Sei sulla pagina 1di 9

International Journal of Hydrogen Energy 27 (2002) 611 619

www.elsevier.com/locate/ijhydene

Solar hydrogen production via a two-step water-splitting


thermochemical cycle based on Zn=ZnO redox reactions
A. Steinfelda; b;
a Department

of Mechanical and Process Engineering, Institute of Energy Technology, ETH - Swiss Federal Institute of Technology,
ETH-Zentrum, CH-8092 Zurich, Switzerland
b Solar Process Technology, Paul Scherrer Institute, CH-5232 Villigen, Switzerland

Abstract
The production of hydrogen from water using solar energy via a two-step thermochemical cycle is considered. The -rst,
endothermic step is the thermal dissociation of ZnO(s) into Zn(g) and O2 at 2300 K using concentrated solar energy as the
source of process heat. The second, non-solar, exothermic step is the hydrolysis of Zn(l) at 700 K to form H2 and ZnO(s);
the latter separates naturally and is recycled to the -rst step. Hydrogen and oxygen are derived in di2erent steps, thereby
eliminating the need for high-temperature gas separation. A 2nd-law analysis performed on the closed cyclic process indicates

a maximum exergy conversion e3ciency of 29% (ratio of 5G298 K |H2 +0:5O2 H2 O for the H2 produced to the solar power input),
when using a solar cavity-receiver operated at 2300 K and subjected to a solar 9ux concentration ratio of 5000. The major
sources of irreversibility are associated with the re-radiation losses from the solar reactor and the quenching of Zn(g) and O2
to avoid their recombination. An economic assessment for a large-scale chemical plant, having a solar thermal power input
into the solar reactor of 90 MW and a hydrogen production output from the hydrolyser of 61 million-kWh=yr, indicates that
the cost of solar hydrogen ranges between 0.13 and 0.15$=kWh (based on its low heating value and a heliostat -eld cost at
100 150$=m2 ) and, thus, might be competitive vis-?a-vis other renewables-based routes such as electrolysis of water using
solar-generated electricity. The economic feasibility of the proposed solar process is strongly dependent on the development
of an e2ective Zn=O2 separation technique (either by quench or by in situ electrolytic separation) that eliminates the need
for an inert gas. The chemical aspects of the reactions involved and the present status of the pertinent chemical reactor
technology are summarized. ? 2002 International Association for Hydrogen Energy. Published by Elsevier Science Ltd. All
rights reserved.

1. Introduction
Single-step thermal dissociation of water, also known as
water thermolysis, although conceptually simple, has been
impeded by the need of a high-temperature heat source at
above 2500 K for achieving a reasonable degree of dissociation, and by the need of an e2ective technique for
Corresponding author. Department of Mechanical and
Process Engineering, ETH - Swiss Federal Institute of
Technology, ETH-Zentrum, CH-8092 Zurich, Switzerland.
Tel.: 41-56-310-3124; fax: 41-56-310-3160.
E-mail address: aldo.steinfeld@eth.ch (A. Steinfeld).

separating H2 and O2 to avoid recombination or end up with


an explosive mixture. Among the ideas proposed for separating H2 from the products are e2usion separation [1,2], and
electrolytic separation [3,4]. Water-splitting thermochemical cycles bypass the H2 =O2 separation problem and further
allow operating at relatively moderate upper temperatures.
Previous studies performed on H2 O-splitting thermochemical cycles were mostly characterized by the use of process
heat at below about 1300 K, available from nuclear and other
thermal sources [5]. These cycles required multiple steps
(more than two) and were su2ering from inherent ine3ciencies associated with heat transfer and product separation at
each step. A status review on multi-step cycles was given

0360-3199/02/$ 22.00 ? 2002 International Association for Hydrogen Energy. Published by Elsevier Science Ltd. All rights reserved.
PII: S 0 3 6 0 - 3 1 9 9 ( 0 1 ) 0 0 1 7 7 - X

612

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

Nomenclature
C
LHV
I
Irr quench
Irr reactor
Irr hydrolyser
n
QFC
Qhydrolyser
Qquench
Qreactor; net
Qreradiation
Qsolar

solar 9ux concentration ratio, dimensionless


low heating value of H2 , 241 kJ mol1
normal beam insolation, kW m2
irreversibility associated with quenching,
kW K 1
irreversibility associated with solar reactor,
kW K 1
irreversibility associated with hydrolyser,
kW K 1
molar 9ow-rate, mol s1
rate of heat rejected to the surroundings by
fuel cell, kW
rate of heat rejected to the surroundings by
hydrolyser, kW
rate of heat rejected to the surroundings by
quenching, kW
net power absorbed by solar reactor, kW
power re-radiated by solar reactor, kW
solar power input to solar reactor, kW

in [5,6] and includes a three-step cycle based on the solar


decomposition of H2 SO4 at 1140 K and a four-step cycle
based on the solar hydrolysis of CaBr 2 and FeBr 2 at 1020
and 870 K, respectively.
In recent years signi-cant progress has been accomplished in the development of optical systems for large-scale
collection and concentration of solar energy, using solar
tower and tower-re9ector technology at power levels of
several megawatts. These concentrating systems are capable of achieving power 9ux intensities equivalent to solar
concentration ratios 1 of 5000 suns and higher by applying
non-imaging secondary concentrators (usually compound
parabolic concentrators, also known as CPCs [7]) in tandem
with the primary focusing heliostat -eld [8,9]. Such high
radiation 9uxes correspond to stagnation temperatures 2
exceeding 3000 K and allow the conversion of concentrated solar radiation to thermal reservoirs at 2000 K and
above. Thus, the door was opened for the more e3cient
two-step thermochemical cycles for splitting water using
solar energy.
1 The solar 9ux concentration ratio C is de-ned as the ratio of
the solar 9ux intensity achieved after concentration to the incident
beam normal insolation. It is a dimensionless number, sometimes
reported in units of suns.
2 The stagnation temperature is the highest temperature an ideal
blackbody receiver is capable of achieving when energy is being
re-radiated as fast as it is absorbed. It is given by (CI=)0:25 , where
I is the normal beam insolation, and  is the StefanBoltzmann
constant.

