You are on page 1of 9

Energy Conversion and Management 51 (2010) 2353–2361

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Hydrogen production by hydrogen sulfide splitting using concentrated solar


energy – Thermodynamics and economic evaluation
W. Villasmil a, A. Steinfeld a,b,*
a
Department of Mechanical and Process Engineering, ETH Zurich, 8092 Zurich, Switzerland
b
Solar Technology Laboratory, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland

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

Article history: Thermodynamic and economic analyses were carried out to evaluate the use of concentrated solar energy
Received 23 August 2009 for driving the endothermic dissociation reaction H2S ? H2 + 0.5S2. Three different schemes were
Received in revised form 25 March 2010 assessed: (1) a pure solar process; (2) a hybrid process, which uses both solar and natural gas combustion
Accepted 3 April 2010
as the energy sources of high-temperature process heat; and (3) the Claus process. This study indicates
Available online 18 May 2010
that the pure solar process has the potential of lowering the disposal costs of H2S vis-à-vis the conven-
tional Claus process while co-producing H2 without concomitant CO2 emissions. An economic assessment
Keywords:
for a 40 MWth chemical plant using solar tower technology indicates savings of approximately 45% in
Solar energy
Hydrogen
comparison to the Claus process. Solar H2 production is estimated at a cost in the range of
Hydrogen sulfide 0.061–0.086 $/kW h, based on its lower heating value and without credit attributed to H2S disposal. A
Thermochemical cycle sensitivity analysis revealed that the quench efficiency represents the parameter with the highest impact
Economic assessment on the economics of the process. A hybrid natural gas/solar plant design able to operate 24 h-a-day is pre-
Claus process dicted to reduce the H2 production cost to 0.058 $/kW h at current fuel prices, however, at the expense of
increased complexity related with the hybrid reactor design and operation plus the associated CO2
emissions of 0.42 kg/kW h.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction Currently, the two-step Claus process is the most widely


applied procedure to remove sulfur from H2S. In the first, thermal
Several thermochemical processes have been proposed to split step, approximately one third of the H2S is partially oxidized above
H2S using concentrated solar energy [1–3], represented by: 1300 K according to [6]:

2H2 S þ 3O2 ! 2SO2 þ 2H2 O ð2Þ


1
H2 S
H2 þ S2 ð1Þ The remaining H2S enters the second catalytic step where it
2
reacts with the SO2 formed in the previous stage according to the
The advantages of this solar-driven process are multifold: (1)
so-called Claus reaction:
the upper temperature requirement is lower than that for direct
water thermolysis or metal oxides based thermochemical cycles 2H2 S þ SO2
þ 2H2 O þ 3S ð3Þ
for H2 generation [4]; (2) a highly toxic waste product is disposed
while H2 is co-produced; and (3) no greenhouse gases are emitted. Generally, between 1 and 3-stage units are required for the cat-
The oil and natural gas industry is the main source of H2S, derived alytic step, each unit consisting of: heating, catalytic reaction, cool-
from the sweetening of natural gas and the removal of organically ing and condensation. One important disadvantage is that by
bound sulfur from petroleum. Since stringent environmental regu- introducing air as oxidant, a large volume of tail-gas is produced,
lations require the removal of nearly all the sulfur present in fossil which needs to be further treated to comply with environmental
fuels, the economic impact of the removal and disposal units are so regulations. Moreover, the process wastes H2 by oxidizing H2S to
adverse that some natural gas wells are simply not exploited. The H2O to produce low-grade process heat.
annual production of sulfur has been estimated in 2.4  106 tons The proposal of decomposing H2S to co-produce H2 and S2 dates
from refinery streams and 1.8  106 tons from natural gas [5]. back more than 30 years. Raymont first proposed the use perme-
able membranes to remove H2 from the hot gas mixture resulting
* Corresponding author at: Department of Mechanical and Process Engineering,
from the dissociation reaction [7]. Fletcher, Noring et al. [1,8] stud-
ETH Zurich, 8092 Zurich, Switzerland. Tel.: +41 44 6327929; fax: +41 44 6321065. ied the solar thermal dissociation of H2S at up to 2000 K and
E-mail address: aldo.steinfeld@ethz.ch (A. Steinfeld). 0.5 atm. Their original idea was to use a reactor-separator

0196-8904/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2010.04.009
2354 W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361

