You are on page 1of 55

Master Semester Project

Cost evaluation of large scale hydrogen


production for the aviation industry

Paul Stadler

Supervisors:
Priscilla Caliandro
Claudio Leonardi
Dr. Jan Van Herle

January 9, 2014

Abstract
The Clip-Air is novel aircraft concept which is based on a single wing design and multiple
passenger and/or cargo vessels. In order to face both economical and environmental issues,
the airplane is considered to fly on liquid hydrogen (LH2). The aim of this study is to assess
the daily fuel requirement of a small Clip-Air fleet and to select the most suited solution for
its generation. A production plant is finally designed in order to evaluate the different costs
related to the cryogenic fuel preparation. Since no greenhouse gases should be released during
the entire cycle, solely conversion methods using renewable energy sources are considered in
this paper.
TABLE OF CONTENTS

Table of contents

1 Introduction 2

2 Fuel assessment 4
2.1 Scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

3 Production technologies 8
3.1 Fossil fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2 Renewable energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.3 Technology selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

4 H2 Production plant design 18


4.1 Copper-chloride cycle step number . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.2 Four-step copper-chloride cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.3 Complementary utilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.4 Energy requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.5 Production plant cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

5 Power supply plant design 33


5.1 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 Solar irradiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.3 Plant setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.4 Cost evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

6 Delivery costs 43
6.1 Delivery methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.2 Methods selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.3 Delivery cost assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

7 Conclusion 47
7.1 Hydrogen cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
7.2 Further work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

8 References 49

List of Figures 51

List of Tables 52

Appendix A Fuel assessment i


A.1 A320 fuel consumption sheet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Appendix B Power plant design ii


B.1 Capital cost reference values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

1
1 INTRODUCTION

1 Introduction
Since its beginning in the last half century, commercial air travel has continuously grown all
over the globe, developing itself into a safe, comfortable and fast type of transportation. The
economic liberalization of the airline industry which has gradually spread through all countries
since the 1980s, has generated an aggressive market with low entry barriers and an extremely
high level of competition. The resulting drop in airfares enabled some companies to target a
wider range of customers, making air travel more accessible for the general public. The following
democratization has led to a further strong increase of demand, making air travel one of the
most popular means of transportation existing.
Nevertheless, regarding today’s environmental issues, the aviation industry is constrained to
react; increasing political and economical pressure has pushed companies to produce more and
more fuel-efficient planes. Recent designs such as the Boeing 787 Dreamliner or the Airbus 350
have shown the improvements feasible by introducing composite materials in most parts of the
fuselage and wings.
Nevertheless, these innovative improvements are considered to be insufficient for agencies
such as NASA and the Advisory Council for Aeronautic Research in Europe (ACARE). Indeed,
recent forecasts from the International Air Transport Association (IATA) showed that the future
growth in air travel traffic is going to skyrocket1 and therefore, despite the latest efficiency
improvements, leading to an overall increase of fuel demand. For these ecological and economical
reasons, ACARE has targeted the following objectives for the year 20202 :

• Lower CO2 emissions by 50 %


• Lower NOx emissions by 80 %
• Increase transport capacity
• Increase aircraft safety design
• Reduce airline noise pollution by 50 %

Regarding these objectives, the novel aircraft concept Clip-Air has been designed. Unlike
traditional airplane designs, it is composed of a single wing under which up to three capsules
can be attached. Due to the lower drag of the wing shape and the high modularity resulting
from its unique sizing system, significant fuel savings can been made by operating such type of
airplane. Indeed, the aircraft capacity can be instantaneously modified by removing or adding
a passenger/cargo capsule just before the flight.

Figure 1.1: The Clip-Air aircraft


1
http://www.iata.org/pressroom/pr/pages/2012-12-06-01.aspx
2
http://infoscience.epfl.ch/record/152404/files/Bilge_AGIFORS_2010.pdf

2
1 INTRODUCTION

In order to further reduce the production of toxic exhaust gases, the Clip-Air project is
studying the feasibility of using liquid hydrogen (LH2) as a substitute to the current Jet-A
fuel. Since hydrogen is an unstable element, it still presents strong challenges for many of
its production steps, its transportation and storage. This report reviews the different existing
technologies for hydrogen generation and tries to give a preliminary cost evaluation for a specific
chosen case, when used in the airline industry.

3
2 FUEL ASSESSMENT

2 Fuel assessment
The objective of this paper is to evaluate the costs of hydrogen production for commercial
air travel use. Instead of analyzing various existing technologies and processes for hydrogen
generation, a particular scenario and the related fuel consumption has been defined in order to
selected the most suitable solution. The aim of this chapter is to assess the daily amount of
liquid hydrogen (LH2) required for a given Clip-Air fleet size and a specific type of operation.