T
WFC
5G
5H
5S
absorption
Carnot
exergy
FC
optics
solar

thermal

temperature, K
rate of work output by fuel cell, kW
Gibbs free energy change, kJ mol1
enthalpy change, kJ mol1
entropy change, kJ mol1 K 1
solar energy absorption e3ciency
e3ciency of a Carnot heat engine
exergy e3ciency of the cycle
e3ciency of H2 =O2 fuel cell
optical e3ciency of the solar
concentrating system
thermal e3ciency of solar concentrating system + solar reactor
StefanBoltzmann
constant,
5:6705108 W m2 K 4

Subscripts
e
th

electric
thermal

Several two-step water-splitting cycles based on metal


oxides redox reactions have been proposed [10, and the literature cited therein]. The -rst, endothermic step is the solar thermal dissociation of metal oxide to the metal or the
lower-valence metal oxide. The second, non-solar, exothermic step is the hydrolysis of the metal at moderate temperatures to form hydrogen and the corresponding metal oxide.
The net reaction is H2 O=H2 +0:5O2 . Hydrogen and oxygen
are formed in di2erent steps, thereby eliminating the need
for high-temperature gas separation. These cycles have been
thermodynamically examined and tested in solar reactors for
ZnO=Zn and Fe3 O4 =FeO redox pairs [10 15]. Other redox
pairs, such as TiO2 =TiOx , Mn3 O4 =MnO, and Co3 O4 =CoO
have also been considered, but the yield of H2 has been too
low to be of any practical interest. Partial substitution of iron
in Fe3 O4 by other metals forms mixed metal oxides of the
type (Fe1x Mx )3 O4 that may be reducible at lower temperatures than those required for the reduction of Fe3 O4 , while
the reduced phase (Fe1x Mx )1y O remains capable of splitting water [16,17].
Of special interest is the solar thermochemical cycle based
on ZnO=Zn redox reactions, represented by
1st step (solar):

ZnO = Zn + 0:5O2 ;

2nd step (non-solar):

Zn + H2 O = ZnO + H2 :

(1)
(2)

Several chemical aspects of the thermal dissociation of ZnO


[Eq. (1)] have been previously investigated [13,1824]. At

2235 K, 5G = 0. Thermogravimetric studies reported an


apparent activation energy in the range 310 350 kJ=mol

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

[18,19]. The reaction rate for directly irradiated ZnO pellets,


assumed to obey an Arrhenius-type law, was derived in [20].
The product gases needed to be either quenched or separated
at high temperatures to prevent their recombination. The
condensation of zinc vapor in the presence of O2 was
studied by fractional crystallization in a temperature-gradient
tube furnace [21]. It was found that the oxidation of Zn is
a heterogeneous process and, in the absence of nucleation
sites, Zn(g) and O2 can coexist in a meta-stable state. In
particular, the quench e3ciency is sensitive to the dilution
ratio of Zn(g) in an inert gas 9ow and to the temperature of the surface on which the products are quenched.
Alternatively, electrolytic methods have been proposed for
in situ separation of Zn(g) and O2 at high temperatures, and
experimentally demonstrated to work in small-scale reactors
[4,22,23]. Various exploratory tests on the dissociation of
ZnO were carried out at the solar furnaces and solar simulators of PSI=ETH (Switzerland) and CNRS-Odeillo=Nancy
(France) for gaining kinetic information and acquiring experience with the reduction of ZnO using concentrated solar
energy [13,20,24 26]. Based on these previous studies and
on the constraints imposed by the chemistry of the decomposition reaction, a solar chemical reactor was designed
and a 10 kW prototype was fabricated and recently tested
[27]. It features a windowed rotating cavity-receiver lined
with ZnO particles that are held by centrifugal force. With
this arrangement, ZnO is directly exposed to high-9ux solar
irradiation and serves simultaneously the functions of
radiant absorber, thermal insulator, and chemical reactant.
The reactor development work is in progress and currently
focused on experimental evaluation and on -nding optimal
operating conditions for maximum exergy e3ciency [27].
As far as the second step [Eq. (2)] is concerned, laboratory studies on the kinetics and preliminary tests with
a novel concept of a hydrolyser reactor indicate that the
water-splitting reaction proceeds exothermally at reasonable
rates when steam is bubbled through molten zinc at above
about 700 K [28,29]. In principle, the heat liberated could
be used in an auto-thermal type of hydrolyser to melt zinc
and produce steam. Alternatively, if the H2 production plant
is realized next to the solar plant, molten zinc could be
withdrawn from the quencher at 700 K (or higher) and fed
directly to the hydrolyser. This might be accomplished by
implementing a splash-condenser at the exit of the solar
reactor: a quench technique widely used in the industrial
carbothermic reduction of ZnO by the imperial smelting process for simultaneously avoiding Zn(g) re-oxidation with
CO2 and obtaining molten zinc [30]. Yet, the e2ectiveness
of the splash-condenser to avoid Zn(g) re-oxidation with
O2 needs demonstration. The technical feasibility of the
solar reactor and hydrolyser at an industrial scale also needs
demonstration.
This paper presents a Second-law analysis for establishing the ideal maximum exergy e3ciency of the proposed water-splitting cycle and for identifying the major
sources of irreversibility. It further presents an economic

613

assessment of the costs of solar H2 and of electricity generated by solar H2 =O2 fuel cell for a large-scale solar chemical
plant. Both analyses require a de-nition of the material and
energy 9ows.