Nomenclature

C solar flux concentration ratio EPCM engineering, procurement, commissioning, manage-


h enthalpy (kJ mol1) ment
I normal beam insolation (kW m2) IHR integrated heat recovery
n_ molar flow rate (mol s1) LHV lower heating value (kJ mol1)
Q_ chemical power absorbed to drive the chemical dissociation reac- NG natural gas
tion (kW) NHR no heat recovery
Q_ cooler rate of heat rejected to cooling water in the cooler (kW) TT tower-top
Q_ heating power absorbed to heat up the reactants to the reactor
temperature (kW)
Q_ quench rate of heat rejected by the products in the quenching Greek symbols
apparatus (kW) aeff effective absorptance
Q_ reactor;net net power absorbed by solar reactor (kW) eeff effective emittance
Q_ solar solar power input to reactor (kW) gabsorption solar energy absorption efficiency
Q_ sulfur condenser rate of heat rejected by the products in the sulfur gquench quench efficiency
condenser (kW) gprocess process efficiency
W _ compression compression work (kW)
r Stefan–Boltzmann constant (W m2 K4)
T temperature (K)
X extent of dissociation
Subscripts
Acronyms in inlet to solar reactor
BD beam-down out outlet of solar reactor
CPC compound parabolic concentrator

containing a porous material for dissociation and in situ effusional unreacted is removed from the product stream and recycled. Sulfur
separation of H2 and S. Later studies [2,9,10] demonstrated that a is recovered at atmospheric pressure in its various liquid allotropes
simple quench process can lead to almost complete product phase (Sx) and H2 is recovered at a pressure of approximately 24 bars.
separation. The kinetics of the dissociation reaction has been char-
acterized for homogeneous and heterogeneous components with
2.1. Storage/recycling
various catalysts (e.g. SiO2, MoS2, WS2 and Al2O3) [11–14].
This paper presents the thermodynamic and economic analyses
When solar energy is not available, the H2S fed to the plant is
of the production of H2 via reaction (1) using solar process heat.
compressed, liquefied and stored at 25 bars and 298 K. Once solar
The technical feasibility of solar-driven systems has been thor-
energy becomes available, stored H2S is recycled by mixing it with
oughly studied, but the various solar/fossil hybrid scenarios and
the external feed stream. In designing the process, the ratio of both
their economic viability has not been yet treated with the same
streams was set to ensure a constant flow rate of H2S is fed to the
rigor. The objective of this study is to identify the economical
plant throughout the whole day.
potentials and limitations of the proposed solar process by com-
paring it with the conventional technologies. Three different
schemes were evaluated: (1) pure solar process; (2) hybrid process, 2.2. Preheating/sulfur condensation
which uses both concentrated solar radiation and combustion of
natural gas as the energy sources of high-temperature process The resulting saturated H2S stream leaving the mixer is pre-
heat; and (3) the Claus process. These analyses require a definition heated in an adiabatic counter-flow heat exchanger where part
of the material and energy flows and an estimation of the costs of of the sensible and latent heat of the products is recovered while
the plant components and the associated operational and mainte- sulfur is condensed. For the NHR design, the outlet temperature
nance costs. of the products is limited to 675 K to avoid sulfur solidification.
For the IHR design, the outlet temperature of the reactants is set
to 300 K to guarantee an effective quench of the product stream.
2. The solar H2S splitting process
2.3. Quench
The flow diagrams of the proposed solar thermal dissociation
process of H2S are depicted in Fig. 1. For both thermodynamic
Pressurized water is used as the cooling fluid in the quench unit
and economic analyses, two plant designs were considered: (1) a
for the NHR design. For the IHR design, it has been experimentally
plant without heat recovery (NHR), where the sensible heat of
demonstrated that it is possible to reach quench efficiencies
the product gases are not recovered during the quench; and (2) a
higher than 70% by using the reactants as the cooling fluid in this
plant with integrated heat recovery (IHR), where some portion of
process [9].
the sensible heat is recovered by using the reactants as the cooling
fluid in the quench unit. The solar plant is comprised of a solar
reactor, three heat exchangers, a liquid–gas mechanical separator, 2.4. Solar reactor
an amine scrubber, a storage vessel and two compressors. An
external source of H2S gas is fed to the process at 298 K and Concentrated solar energy is absorbed by the reactor to heat up
1 bar, mixed with recycled H2S, preheated and finally injected to the H2S and drive the endothermic dissociation reaction. Besides
the solar reactor. In the reactor, concentrated solar energy provides the main products H2 and S2, other minor species such as S3–S8,
the high-temperature process heat to drive the chemical reaction HS and H2S2–H2S9 might also be present in the product mixture
(1) and dissociate H2S to H2 and S2. The portion of H2S that remains depending on the operating conditions.
W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361 2355

Fig. 1. Conceptual system design of a solar chemical plant for the production of H2 and Sx via H2S thermal dissociation: (a) NHR – No heat recovery: the products are
quenched using cooling water; (b) IHR – Integrated heat recovery: the products are quenched with the H2S fed to the solar reactor.

