials caused by energy being stored ( Q& storage > 0 ) or released
( Q& storage < 0 ). Only part of the effective power input
Q& input = Q& solar Q& storage = Q& diss + Q& heat + Q& rerad + Q& cond + Q& other

(3)

is converted to chemical energy per unit time, Q& diss , while a substantial amount leaves the reactor via
various heat transfer paths. Q& heat accounts for the loss in the hot product's sensible heat during the cooling
phase. It may be partly recovered with a suitable heat exchanger. The value of the re-radiation heat loss
Q& rerad through the aperture is determined from the apparent emissivity of the cavity, = 1 , and the cavity
temperature. Q& cond is the conduction heat loss through the cylindrical walls of the reaction chamber. The
term Q& other includes all unaccounted heat losses such as free convection through the aperture and the
energy contained in non-recovered fine product powder leaving the reactor together with the CO2 exhaust
gas.
Figure 3 shows the relative contribution of the heat of dissociation and the various heat losses for a variety
of solar experiments that have been conducted under specific operating conditions depending on the
rotational speed of the reactor (8-18 rpm), the particle feed rate (36-136 g min-1), the limestone grain size
(2-3 mm), and the temperature (1200-1400 K).
1.0
0.9

Relative Power Fraction

0.8
0.7
Q_other
Q_rerad
Q_cond
Q_heat
Q_diss

0.6
0.5
0.4
0.3
0.2
0.1
0.0
1

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31

Experiment number

Fig. 3: Relative power fraction for solar experiments conducted under various experimental conditions. Exp.19 is
considered as best solar experiment. See text for details.

For a particular solar experiment (Exp.19, see Fig. 3) performed at 1395 K, the degree of calcination was
98.2%, and the CaO production rate was 64.2 g min-1. The solar energy to chemical energy conversion
efficiency was 34.8%, and the maximum error associated with this measurement was 15.2%. Table 1 lists
the loss terms of Eq. (3) and their uncertainties, and one can see how each of these terms contributes to
lowering the reactor efficiency. Note that Q& diss (Ta ) + Q& product ,heat = Q& diss (Tr ) + Q& CaCO3 ,heat according to the H-

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

T diagram (see e.g. [8]). In the case of Exp.19, heat was stored in the reactor wall materials ( Q& storage > 0 )
due to non-steady-state experimental conditions. Hence, the effective power input was lower than the solar
power input, and consequently eff > solar as seen from comparing the last two columns of Table 1. The
partitioning of the power input to the various heat transfer modes for the reactor is slightly changed if
transient effects are included in the calculation. For example, from the effective power input (solar power
input), 13.4% (12.4%) is conducted through the walls of the reactor, 22.0% (20.4%) is lost by re-radiation,
22.3% (20.7%) is carried away in the hot product's, and 7.5% (7.0%) leaves the reactor as "other heat
losses" (mainly convection heat losses at the aperture and un-recovered calcined powder). Obviously,
design changes that reduce any of these terms will lead to better reactor efficiency. In practice, at least part
of the sensible heat stored in the products (CaO, CO2, and non-reacted CaCO3) can be used to preheat the
limestone raw material. Similarly, the conduction heat losses can be reduced by better insulating the
reaction chamber. Assuming only 50% heat recovery, we expect the efficiency of an industrial solar lime
kiln reaching 45-55% [8].
Table 1: Heat balance for selected solar experiment (Exp.19 from Fig. 3) performed at 1395 K
with DOC>98%. Note that Q& diss (Ta ) + Q& product ,heat = Q& diss (Tr ) + Q& CaCO3 ,heat .

Typical value
Q& parameter Q& solar

Typical value
Q& parameter Q& input

10631

1830

1.077

764

2413

0.072

0.077

9867

4243

0.928

3146

269

0.296

0.319

2492

106

0.234

0.252

Q& cond
Q&

1321

953

0.124

0.134

2167

121

0.204

0.220

Q& other
Q& diss (Ta )
Q&

741

5693

0.070

0.075

3431

293

solar =0.323

eff =0.348

2206

313

0.207

0.223

Parameter

Symbol

W
Solar power input

Q& solar
Q&

Q& input
Q& diss (Tr )
Q&

Stored power
Effective power input
Heat of dissociation at Tr
Heating of CaCO3
Conduction heat losses
Re-radiation heat losses
Other heat losses
Heat of dissociation at Ta
Sensible heat of products