2. Exergy analysis
A 2nd-law (exergy) analysis of an ideal ZnO=Zn cyclic
process, in which Zn and O2 recombine in a reversible ideal
fuel cell, is found in [31] along with all the pertinent energy
balance equations for calculating the maximum exergy e3ciency. Here, the analysis is extended to include the hydrolysis of Zn for producing H2 and a H2 =O2 fuel cell to close
the cycle.
A model 9ow diagram for the proposed two-step
solar thermochemical cycle is shown schematically in Fig. 1.
It uses a solar reactor, a quenching device, a hydrolyser
reactor, and a H2 =O2 fuel cell. The complete process is
carried out at a constant pressure of 1 bar. In practice, pressure drops will occur throughout the system and pumping
work will be required. The solar reactor is assumed to
be a cavity-receiver operating at 2300 K. Its solar energy
absorption e3ciency, absorption , is de-ned as the net rate
at which energy is being absorbed, Qreactor; net , divided by
the solar power input through the solar receivers aperture,
Qsolar . For a perfectly insulated blackbody cavity-receiver
(no convection or conduction heat losses; cavitys e2ective
absorptivity and emissivity approaching 1), it is given by
 4
Qreactor; net
T
=1
absorption =
;
(3)
Qsolar
IC
where I is the normal beam insolation (taken in this study
equal to 1 kW=m2 ), C is the 9ux concentration ratio of the
solar concentrating system, T is the nominal reactor temperature (2300 K), and  is the StefanBoltzmann constant.
absorption expresses the capability of a solar reactor to absorb
incoming concentrated solar energy but does not account
for the optical losses incurred in collecting and concentrating solar energy. The optical losses will be discussed later
in the economic analysis.
The reactant ZnO(s) enters the solar reactor at 298 K
and is further heated to the reactor temperature at 2300 K.
The molar feed rate of ZnO to the reactor, n,
is set to
1 mol=s, and is equal to that of H2 O fed to the hydrolyser reactor. Chemical equilibrium is assumed inside
the reactor. The net power absorbed in the solar reactor
should match the enthalpy change per unit time of the
reaction,
Qreactor; net = n5H

|ZnO(s)

at 298 KZn(g)+0:5O2 at 2300 K :

(4)

Irreversibilities in the solar reactor arise from the


non-reversible chemical transformation and re-radiation
losses to the surroundings at 298 K, given by

614

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

T = 2300 K

Qsolar
Concentrated
Solar Energy
ZnO
@ 298 K

Qreradiation

Zn + O2
@ 2300 K

WF.C.
QF.C.

O2 @ 298 K
Fuel
Cell

Qquench

Quench
Zn @ 298 K

H2O

Qhydrolyser

Hydrolyser
H2

Zn

Fig. 1. Model 9ow diagram of the water-splitting solar thermochemical cycle used for the exergy analysis.


Irr reactor =

Qsolar
2300 K


+

+ (n5S|

ZnO(s)

Qreradiation
298 K

by
Qhydrolyser = n5H

|Zn+H2 O

at 298 KZn(g)+0:5O2 at 2300 K );

(5)

where Qreradiation denotes the radiation heat loss by the reactor, Qreradiation = (1 absorption )Qsolar . Products Zn(g) and O2
exit the solar reactor at 2300 K and are cooled rapidly to
298 K. It is assumed that the chemical composition of the
products remains unchanged upon cooling in the quencher.
Since the quench step is required for avoiding recombination of products, no heat exchanger is used for recovering
their sensible and latent heat. Thus, the amount of power
lost during quenching is
Qquench = n5H

|Zn(g)+0:5O2

at 2300 KZn(s)+0:5O2 at

298 K : (6)

at 2300 KZn(s)+0:5O2 at 298 K ):

(7)
After quenching, the products separate naturally (without
expending work) into gaseous O2 and condensed phase zinc.
Zinc is sent to the hydrolyser reactor to react exothermally
with water and form hydrogen, according to Eq. (2). The
heat liberated is assumed lost to the surroundings, as given

(8)

Thus, the irreversibility associated with the hydrolyser is:




Qhydrolyser
Irr hydrolyser =
298 K
+ (n5S|

Zn+H2 O

at 298 KZnO+H2 at 298 K ):

(9)

The cycle is closed by introducing an H2 =O2 fuel cell, in


which the products recombine to form H2 O and thereby
generate electrical power. WFC and QFC are the work output
and heat rejected, respectively, given by
WFC = FC n5G|

H2 +0:5O2

at 298 KH2 O at 298 K ;

QFC = TL n5S|
H2 +0:5O2

and the irreversibility associated with it is




Qquench
Irr quench =
298 K
+ (n5S|

Zn(g)+0:5O2

at 298 KZnO+H2 at 298 K :

at 298 KH2 O at 298 K ;