2.5. Separation processes (aeff = eeff = 1, convection/conduction losses neglected) and a solar
concentration ratio of 2500 suns1. The molar feed rate of H2S to
Liquid sulfur leaving the condenser is removed from the gas the solar reactor, n_ H2 S , was calculated such that the net power ab-
stream by means of a two-phase mechanical separator. The sorbed by the reactor matches the enthalpy change per unit time:
remaining gas mixture, composed mainly by H2 and unreacted  
H2S, is further cooled, compressed and finally injected to the amine
Q_ reactor;net ¼ n_ H2 S;out  hH2 S jTout þ n_ H2 ;out  hH2 jTout þ n_ S2 ;out  hS2 jTout
 
scrubber where the H2S is removed and recycled, while H2 is recov-  n_ H2 S;in  hH2 S ð4Þ
ered as the final product.
where the quantities inside the first brackets correspond to the
products while those within the second brackets correspond to
3. Thermodynamic analysis the reactants. The value of the H2S inlet temperature, Tin, depends
on the amount of heat recovered in the heat exchangers. The net
The HSC Outokumpu code [15] was used to compute the equi- power absorbed by the solar reactor is given by:
librium composition in the ranges 600–2000 K and 0.1–10 bars;
species whose mole fraction is <105 have been omitted from the 1
Solar concentration ratio is defined as the ratio of mean solar radiative flux over
figures. The energy balance was carried out assuming the solar the reactor aperture to the normalized value of 1 kW/m2, and is usually given in units
reactor to be a perfectly insulated blackbody cavity-receiver of ‘‘suns”.
2356 W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361

4
!
r  T cavity
Q_ reactor;net ¼ Q_ solar  gabsorption ¼ Q_ solar  1  ð5Þ
IC

where Q_ solar is the concentrated solar power intercepted by the solar


reactor, I is the normal beam insolation (taken in this study equal to
1 kW/m2), C is the flux concentration ratio of the solar concentrat-
ing system, Tcavity is the nominal cavity temperature, and r is the
Stefan–Boltzmann constant. The solar absorption efficiency,
gabsorption , is then defined as the portion of incoming solar power that
is net absorbed by the solar reactor.
Fig. 2 shows the equilibrium composition of the system H2S as a
function of temperature at 1 bar. H2S, H2 and S2 are the main spe-
cies in the 1500–1700 K range, along with small amounts (<1%) of
HS and H2S2. The reaction extent of H2S dissociation, X, is defined
as:
n_ H2 S;out Fig. 4. Process efficiency, defined in Eq. (7), as a function of the reactor temperature
X ¼1 ð6Þ at 0.1 and 10 bars.
n_ H2 S;in
where n_ H2 S;in and n_ H2 S;out denote the molar flow rate of H2S entering
and leaving the solar reactor respectively. X is shown in Fig. 3 as a where n_ H2 and n_ S2 represent the amount of hydrogen and sulfur
function of temperature at 0.1, 1, and 10 bars. The process effi- recovered as final products. The lower heating values (LHV) for
ciency, gprocess, is defined as the ratio of the products energy content H2S, H2, and S2 are 517, 241 and 119 kJ/mol respectively. Isothermal
to the total energy input to the system: compression and a perfect quench were assumed when using Eq.
(7) to calculate the maximum gprocess that could be reached under
H2n_  LHV
H2 S2þ n_  LHV
S2 ideal conditions. gprocess is plotted in Fig. 4 as a function of the reac-
gprocess ¼ _ _ compression
ð7Þ
Q solar þ n_ H S;in  X  LHVH2 S þ W
2 tor temperature at 0.1 and 10 bars. For a given pressure there is an
optimal temperature at which maximum efficiency is achieved.
Operating above this temperature would imply increased re-radia-
tion losses that would not be compensated by the increase in X.
Fig. 4 also points out to the importance of recovering the sensible
heat rejected during the quench. However, this effect becomes less
_
important as the pressure is reduced because of the increased QQ_chemical
heating

ratio, which in turn reduces the sensible energy rejected during the
quench.