storage

CaCO3 ,heat

rerad

product,heat

Typical value Max. Error

In this paper, we only present results from solar experiments that have been performed with the indirect
heating 10 kW solar multi-tube rotary kiln. It is important to note, however, that corresponding
experimental results from tests with the electric heating system are in good agreement with the results
obtained in PSIs solar furnace.
3.2 Conceptual Design of an Industrial Solar Lime Plant
Figure 4 represents a conceptual design of an industrial solar lime kiln. The dimensions of the scaled-up
solar multi-tube rotary kiln are determined by the maximum feed rate of the limestone particles, which in
turn yields the minimum cross section of the absorber tubes. Arranging these tubes tightly along the
circumference of the absorber cavity defines the minimum diameter of the cavity. The length of the
absorber tubes within the cavity is determined by the retention time of the limestone particles. A secondary
concentrator is required to enhance the solar radiation entering the kiln, since the concentrated solar flux of
large solar concentrating systems is typically limited to a maximum of about 1200 suns (1 sun = 1 kW m-2),
which is not sufficient to reach calcination temperatures in the range 1200-1600 K. Typical dimensions of
the absorber tubes, the cavity, and the non-imaging compound parabolic concentrator (CPC, [9]) are shown
in Table 2 for rotary kilns of 0.5 MWth, 3 MWth, and 20 MWth solar power input, respectively.
Peripheral components of the solar lime kiln comprise state-of-the-art feeding, discharging, and heat
recovery systems. For fossil or hybrid mode of operation, both oil and gas burners may be conveniently

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

inserted into the cavity of the kiln. The implementation of these conventional system components is
straightforward and therefore is not discussed here.

Fig. 4: Conceptual design of a megawatt-scale indirect heating rotary kiln showing the well-insulated kiln including
CPC (left) and the arrangement of the absorber tubes (right). For typical dimensions see Table 2.

Table 2: Estimated dimensions of rotary kiln and CPC for 0.5 MWth, 3 MWth, and
20 MWth solar power input. Assumptions: Reactor efficiency = 0.5 ; solar
~
concentration ratio C = 2000 ; heliostat field rim angle rim = 39 ; particle
residence time t res = 300 s .
Parameter

Symbol

Unit

Solar power input


0.5 MWth

3 MWth

20 MWth

t/h

0.5

3.0

20.0

Feed rate (peak)

m& CaCO 3

Number of tubes

nt

35

67

105

Tube diameter

dt

0.06

0.07

0.09

Tube length

lt

0.60

1.40

3.60

Cavity diameter

d cav

1.17

2.55

5.00

CPC entrance diameter

d in

0.91

2.22

5.73

CPC exit diameter

d out

0.57

1.40

3.60

CPC length

lCPC

0.91

2.23

5.76

Principally, two different types of solar concentrating systems are conceivable for the solar lime plant: (1)
central receiver tower top (TT) system consisting of a field of heliostats reflecting the sunlight to the lime
kiln mounted on top of a tower, and (2) central receiver tower reflector or beam down (BD) system,
where the solar lime kiln is sitting on the ground. Although BD systems offer advantages for materials
handling, the indirect heating rotary kiln seems not well adapted to concentrated solar radiation impinging
nearly perpendicularly on the absorber tubes. Especially, convective heat losses due to buoyancy effects are
expected to be much larger than for applications where the solar radiation enters axially through the
aperture into the cavity of the kiln. Furthermore, BD systems tend to be more expensive than TT systems
[10]. Thus, for both technical and economic reasons, the BD option has not been pursued further.
Conventional TT systems with planar heliostat field are well proven technology and no more fundamental
planning is required [11]. However, one disadvantage of such large solar plants is the need for extended

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

land area to minimize blocking and shading of the heliostats. This results in large distances from the
heliostats to the kiln, requiring extremely high quality mirrors and thus affecting the cost of the solar lime
plant. Minor technical concerns are associated with the handling of the material on the rather tall tower and
the CPC being off-axis with the kiln in order to accommodate for the rim angle of the heliostat field.
For solar lime plants smaller than about 1 MWth thermal input, an alternative arrangement of the heliostat
field and the rotary kiln is proposed (Fig. 5). Here, the heliostat field is built on a natural or artificial hill,
and the heliostats are arranged in lines on a number of terraces. The advantages of such a configuration vis-vis conventional TT systems are obvious: 1) Less land area is needed; 2) Less blocking and shadowing of
heliostats occurs; 3) Distances between heliostats and kiln are smaller and more uniform; 4) The heliostat
field may be arranged such that the CPC is on-axis with the kiln and illuminated more symmetrically; 5)
Materials handling is easier because the kiln is mounted close to ground level, and no high tower is thus
needed. Major disadvantages are due to the additional civil engineering work needed for preparing the
trough-shaped slope and the terraces for the heliostats. Furthermore, installation and maintenance of the
heliostats is rendered more difficult.

Fig. 5: Solar lime plant (<1 MWth thermal input): Alternative natural or artificial arrangement of the lime kiln on a hill
and the heliostat field on a south-facing slope. Indicated are the extreme positions of the heliostats and the sunrays at
both winter and summer solstice for typical Mediterranean regions.