(10)
(11)

where FC is the e3ciency of the H2 =O2 fuel cell, assumed


100% when based on the maximum work output of H2
(also known as its exergy), and 70% when based on the
conversion to electricity with current PEM fuel cells. Finally, the exergy e3ciency is de-ned as the portion of solar
energy converted into chemical energy of the reaction
products (given by the Gibbs free energy change of product
oxidation), and is thus calculated as the ratio of the fuel
cells work output to the solar power input,
exergy =

WFC
:
Qsolar

(12)

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619


Table 1
Exergy analysis of the two-step water-splitting thermochemical
cycle using the process modeling depicted in Fig. 1. The ZnO and
H2 O molar rates are set to 1 mol=s
C

5000

10,000

Qsolar
Qreradiation
Qreactor; net
Irr reactor
Qquench
Irr quench
Qhydrolyser
Irr hydrolyser
QFC
WFC
absorption
exergy
exergy ( FC = 0:7)

815 kW
258 kW

662 kW
105 kW

68%
29%
20%

557 kW
0:81 kW K 1
209 kW
0:52 kW K 1
64 kW
0:27 kW K 1
49 kW
237 kW

84%
36%
25%

The veri-cation of this thermodynamic analysis is shown in


the appendix.
Table 1 gives the results of the calculation for two values
of the solar concentration ratio: 5000 and 10,000. The exergy
e3ciency reaches 29% and 36%, respectively. As expected,
the exergy e3ciency increases with the solar concentration
ratio as a result of being able to use a smaller solar receivers
aperture to intercept the same amount of solar power, and,
consequently, reduce re-radiation losses and increase the
absorption e3ciency [see Eq. (3)]. Note that, for a given
concentration ratio, smaller apertures intercept a reduced
fraction of the incoming solar power. Thus, the optimum
aperture size of the solar receiver becomes a compromise
between maximizing solar radiation capture and minimizing
re-radiation losses [32].
The major sources of irreversibility are associated with the
re-radiation losses from the solar reactor and the quenching
of products: 32% and 26% of the input solar power is lost by
re-radiation and by quenching, respectively. Besides using
smaller apertures to match higher solar concentration ratios,
re-radiation losses can also be diminished to some extent by
implementing selective windows, with high transmissivity
in the solar spectrum around 0.5 m where the solar irradiation peaks, and high re9ectivity in the infrared range around
1:26 m where the Planks spectral emissive power for a
2300 K blackbody peaks. 3 Quenching losses can be somewhat reduced if the quench takes e2ect from 2300 K to only
700 K (and not to 298 K), and molten zinc is further sent to
an on-site hydrolyser. Still, the largest heat loss occurs
during the condensation of Zn(g): 116 kJ=mol or 55% of
the total heat rejected during quench. This analysis assumes
a perfect quench, i.e. no recombination of Zn and O2 .
3 From Wiens displacement law: 
max T =2897:8 m K, where
max is the wavelength for maximum spectral hemispherical emissive power of a blackbody at temperature T .

615

It further assumes no energy expenditure associated with


the use of a quenching inert gas. In practice, some product
recombination will occur and supplemental work will be
needed to separate and recirculate the inert gas. A good
maximum e3ciency can be obtained as long as the inert gas
to Zn(g) molar ratio at the decomposition temperature is
kept below 10 [13]. Alternatively, electrolytic methods for
in situ separation of Zn(g) and O2 eliminate the quenching
step and further enable the recovery of some portion of the
sensible and latent heat of the products.
The exergy e3ciency is an important criterion for judging
the relative industrial potential of the cycle. The higher the
exergy e3ciency, the lower is the required solar collection
area for producing a given amount of solar H2 . Hence, the
lower are the costs of the solar concentrating system which,
as it will be shown in the analysis that follows, usually
correspond to about half of the total investments for the
entire solar chemical plant. Thus, high exergy e3ciency
implies favorable economic competitiveness.
3. Economic assessment
The cost analysis is carried out using the general procedure developed in [33], applied previously in the solar combined ZnO-reduction and CH4 -reforming [34]. The baseline
operating conditions and enthalpy 9ows are the ones indicated in Table 1. The solar concentrating plant is assumed to
be a solar tower system with a Cassegrain optical con-guration of the type being developed at the Weizmann Institute
of Sciences [8,9]. This innovative beam-down concentrating system uses a -eld of heliostats (two-axis tracking
parabolic mirrors) to focus the sunrays onto a
hyperboloidal re9ector positioned at the top of the tower,
which further re-directs the concentrated sunlight down to
a CPC at the ground level. With this arrangement, concentrated solar radiation emerging out of the CPC (or an
array of CPCs) can enter the solar reactor (or an array of
solar reactors) located on the ground level, eliminating the
need for massive and expensive tower, piping, and frequent
personnel access to the tower top. The heliostat -eld is the
largest single cost item in the solar plant. The assumed
baseline cost of 150$=m2 is consistent with current estimates of large-scale production of silvered glass heliostats
[35,36]. The amount of area required is dependent on the
e3ciency of delivering solar heat to the receiver-reactor.
The annual solar thermal e3ciency, solar thermal , is de-ned
as the fraction of annual solar radiation used for process
heat. It is calculated, on a yearly basis, as the ratio of the
enthalpy change of the reaction to the solar beam radiation incident over the heliostat area. Thus, solar thermal is
the product of the solar receivers absorption e3ciency,
absorption , and the optical e3ciency of the solar concentrating system, optics . The latter is assumed to be 58%,
and is based on ray-tracing calculations that account for
re9ectivity, shading, alignment, and spillage losses due to