4. Economic assessment

All costs are reported in 1st quarter 2009 USD. The baseline
operating conditions for the economic assessment are listed in
Table 1. The concentrating solar system is based on the solar tower
technology consisting of a field of heliostats – two-axis tracking
parabolic mirrors – that focus the sun rays onto a solar reactor
mounted on top of a centrally located tower [16]. Two optical con-
Fig. 2. Thermodynamic equilibrium composition of the system H2S as a function of figurations are evaluated in this study: (1) tower-top (TT), in which
temperature at 1 bar. the solar reactor is installed at the top of the tower and directly re-
ceives the concentrated solar irradiation from the heliostat field;
and (2) beam-down (BD), in which the solar reactor is placed on
the ground and the focused solar energy is redirected by a hyper-
bolic reflector located at the top of the tower [17,18]. In the beam-
down case, a compound parabolic concentrator (CPC, [19]) is

Table 1
Baseline operating conditions and design parameters of the solar thermochemical
plant.

Parameter Unit Value


Solar reactor temperature K 1700
Solar reactor pressure bar 1
Solar concentration ratio (tower-top) suns 2500
Solar concentration ratio (beam-down) suns 5000
Number of heliostats 624
Heliostat field reflective area m2 75,504
Equivalent full power hours h/yr 2300
Optical efficiency % 65
Quench efficiency for NHR design % 80
Fig. 3. Reaction extent, defined as in Eq. (6), as a function of the reactor Quench efficiency for IHR design % 70
temperature at 0.1, 1, and 10 bars.
W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361 2357

incorporated at the reactor’s aperture to boost the solar concentra- 4.2. Economic assumptions
tion ratio. In sizing and evaluating the annual performance of the
solar concentration system, the heliostat field of the commercial Capital costs were updated using the Component Cost Index for
PS10 solar power plant located in Seville [20] was taken as a Chemical Processes. The operation and maintenance cost was as-
reference. sumed to be 2% of the capital cost. The assumed discount rate (real
interest rate) of 15% per annum is representative of private owner-
ship with a significant fraction of borrowed money. Early plants
4.1. Single equipment cost
would probably have a higher discount rate because of the per-
ceived higher risk. Plant design costs, such as engineering, procure-
4.1.1. Heliostat field
ment, commissioning and management (EPCM), were taken into
The heliostat field represents the largest single cost in the solar
account as part of the indirect cost. Generally, indirect costs
plant. A specific cost of 160 $/m2 was assumed based on Refs.
amount to 20% of the initial capital cost except those for the helio-
[21,22].
stats that amount to only 10% of the total capital cost [23]. On top,
a contingency of 15%, typical for high risk projects, was also added.
4.1.2. Tower, tower reflector and CPCs Government subsidies have been excluded from consideration. The
Correlations provided in Ref. [23] were used for sizing the helio- credit for sulfur sale was accounted at 30 $/ton based on historical
stat field area, tower height, and reflector area and further estimate values and actual markets [27,28], while H2S was assumed to have
their cost. CPC cost was assumed to be proportional to the solar zero value.
reactor power input and the tower reflector area was assumed to
be 2.3% of the heliostat field area.
4.3. Results