In summary, the TT system is a viable technical concept for an industrial solar lime plant. For solar lime
plants with power input less than about 1 MWth, it may be worth considering the alternative plant
configuration with heliostats on a slope and the lime kiln on a hill. For a 20 MWth industrial solar
calcination plant, cost of solar produced lime are estimated in the range 131-158 $/t, about 2-3 times the
current selling price of conventional lime [10]. The solar production of high purity lime might be
competitive with conventional fossil fuel based calcination processes at current fuel prices.
4. Conclusions and Outlook

We have developed the solar chemical reactor technology that aims at reducing CO2 emissions in the lime
and cement industry by replacing fossil fuels with solar energy. Concentrated sunlight is used as process
heat for driving the calcination reaction: CaCO3 CaO + CO2. Our work was primarily focused on
developing a scaleable solar calcination reactor for efficiently processing limestone (CaCO3) particles and
producing high purity quicklime (CaO) for applications in specific market sectors of the chemical and
pharmaceutical industry.
The question of technical feasibility was addressed by first designing, constructing, and then experimentally
evaluating the performance of an indirect heating 10 kW solar multi-tube rotary reactor prototype for
effecting the calcination reaction at temperatures up to 1400 K. Experimental results obtained at the small
scale in PSIs high-flux solar furnace confirm that limestone particles in the range of 1-5 mm can be
efficiently calcined with concentrated sunlight and that a high degree of chemical conversion can be
achieved. The solar produced high purity lime is not contaminated with combustion by-products and meets
the highest industrial standards. The reactors efficiency, defined as the enthalpy of the calcination reaction

Paper submitted to 12th SolarPACES International Symposium, October 6-8, 2004, Oaxaca, Mexico

at room temperature (3184 kJ kg-1) divided by the solar energy input, reached 30-35% for solar flux inputs
of about 2000 kW m-2 and for quicklime production rates up to about 4 kg h-1.
We are convinced that our solar lime technology has the potential of meeting an industrial standard for
reactor performance. We expect the efficiency of an industrial reactor to be higher than that of the smallscale experimental version if conduction losses were reduced and sensible heat were recovered for
preheating the limestone particles. Such a reactor with a thermal efficiency near 45-55% that also produces
high quality quicklime would demonstrate that an Industrial Solar Lime Plant could be an economically
viable path for reducing CO2 emissions in specific market sectors of the lime industry.
Acknowledgements

This work has been funded in part by the Swiss Federal Office of Energy (BFE) and performed at the Solar
Furnace, Paul Scherrer Institut, Villigen, Switzerland. We thank R. Palumbo, A. Steinfeld, and C. Wieckert
for fruitful discussions. We also thank M. Brack for help with calibrating the flux measurement system and
P. Hberling for technical assistance.
References

[1] Oates J.A.H. (1998), Lime and Limestone: chemistry and technology, production and uses,
Weinheim: Wiley-VCH Verlag GmbH, Germany.
[2] Boynton R.S. (1980), Chemistry and Technology of Lime and Limestone, New York: John Wiley &
Sons Inc., USA.
[3] World Business Council for Sustainable Development (WBCSD), (2002), The cement sustainable
initiative our agenda for action, http://www.wbcsd.org.
[4] Bonaldi E., Cella G.M., Lipinski W., Palumbo R., Steinfeld A., Wuillemin D., Meier A. (2002) CO2
Mitigation in the Lime Industry: Replacing Fossil Fuels with Concentrated Solar Energy, In Proc. 10th
International Lime Association Congress, Washington D.C., USA, May 7-10, 2002.
[5] Meier, A., Bonaldi E., Cella G.M., Lipinski W., Palumbo R., Wuillemin D. (2004), Design and
Experimental Investigation of a Horizontal Rotary Reactor for the Solar Thermal Production of Lime,
Energy The Int. J. 29 (5-6), 811-821.
[6] Meier A., Bonaldi E., Cella G.M., Lipinski W., Wuillemin D. (2003), Reactor for indirect utilization
of external radiation heat for thermal or thermochemical material processes, European Patent Appl.,
EP 03 010 448.3, May 9, 2003.
[7] Meier A., Cella G.M., (2004), Harnessing the Power of the Sun, World Cement Magazine. Volume
35, Number 8, August 2004.
[8] Meier A., Bonaldi E., Cella G.M., Lipinski W. (2004), Multi-Tube Rotary Kiln for the Industrial Solar
Production of Lime, J. Solar Energy Eng. Submitted for publication 2004.
[9] Welford W.T., Winston R. (1989), High collection nonimaging optics, San Diego, CA: Academic
Press.
[10] Meier A., Gremaud N., Steinfeld A. (2004), Economic Evaluation of the Industrial Solar Production
of Lime, Energy Conv. & Mgmt. In press.
[11] Romero M., Buck R., Pacheco J.E. (2002), An Update on Solar Central Receiver Systems, Projects,
and Technologies, J. Solar Energy Eng. 124, 98-108.

Das könnte Ihnen auch gefallen