616

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

geometrical and tracking imperfections [37]. The costs for


the tower, tower re9ector, and CPC were derived from
[37], designed for a 34 MWe solar-driven combined cycle
power plant. The cost of the receiver-reactor was based on
judgment considering the earlier data [3739]. Costs associated with balance of plant were based mainly on [38] which
displays detailed breakdowns of a variety of chemical systems. It includes site development, infrastructure, piping,
instrumentation and control, security provisions, contingency, etc. The O&M cost is taken to be 2% of the capital
cost. A -xed charge rate of 15% per annum is assumed, but
early plants would probably have a higher -xed charge rate
because of the perceived high risk. The solar plant size is
selected to be one that delivers 90 MW of concentrated
solar power to the solar reactor, Qsolar . The only reason
for this selection is that it allows the cost for the tower,
tower re9ector, and CPC to be directly extracted from
[37] without inter=extrapolations. An important implicit
assumption is that the solar tower technology becomes cost
e2ective at large scales, but the question of the optimal
size of the solar plant is not addressed in this analysis.
Other baseline parameters are an annual beam irradiation
of 2300 kWhth =m2 =yr and 2300 equivalent full power hours
per year (obviously intermittently). For all cases, the H2 produced is assumed to have an energy content of 241 kJ=mol,
which corresponds to its low heating value (LHV).
Government subsidies or any credit for CO2 mitigation
have been excluded from consideration. Credit for O2 sale
(O2 derived from the solar reactor) has also been excluded.
Also not accounted for is any cost incurred from storage and
transport of reactants and products, and transporting H2 to
the consumer site. Since the hydrolyser does not necessarily
need to be located next to the solar plant, the transportation
and storage costs for H2 can be minimized or even completely eliminated if the hydrolyser is located next to the
consumer site, but the transportation and storage costs for
Zn and ZnO, which are comparable to those for coal, will
then have to be added. The recycling of a quenching inert
gas is not included explicitly in the cost breakdown, but its
implications are discussed.
Table 2 shows the cost breakdown of the solar chemical
plant and the unit cost of hydrogen, in US$=kWh. Included
is also the cost of electricity generated using 70% e3cient
PEM fuel cells fueled by solar H2 . Similar to Table 1, calculations were carried out for two values of the solar concentration ratio: 5000 and 10,000. As expected, the heliostat
-eld causes the largest single cost item and is responsible
for 44% of the total investment costs for the entire chemical
plant. In contrast, the cost of the solar reactor represents
only 13%. However, for a -xed product throughput, the
solar reactors e3ciency dictates the size of the heliostat
-eld (or, vice versa, for a -xed size of the heliostat -eld,
the solar reactors e3ciency dictates the product throughput). Thus, reaching high solar reactor e3ciencies and
reducing the cost of the heliostats per unit area will have a
signi-cant impact on reducing the unit cost of H2 . For

example, a reduction of the H2 cost by 13% is possible for


a heliostat -eld at 100$=m2 . The same magnitude of cost
reduction can also be obtained by doubling the concentration ratio to 10,000 and, consequently, increasing the solar
reactors absorption e3ciency by 23%. Note that doubling
the concentration would require higher precision optics of
the solar tower re9ector and CPC, estimated to incur double
costs. It may also require more accurate heliostats arranged
to have more ground coverage, resulting in higher blocking
and shadowing and, therefore, higher costs per m2 of heliostat -eld. Thus, the economical optimization of the solar 9ux
concentration is a complex problem, not solved in the present
study.
The speci-c cost of solar H2 is estimated to be in the
range 0.14 0:15$=kWh (based on its LHV), assuming
baseline parameters and a solar concentration ratio varying
between 5000 and 10,000. A sensitivity analysis for
C = 5000 revealed that the cost of H2 varies in the range
0.11 0.17$=kWh when the heliostat -eld cost varies between 50 and 200$=m2 . This hydrogen cost has to be compared to that of hydrogen produced from water using other
renewable energy based routes, e.g., solar, wind, biomass,
geothermal, hydro, etc. The reference route is generally
taken to be the production of hydrogen by electrolysis of
water using solar-generated electricity. For solar thermal
troughs systems (e.g. SEGS plant in California [40]), currently generating electricity at 0:12$=kWhe , the reference
cost of H2 is cited as $0.20=kWh [40,41]. Thus, the proposed
thermochemical route has the potential of becoming economically competitive. For wind electricity, currently at about
0:06$=kWhe , the cost of H2 by H2 O-electrolysis is estimated to be 0.17$=kWh at present, and expected to come
down to 0.10$=kWh by the year 2005, as both the wind and
electrolysis technologies mature [41]. It will be di3cult
for any solar technology to compete with the cost of H2
from wind electricity unless the heliostat -eld cost drops
under 100$=m2 and the annual solar thermal e3ciency
exceeds 40%. Obviously, for locations with rich solar insolation but poor wind resources, solar is going to be more
competitive.
The cost of separating an inert quenching gas from the
gaseous products exiting the solar reactor is not included in
the cost breakdown of Table 2 because of the uncertainty
regarding the need of such a quenching gas. However, if
needed, it could play a decisive factor on the economics of
the entire solar process. If N2 were to be used as the
inert quenching gas by an N2 =Zn dilution ratio of 10, the
annual N2 requirement would amount to about 9106 kmol.
Assuming current costs of N2 production with membrane
separation techniques at 0.06 0:08$=Nm3 (by an electricity cost of 0.06$=kWhe ), the annual capital costs would
be increased by 12 M$, resulting in about doubling the
speci-c cost of solar H2 [42]. Thus, the economic feasibility of the proposed process is strongly dependent on the
development of an Zn=O2 quench technology (or, alternatively, in situ Zn=O2 electrolytic separation technology)