4.1.3. Solar reactor Table 2 shows the material and energy flows for the different
The proposed solar reactor consists of a volumetric-type recei- plant designs evaluated. A plant of the size considered could pro-
ver equipped with an Al2O3 honeycomb structure that serves as duce between 1% and 1.5% of the world sulfur recovered from fossil
the radiant absorber. The reactor was sized using a simplified mod- fuels. Table 3 shows the cost breakdown of the solar chemical
el based on the H2S dissociation kinetics [11]. Direct industrial plant. The specific cost of solar H2 is estimated to be in the range
quotes were obtained to estimate the cost of the honeycomb of 6.7–8.6 ¢/kW h (based on its LHV), assuming the baseline
(4100 $/m3) and the refractory layer (2200 $/m2). The cost of a car- parameters shown in Table 1. Regardless of the reduced quench
bon steel shell was added and a multiplier of 2.0 was applied to efficiency in the IHR scheme, the effect of recovering the sensible
estimate the final installed cost. heat has a positive impact on the process economics. This is indi-
cated by a reduction of nearly 20% in the H2 specific cost for both
4.1.4. Storage vessels optical configurations studied. For the solar concentration ratios
The storage system was sized to allow 15 days of continuous considered, both optical configurations exhibit similar H2 specific
storage. Stainless steel grade 304 was selected as the vessel mate- costs. However, for concentration ratios less than 2500, the perfor-
rial based on the highly corrosive nature of H2S. Due to the high mance of the TT configuration decreases significantly while the BD
toxicity of H2S, a 50% reserve storage capacity was added to enable configuration shows less sensitivity to the solar concentration ratio
rapid emptying of liquid H2S in case of accident. Capital costs were (Fig. 5).
estimated based on Ref. [24].
4.3.1. Sensitivity analysis
A sensitivity analysis was performed for the TT configuration
4.1.5. Amine scrubber
and a solar concentration ratio of 2500. Four different parameters
Based on recommendations by industry experts, an amine solu-
were considered: (1) specific heliostat cost; (2) reactor pressure;
tion rate of 0.22 l/mol of H2S absorbed is required to effectively
(3) quench efficiency; and (4) sulfur price. Although the heliostat
separate the H2S–H2 mixture. Such a rate would achieve high pur-
field represents nearly a third of the total direct capital cost,
ity outlet streams of 99.8% H2 and 99.9% H2S. The cost of the com-
plete amine unit was estimated based on an industrial quotation
and Ref. [25]. Table 2
Material and energy flows of the solar thermochemical plant for H2S dissociation
cycle shown in Fig. 1. Two optical configurations are evaluated: tower-top (C = 2500)
4.1.6. Compressors and beam-down (C = 5000). For each case, two plant designs are considered: no heat
The compression process was assumed to take place isother- recovery (NHR) and integrated heat recovery (IHR).
mally at 320 K and a mechanical efficiency of 85% was used. The
Solar concentration ratio 2500 5000
capital cost of each compression unit was estimated based on
Plant design Unit NHR IHR NHR IHR
Ref. [26] and a 10% increase was added to account for the cost of
the drives. Material flows
Design H2S feed to solar reactor mol/s 323 562 361 628
H2 recovered after scrubber mol/s 147 223 164 249
4.1.7. Quencher, heat exchangers, sulfur separator S2 recovered after the separator mol/s 71 107 79 120
The quench efficiency was taken as 80% and 70% for NHR and H2S recycled after the scrubber mol/s 173 334 194 373
H2S external feed mol/s 40 60 44 67
IHR designs, respectively, the difference being the result of a de- Amine flow circulation l/s 38.1 73.3 42.6 81.9
creased heat transfer coefficient and a smaller mean temperature
Energy flows
difference for the IHR design. Stainless steel grade 304 was selected Qsolar MW 75.5 75.5 75.5 75.5
as the material for the cooler and the sulfur condenser. Ceramic Qreactor,net MW 39.8 39.8 44.4 44.4
materials are proposed for the quench unit. A cyclone-type Qquench MW 14.4 25.2 16.1 28.1
mechanical separator is used to remove the liquid sulfur from Qsulfur condenser MW 1.9 7.8 2.1 8.7
Qcooler MW 4.8 4.0 5.4 4.4
the H2S–H2 gas stream after the sulfur condenser. In estimating
Wcompression MW 5.3 9.5 5.9 10.6
the capital costs, the correlations provided in Ref. [26] were used.
2358 W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361

Table 3
Plant size, annual production, and estimated costs of the solar thermochemical plant for H2S dissociation. Two optical configurations are evaluated: tower-top (C = 2500) and
beam-down (C = 5000). For each case, two plant designs are considered: no heat recovery (NHR) and integrated heat recovery (IHR).

Plant design Tower-top (TT) Beam-down (BD)


NHR IHR NHR IHR
Solar plant size (solar power input to reactor) (MWth) 39.8 39.8 44.4 44.4
H2S storage capacity for 15 days (m3) 3375 5158 3770 5761
Solar input on heliostat field (MW h/yr) 173,659 173,659 173,659 173,659
H2 production (based on the LHV) (million-kW h/yr) 81 123 90 138
S2 production (tons/yr) 37,537 56,906 45,419 63,555
Efficiencies
Solar concentration ratio, C 2500 2500 5000 5000
Optical efficiency of solar concentrating system (%) 65 65 65 65
Solar reactor absorption efficiency, gabsorption (%) 81 81 91 91
Overall dissociation, X  gquench (%) 46 40 46 40
Process efficiency, gprocess (%) 29 34 30 36

Capital cost (million $)


Heliostat field (assuming 160 $/m2) 12.1 12.1 12.1 12.1
Land (2 $/m2) 0.4 0.4 0.4 0.4
Tower 1.7 1.7 1.2 1.2
Tower reflector + CPCs – – 2.8 2.8
Solar receiver-reactor 1.2 1.9 1.1 1.7
Compressors 4.0 5.7 4.2 6.1
Heat exchangers + sulfur separator 1.4 3.2 1.6 3.5
Amine scrubber 7.9 11.4 8.5 11.7
Storage vessels 7.5 9.2 7.9 9.7
Total direct cost 36.3 45.6 39.9 49.3
EPCM (indirect) 6.0 7.9 6.8 8.7
Contingency 5.4 6.8 6.0 7.4
TOTAL capital cost 47.8 60.3 52.7 65.4
Annual cost (million $)
Capital cost 7.0 9.0 7.9 9.8
Operations and maintenance 1.0 1.2 1.1 1.3
Credit for sulfur sale (30 $/ton) 1.1 1.7 1.3 1.9
TOTAL annual cost 7.0 8.5 7.7 9.2
Specific cost
Unit cost of solar H2 (¢/kW h LHV) 8.6 6.9 8.5 6.7

leads to the same H2 specific cost as the IHR plant with


gquench ¼ 70%. This is explained by its twofold effect: (1) it affects
the amount of H2 and S2 produced; and (2) it affects the amount
of H2S recycled and therefore the scrubber size, the cost of which
also represents an important share of the total direct cost (22%).

4.3.2. Comparison with Claus process


A Claus unit with the same sulfur production capacity of the so-
lar chemical plant3 has an estimated cost of 22.1 M$ [24]. In addi-
tion, the cost of the tail-gas clean-up process unit is generally
equal to 75% of the Claus unit alone [29]. Assuming the same eco-
nomic parameters as for the solar plant4, the total annual cost asso-
ciated with the Claus plant would be 5.4 M$. On the other hand, the
solar plant co-produces hydrogen. The credit for H2 production was
conservatively taken as 3.8 ¢/kW h, corresponding to the lowest pro-
duction cost available when using the conventional reforming of nat-
Fig. 5. Hydrogen unit cost as a function of the solar concentration ratio. The
discontinuities at C = 2500 are a consequence of the change in optical configuration
ural gas [30]. This would lead to 3.0 M$ in annual cost for the solar
of the solar concentrating-system. plant, compared with 5.4 M$ for the Claus process, meaning nearly
45% in annual savings. Moreover, since the economics of both plants
depend on the sulfur production in the same manner, the absolute
halving its specific cost results in a reduction of only 18% in the fi- savings would be identical regardless of the sulfur price.
nal H2 specific cost (Fig. 6a). In contrast, the quench efficiency2 re-
veals a much stronger effect on the economics: a 30% relative
5. Hybrid natural gas/solar chemical plant
increase of gquench would lead to a 30% reduction in the H2 unit price
(Fig. 6c). It is noteworthy that the NHR plant with gquench ¼ 90%
Considering the high capital cost associated with the storage
system in the solar chemical plant (Table 3), a NG/solar-hybrid
2
3
H2 molar flow after quench Sulfur production of 111 tons/day corresponding to the tower-top (C = 2500) and
gquench ¼ no heat recovery design.
H2 molar flow before quench
4
Sulfur price = 30 $/ton; fixed charge rate = 15%; O and M cost = 2% of capital cost.
W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361 2359

Fig. 6. Hydrogen unit cost sensitivity to: (a) heliostat cost; (b) reactor pressure; (c) quench efficiency; (d) sulfur price. The results shown correspond to the tower-top
configuration and a solar concentration ratio C = 2500.

plant is evaluated, capable of operating 24 h a day and thus elimi- Fig. 1. The economic evaluation demonstrated that recovering the
nating the need for large overnight storage capacity. The baseline heat in the hybrid scheme would lead to higher H2 specific cost.
parameters for the hybrid plant are the same as those indicated This is due to the fact that the heat required to drive the endother-
in Table 1, except for the reactor temperature which was reduced mic reaction would be added at a much lower temperature differ-
to 1500 K due to the inherent limitation of the hybrid reactor de- ence, which in turn would increase the size and thus the cost of the
sign. The plant layout corresponds to the NHR design shown in reactor considerably.