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

617

Table 2
Estimated costs of solar H2 and of solar electricity generated by H2 =O2 fuel cells fueled with solar H2 . The LHV for H2 is assumed
Plant size, energy, and mass 5ows
Solar plant size (solar power input to solar reactor) (MWth )
Solar input on heliostat -eld=year (MWhth =yr)
Heliostat area (m2 )
Design ZnO=H2 O feed (kg mol=h)
Zinc metal production (tons=yr)
Hydrogen production (million-kWh=yr)
Electricity production (million-kWhe =yr)

90
356,896
155,172
398
59,433
61
42

90
356,896
155,172
398
73,169
75
52

E6ciencies
Solar concentration ratio, C
Optical e3ciency of solar concentrating system, optics
Solar reactors absorption e3ciency, absorption
Cycle exergy e3ciency, exergy
Fuel cell e3ciency, FC
Annual solar thermal e3ciency, solar thermal

5000
58%
68%
29%
70%
40%

10,000
58%
84%
36%
70%
49%

Capital cost
Heliostat -eld (M$, assuming $150=m2 )
Tower (M$)
Tower re9ector and CPCs (M$)
Solar receiver-reactor + periphery (M$)
Quencher (M$)
Hydrolyser (M$)
Balance of plant, indirects, contingency (M$)
PEM fuel cells (M$, assuming $1500=kWe installed)
Total for solar H2 (M$)
Total for solar electricity (M$)
Speci-c installation cost for solar H2 ($=kW installed)
Speci-c installation cost for solar electricity ($=kWe installed)

23.28
3.60
5.30
7.00
3.00
4.00
8.90
27.48
55.08
82.56
2069
4506

23.28
3.60
10.60
7.00
3.00
4.00
8.90
33.83
60.38
94.21
1843
4177

Annual cost
Annual -xed charge rate (M$)
Capital cost for solar H2 (M$)
Capital cost for solar electricity (M$)
O& M cost for solar H2 (M$)
O& M cost for solar electricity (M$)
Total cost for solar H2 (M$)
Total cost for solar electricity (M$)

15%
8.26
12.38
1.10
1.65
9.36
14.03

15%
9.06
14.13
1.21
1.88
10.26
16.02

0.15
0.33

0.14
0.31

Speci7c cost
Unit cost of solar H2 ($=kWh)
Unit cost of solar electricity from fuel cells ($=kWhe )

that practically eliminates the need for an inert quenching


gas.
The cheapest non-renewable H2 is obtained at the moment via the catalytic steam-reforming of natural gas (with
process heat supplied by the combustion of natural gas),
at about 0.03 0.04$=kWh, assuming feedstock cost in the
range 10 12$=MWh and excluding any externalities such as
the cost of CO2 mitigation and pollution abatement [43,44].
The external costs may be assessed with the help of a life cycle analysis (LCA) for evaluating the environmental burdens
associated with the process by quantifying energy and

materials used and wastes released during the entire life


cycle. A recent LCA for the conventional syngas production indicates greenhouse gas emissions 4 of over
4 The three most relevant greenhouse gases were considered,
namely: CO2 , CH4 , and N2 O. The amount of CH4 and N2 O emitted over the entire process is converted into CO2 s equivalents
(CO2-equivalent ) by using the global warming potential factor of 21
for CH4 and 310 for N2 O. A European electricity mix with 15%
share of renewables and 32% share of nuclear was assumed in this
study [45].

618

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619

1 kg CO2-equivalent per kg syngas, of which 84% are derived


from the combustion of natural gas in the endothermic
steam-reforming step [45]. Once the external costs are internalized, the cost of solar H2 is expected to become competitive with that of H2 produced using fossil-fuel-based
technologies.
The cost of electricity generated by H2 =O2 fuel cells that
are fueled with solar H2 ranges between 0.310:33$=kWhe ,
for the baseline case. This cost is not competitive against
that of bulk electricity generated directly via solar thermal or
even solar PV conversion systems, provided these systems
are constrained to similar annual solar irradiation conditions
of at least 2300 kW=m2 =yr with 2300 equivalent full power
hours per year. Thus, for large-scale stationary applications
in regions of high insolation, direct solar electricity generation is clearly the preferable path, whereas for mobile applications it is obviously not a viable option. Solar thermal
electricity systems can also feature on-site thermal storage capabilities (e.g. using molten salt as the heat transfer
medium [40]) to allow for solar electricity dispatchability
after sunset at more competitive prices than using chemical
storage in the form of hydrogen. However, for regions
having poor insolation, the cost of direct solar electricity
generation often exceeds 0:30$=kWhe . Transmission of solar electricity from regions of high insolation may then be
a viable option; its cost will need to be compared with the
cost of electricity from fuel cells fueled with solar H2 , including the additional cost of H2 storage and transport. For
mobile applications such as powering electric vehicles, fuel
cells fueled with solar H2 may compare favorably when
judged against the alternative of using rechargeable batteries of similar performance and being charged with solar
electricity. Furthermore, solar H2 decouples the collection
of solar energy and the generation of solar electricity, so
that solar-H2 -fueled fuel cells can supply solar electricity around-the-clock to meet customers energy demands
whenever and wherever needed.
The weaknesses of this economic assessment are related
primarily to the uncertainties in the viable e3ciencies and
investment costs of the various components due to their
early stage of development and their economy of scale. The
technical feasibility of large-scale windowed solar reactors,
quenching techniques, and zinc hydrolysers needs demonstration. The solar tower-re9ector + CPC technology also
needs up-scaling to commercial size, but an alternative solar plant con-guration with the receiver-reactor at the top of
the tower could be developed.
4. Summary and conclusions
The two-step water-splitting cycle, based on the solar
ZnO-dissociation followed by the Zn-hydrolysis, bypasses
the H2 =O2 separation problem and can reach a maximum
exergy e3ciency of 29%. Critical key components are
the solar reactor operating at 2300 K, the Zn(g)=O2