Fig. 7. Design configuration of the hybrid natural gas–solar reactor for H2S dissociation.
2360 W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361

5.1. Hybrid reactor impact on the hybrid plant than on the solar-only plant, as ob-
served by the characteristic cost breakdowns of each plant: while
Fig. 7 shows a scheme of the proposed design configuration of for the hybrid plant the reactor represents half of the total direct
the hybrid reactor. The first section is comprised of a honeycomb cost, for the solar plant it accounts for only 5%. This major differ-
structure directly exposed to concentrated solar irradiation, serv- ence in the reactor share is a direct consequence of the heat trans-
ing as the solar absorber and heat transfer medium to the flow of fer modes and thermal efficiencies associated with each reactor
reactants. A transparent quartz window provides the optical access design. Finally, an important drawback of the NG/solar-hybrid
to the solar power input. The second section consists of a counter- plant is the concomitant CO2 emissions at a rate of 0.42 kg/kW h
flow heat exchanger in which reactants flow through the tubes LHV of H2, which would be completely avoided in the solar-only
while hot product gases flow outside, thus supplying the process plant.
heat required to drive the dissociation reaction. Natural gas (NG)
is oxidized in an adiabatic burner with approximately 60% excess
air to allow the hot flue gases to enter the reactor at 1700 K. The 6. Summary and conclusions
reactor dimensions were determined by applying a simplified
model based on the H2S dissociation kinetics [11]. For Qso- The economic evaluation indicates that the proposed solar-dri-
ven H2S dissociation process is economically advantageous vis-à-
lar = 40 MW, a tube diameter of 50.8 mm was found to be the opti-
mum to minimize the reactor cost. The outlet temperature of the vis the conventional Claus process. The increased income associ-
flue gases was set to 1200 K and the overall heat transfer coeffi- ated with the co-production of solar H2 is predicted to represent
cient was taken as 50 W/m2 K which is typical for this type of con- annual savings of 45% in comparison to the Claus process. Even
figuration [31,32]. 2971 tubes, 9.8 m in length each, are required to without credit for H2S disposal, solar H2 could be produced at a
reach chemical equilibrium at 1500 K at the end of the tubes. The cost in the range of 6.1–8.6 ¢/kW h. A hybrid natural gas/solar
reactor cost was estimated in 21.6 M$ based on material costs plant design, able to operate 24 h a day, can further reduce the
(tubes, ceramic plates, honeycomb, refractory layer and shell) H2 production cost to 5.8 ¢/kW h at current fuel prices. However,
and assuming a multiplier of 2.0 for installed cost. the increased complexity related with the hybrid reactor plus the
associated specific CO2 emissions of 0.42 kg/kW h negate the eco-
nomic advantage of the hybrid scheme over the solar-only design.
5.2. Results Increasing the quench efficiency and implementing heat recovery
results in higher process efficiencies and, consequently, a reduction
Fig. 8a shows that at the current NG price of 4.5 $/GJ, the hybrid of the heliostat field area, which represents a third of the total di-
plant would lead to an H2 specific cost of 5.8 ¢/kW h, which com- rect capital cost.
pares favorably to the solar H2 produced at 6.9 ¢/kW h. However,
for NG prices higher that 6.0 $/GJ, the solar-only plant is able to
produce H2 at a lower cost than the hybrid plant. It is noteworthy References
that although the difference in H2 specific cost is considerable for [1] Fletcher EA, Noring JE, Murray JP. Hydrogen-sulfide as a source of hydrogen. Int
1 bar, this difference is greatly diminished for lower reactor pres- J Hydrogen Energy 1984;9(7):587–93.
sures (Fig. 8b) and, consequently, higher extent of dissociation. [2] Diver RB, Fletcher EA. Hydrogen and sulfur from H2S-III. The economics of a
quench process. Energy 1985;10(7):831–42.
This is due to the fact that increasing the reactor cost has a stronger
[3] Borton DN, Rogers E. Solar dissociation of hydrogen sulfide: a first step in
producing a renewable fuel. p. 315–9.
[4] Steinfeld A. Solar thermochemical production of hydrogen – a review. Sol
Energy 2005;78:603–15.
[5] Weil ED, Sandler SR, Gernon M. Sulfur compounds, Kirk-Othmer encyclopedia
of chemical technology. John Wiley & Sons, Inc.; 2006.
[6] Eow JS. Recovery of sulfur from sour acid gas: a review of the technology.
Environ Prog 2002;21(3):143–62.
[7] Raymont MED. Hydrocarbon Process 1975;54(7):139.
[8] Noring JE, Fletcher EA. High-temperature solar thermochemical processing
hydrogen and sulfur from hydrogen-sulfide. Energy 1982;7(8):651–66.
[9] Kappauf T, Murray JP, Palumbo R, et al. Hydrogen and sulfur from hydrogen
sulfide – IV. Quenching the effluent from a solar furnace. Energy
1985;10(10):1119–37.
[10] Kappauf T, Fletcher EA. Hydrogen and sulfur from hydrogen sulfide – VI. Solar
thermolysis. Energy 1989;14(8):443–9.
[11] Harvey WS, Davidson JH, Fletcher EA. Thermolysis of hydrogen sulfide in the
temperature range 1350–1600 K. Ind Eng Chem Res 1998;37(6):2323–32.
[12] Karan K, Mehrotra AK, Behie LA. On reaction kinetics for the thermal
decomposition of hydrogen sulfide. AIChE J 1999;45(2):383–9.
[13] Kaloidas VE, Papayannakos NG. Kinetic studies on the catalytic decomposition
of hydrogen sulfide in a tubular reactor. Ind Eng Chem Res 1991;30(2):345–51.
[14] Fukuda K, Dokiya M, Kameyama T, et al. Catalytic decomposition of hydrogen
sulfide. Ind Eng Chem Fundam 1978;17(4):243–8.
[15] Roine A. Outokompu HSC chemistry for windows. Pori, Finland: Outokompu
Research; 1997.
[16] Steinfeld A, Palumbo R. Solar thermochemical process technology,
encyclopedia of physical science and technology. New York: Academic Press;
2001. p. 237–256.
[17] Yogev A, Kribus A, Epstein M, et al. Solar tower reflector systems: a new
approach for high-temperature solar plants. Int J Hydrogen Energy
1998;23(4):239–45.
[18] Segal A, Epstein M. Comparative performances of tower-top and tower-
reflector central solar receivers. Sol Energy 1999;65(4):207–26.
[19] Welford WT, Winston R. High collection nonimaging optics. San
Fig. 8. Comparison between the solar-only plant and the NG/solar-hybrid plant. Diego: Academic Press; 1989.
The curves show the variation of the H2 unit cost when varying: (a) NG price; and [20] Romero M, Buck R, Pacheco JE. An update on solar central receiver systems,
(b) reactor pressure. The solar curve corresponds to the tower-top configuration projects, and technologies. J Solar Energy Eng Trans Asme
with heat recovery and a solar concentration ratio C = 2500. 2002;124(2):98–108.
W. Villasmil, A. Steinfeld / Energy Conversion and Management 51 (2010) 2353–2361 2361