quenching device, and the Zn(l)=steam hydrolyser. An


economic evaluation predicts a unit cost of solar H2 varying in the range 0.11 0.17$=kWh for a heliostat -eld at
50 200$=m2 . Thus, the proposed solar cycle, if realized
at an industrial scale, can be competitive with the electrolysis of water using solar-generated electricity. Further
development and large-scale demonstration are warranted.
Acknowledgements
The author thanks Dr. A. Meier, Dr. M. Sturzenegger,
and Dr. C. Wieckert for their critical review. Financial support from the Swiss Federal O3ce of Energy is gratefully
acknowledged.
Appendix
The 2nd -law analysis is veri-ed by performing an energy
balance and by evaluating the maximum achievable e3ciency (Carnot e3ciency) from the total available work and
from the total power input. The energy balance con-rms that
WFC = Qsolar (Qreradiation + Qquench + Qhydrolyser + QFC ):
The available work is calculated as the sum of the ideal
fuel-cell work and the lost work due to irreversibilities in
the solar reactor, the hydrolyser, and during quenching. This
maximum e3ciency must be equal to that of a Carnot heat
engine operating between 2300 and 298 K. Thus,
max =

WFC = FC + 298 K(Irr reactor + Irr quench + Irr hydrolyser )


Qsolar

= Carnot = 1

298 K
= 87%:
2300 K

References
[1] Fletcher EA, Moen RL. Hydrogen and oxygen from water.
Science 1977;197:1050 6.
[2] Kogan A. Direct solar thermal splitting of water and on-site
separation of the products. II. Experimental feasibility study.
Int J Hydrogen Energy 1998;23:8998.
[3] Ihara S. On the study of hydrogen production from water using
solar thermal energy. Int J Hydrogen Energy 1980;5:52734.
[4] Fletcher EA. Solarthermal and solar quasi-electrolytic
processing and separations: Zinc from Zinc Oxide as an
example. Ind Eng Chem Res 1999;38:227582.
[5] Funk J. Thermochemical hydrogen production: past and
present. Int J Hydrogen Energy 2001;26:18590.
[6] Serpone N, Lawless D, Terzian R. Solar Fuels: Status and
Perspectives. Solar Energy 1992;49:22134.
[7] Welford WT, Winston R. High collection nonimaging optics.
San Diego, CA: Academic Press, 1989.
[8] Yogev A, Kribus A, Epstein M, Kogan A. Solar tower re9ector
systems: A new approach for high-temperature solar plants.
Int J Hydrogen Energy 1998;23:239 45.

A. Steinfeld / International Journal of Hydrogen Energy 27 (2002) 611 619


[9] Segal A, Epstein M. Comparative performance of tower-top
and tower-re9ector central solar receivers. Solar Energy
1999;65:20726.
[10] Steinfeld A, Kuhn P, Reller A, Palumbo R, Murray J,
Tamaura Y. Solar-processed metals as clean energy carriers
and water-splitters. Int J Hydrogen Energy 1998;23:76774.
[11] Bilgen E, Ducarroir M, Foex M, Sibieude F, Trombe F. Use
of solar energy for direct and two-step water decomposition
cycles. Int J Hydrogen Energy 1977;2:2517.
[12] Nakamura T. Hydrogen production from water utilizing solar
heat at high temperatures. Solar Energy 1977;19:46775.
[13] Palumbo R, Lede J, Boutin O, Elorza-Ricart E, Steinfeld
A, Moeller S, Weidenka2 A, Fletcher EA, Bielicki J. The
production of Zn from ZnO in a single-step high temperature
solar decomposition process. Chem Eng Sci 1998;53:
250318.
[14] Sibieude F, Ducarroir M, To-ghi A, Ambriz J.
High-temperature experiments with a solar furnace: the
decomposition of Fe3 O4 , Mn3 O4 , CdO. Int J Hydrogen
Energy 1982;7:7988.
[15] Steinfeld A, Sanders S, Palumbo R. Design aspects of solar
thermochemical engineering. Solar Energy 1999;65:4353.
[16] Ehrensberger K, Frei A, Kuhn P, Oswald HR, Hug P.
Comparative experimental investigations on the watersplitting reaction with iron oxide Fe1y O and iron manganese
oxides (Fe1x Mnx )1y O. Solid State Ion 1995;78:15160.
[17] Tamaura Y, Steinfeld A, Kuhn P, Ehrensberger K. Production
of solar hydrogen by a novel, 2-step, water-splitting
thermochemical cycle. Energy 1995;20:32530.
[18] Hirschwald W, Stolze F. Zur Kinetik der thermischen
Dissoziation von Zinkoxid, Ze. Physikalische Chemie Neue
Folge 1972;77:21 42.
[19] Weidenka2 A, Reller A, Wokaun A, Steinfeld A.
Thermogravimetric analysis of the ZnO=Zn water splitting
cycle. Thermochimica Acta 2000;359:6975.
[20] Moeller S. Entwicklung eines Reaktors zur solarthermischen
Herstellung von Zink aus Zinkoxid zur Energiespeicherung
mit Hilfe konzentrierte Sonnenstrahlung. Ph.D. thesis, ETH Swiss Federal Institute of Technology, Zurich, 2001.
[21] Weidenka2 A, Steinfeld A, Wokaun A, Eichler B, Reller A.
The direct solar thermal dissociation of ZnO: condensation
and crystallization of Zn in the presence of oxygen. Solar
Energy 1999;65:5969.
[22] Palumbo RD, Fletcher EA. High temperature solar
electro-thermal processing III. Zinc from zinc oxide at 1200
1675 K using a non-consumable anode. Energy 1988;13:
31932.
[23] Parks DJ, Scholl KL, Fletcher EA. A study of the use of
Y2 O3 doped ZrO2 membranes for solar electrothermal and
solar thermal separations. Energy 1988;13:12136.
[24] Weidenka2 A, Reller A, Sibieude F, Wokaun A, Steinfeld A.
Experimental investigations on the crystallization of zinc by
direct irradiation of zinc oxide in a solar furnace. Chem Mater
2000;12:217581.
[25] Elorza-Ricart E, Martin PY, Ferrer M, Lede J. Direct
thermal splitting of ZnO followed by a quench. Experimental
measurements of mass balances. J Phys IV France
1999;9:32530.
[26] Lede J, Boutin O, Elorza-Ricart E, Ferrer M. Solar thermal
splitting of zinc oxide: a review of some of the rate controlling
factors. J Solar Energy Eng 2001;123:917.