[21] Kolb G, Jones S, Donnelly M, et al. Report SAND2007-3293. Heliostat cost [27] Ober JA. 2006 Minerals Yearbook: sulfur. US Department of the Interior and US
reduction study. Sandia National Laboratories; 2007. Geological Survey; 2008.
[22] Science applications international corporation, phase ii final report. Heliostat [28] Ober JA, Kaiser RC. Sulfur, Mineral Industry Surveys, US Department of the
manufacturing for near-term markets. National Renewable Energy Laboratory, Interior and US Geological Survey, eds.; 2009.
NREL/SR-550-2583; 1998. [29] Maples RE. Tail gas cleanup, petroleum refinery process economics. Pennwell
[23] Meier A, Gremaud N, Steinfeld A. Economic evaluation of the industrial solar Books; 2000. p. 353–354.
production of lime. Energy Convers Manage. 2005;46(6):905–26. [30] Steinfeld A, Spiewak I. Economic evaluation of the solar thermal co-
[24] Erwin D. Process equipment cost determination, industrial chemical process production of zinc and synthesis gas. Energy Convers Manage 1998;39(15):
design. New York: McGraw-Hill; 2002. p. 299–355. 1513–8.
[25] Gary JH, Handwerk GE. Supporting processes, petroleum refining: technology [31] Sinnott RK, Coulson JM, Richardson JF. Heat transfer equipment,
and economics. CRC Press; 2001. Coulson & Richardson’s chemical engineering. Butterworth-Heinemann;
[26] Couper JR, Penney WR, Fair JR, et al. Costs of individual equipment, chemical 2005.
process equipment. second ed. Burlington: Gulf Professional Publishing; 2005. [32] Szczepanski R. Chemical engineering. In: Heaton CA, editor. An Introduction to
p. 719–728. industrial chemistry. Springer; 1996.

You might also like