619

[27] Haueter P, Moeller S, Palumbo R, Steinfeld A. The production


of zinc by thermal dissociation of zinc oxide solar chemical
reactor design. Solar Energy 1999;67:1617.
[28] Berman A, Epstein M. The kinetics of hydrogen production in
the oxidation of liquid zinc with water vapor. Int J Hydrogen
Energy 2000;25:95767.
[29] Cortina C. M.Sc. thesis, ETH-Swiss Federal Institute of
Technology, Zurich, 2001.
[30] Graf GG. Zinc. In: Gerhartz W, editor. Ullmanns
encyclopedia of industrial chemistry A, Vol. 28, 5th ed.
Weinheim, Germany: VCH Verlag, 1996.
[31] Steinfeld A, Palumbo R. Solar thermochemical process
technology. In: Meyers RA, editor. Encyclopedia of physical
science and technology. New York: Academic Press, ISBN
0-12-227410-5, 2001;15:23756.
[32] Steinfeld A, Schubnell M. Optimum aperture size and
operating temperature of a solar cavity-receiver. Solar Energy
1993;50(1):1925.
[33] Spiewak I. In: In: Roy A, editor. Introductory guidance for
economic evaluation of solar-thermal power plants. IEA Solar
PACES, 1997.
[34] Steinfeld A, Spiewak I. Economic evaluation of the solar
thermal co-production of zinc and synthesis gas. Energy
Conversion Management 1998;39:15138.
[35] Sanchez M, Romero M, Ajona J. Proceedings of the
Eighth International Symposium Solar Thermal Concentrating
Technologies, Vol. I, Cologne, Germany, October 6 11, 1996.
p. 315 31.
[36] Science Applications International Co. Heliostat manufacturing for near-term markets. Phase I Summary Report,
NREL-Contract ZAP-5-14168-02, 1996.
[37] Kribus A, Zaibel R, Carey D, Segal A, Karni J. A
solar-driven combined cycle power plant. Sol Energy 1998;62:
1219.
[38] Larson ED, Katofsky RE. Production of methanol and
hydrogen from biomass. Princeton University Center for
Energy and Environmental Studies Report No. 271,
1992.
[39] Spiewak I, Tyner CE, Langnickel U. Solar Reforming
Applications Study Summary. Proceedings of the Sixth
International Symposium on Solar Thermal Concentrating
Technologies, Vol. 1, Mojacar, Spain, September 28October
2, 1992. p. 955 68.
[40] Tyner CE, Kolb GJ, Geyer M, Romero M. Concentrating solar
power in 2001 an IEA=SolarPACES summary of present
status and future prospects. SolarPACES, Sandia National
Laboratories. Albuquerque, New Mexico, USA, 2001.
[41] Glatzmaier G, Blake D, Showalter D. Assessment of
Methods for Hydrogen Production using Concentrated
Solar Energy, NREL=TP-570-23629, National Renewable
Energy Laboratory, Golden, CO, 1998 (available From
the National Technical Information Service, Spring-eld,
VA 22161).
[42] Sturzenegger M. Private communication 2001.
[43] Steinberg M, Cheng HC. Modern and prospective technologies
for hydrogen production from fossil fuels. Int J Hydrogen
Energy 1989;14:797820.
[44] Basye L, Swaminathan S. Hydrogen production costs a
survey, DOE=GO=10170-778, 1997.
[45] Werder M, Steinfeld A. Life cycle assessment of the
conventional and solarthermal production of zinc and
synthesis gas. Energy Int J 2000;25:395 409.

Potrebbero piacerti anche