2.1 Scenario
In order to predict the social and economical impact of the Clip-Air concept in the airline
industry, several scenarios have been developed. Specific simulations made by Atasoy et al. [1]
have shown the different types of performances achievable through the Clip-Air system. By
varying parameters such as the network type, fleet and plane sizes, the study brought to light
significant advantages and drawbacks of this novel aircraft concept.
As the Clip-Air plane does not exist yet, major assumption concerning the operating costs
(fuel consumption, crew cost, airport and air navigation charges) have been made by considering
similarities of an A320 since their fuselages are almost identical. By using the weight difference
between each Clip-Air configuration and the respective capacity of A320, the proportional fuel
consumption increase/decrease can be defined. As presented in Table 2.1, Clip-Air is lighter
than the respective number of A320 planes when flying with three capsules but heavier when it
is flying with less capacity. Similarly to the fuel consumption approximation, the navigation and
crew charges were determined by using analogies with existing cost functions, although they are
not relevant for this study.
Table 2.1: Clip-Air and A320 specific weight [1]

Clip-Air A320
Total seat capacity [-] 450 150
Turbofans [-] 3 2
Maximum takeoff 1 plane/capsule 139 (+78%) 78
weight [mt] 2 planes/capsules 173.5 (+11%) 156
3 planes/capsules 208 (-11%) 234

2.1.1 Data extrapolation


The fuel utilization of a standard A320-100/200 is assessed by using a classical Jet-A consump-
tion sheet (Appendix A.1). By selecting significant operation parameters like flight distance,
wind speed and landing weight, the required fuel amount can be defined through the different
curves. Nevertheless, due to the narrow ranges of significant parameters, the application of these
consumption graphs is quite limited. Indeed, the flight distance is bounded between 1600 and
2900 nmi (2963 and 5370 km respectively) whereas the average value in Europe is about 1200
nmi [9]. Hence, an extrapolation of the existing consumption curve must be performed in order
to include shorter journeys into the fuel assessment.
In order to approach the lower consumption area, a polynomial function is generated by
using a non-linear least-square fit of the available data. The resulting curve is represented in
Figure 2.1, the initial lower bound (gray line) being decreased to 500 nmi. Considering the
magnitude of the mean of squared error (Table 2.2), the approximate graph can be validated
for the desired domain.

4
2 FUEL ASSESSMENT

Figure 2.1: A320-100/200 fuel consumption (FC) graph fitting for the respective flight distance (FD)
assuming a landing weight of 57 [mt]

Table 2.2: Fitted function properties

Error analysis
Degrees of freedom 6
Sum of squared residuals 4207.46
Mean of squared error 701.343

2.1.2 Flight assessment


The A320-100/200 fuel consumption being evaluated, the respective Clip-Air Jet-A utilization
can be defined by using the relative weight difference between both aircrafts. Since this study
is only considering the European air transport market, the related average flight distance, i.e.
1200 nmi (≈1930 km), is selected as a reference value for medium-haul flights. Furthermore,
a short-haul flight is defined in order to include domestic and small transnational flights. The
respective length is set to 600 nmi (≈965 km), half of a medium-haul flight. Table 2.3 presents
the resulting fuel consumption data for each Clip-Air configuration and the respective A320-
100/200 seat capacity.

Table 2.3: Fuel consumption of Clip-Air and A320-100/200

Short-haul flight Medium-haul flight


A320-100/200 Clip-Air A320-100/200 Clip-Air
Fuel Jet-A 4350 7740 7030 12510 1 plane/capsule
consumption [kg] 8700 9660 14060 15610 2 planes/capsules
13050 11610 21090 18770 3 planes/capsules

2.1.3 Network type


Since the reference flight distances have been determined, a network structure is selected based
on previous simulation results. Indeed, the Atasoy et al. study [1] compared three different

5
2 FUEL ASSESSMENT

operation types and the resulting economical impact generated by the utilization of a Clip-
Air fleet. The peer-to-peer network showed itself to be the most suitable solution in terms of
revenues, transported passengers and aircraft usage. In this type of network, the airports are
connected pairwise which enables to change the number of capsules after each flight and hence,
a more efficient capacity management. Consequently, a profit increase of 4.1% was achieved by
carrying 2.8% more customers with 21.3% less seat capacity.

Figure 2.2: Short-haul and medium-haul flight ranges from LSGG

In light of these results, a fleet of five Clip-Air wings operating under a similar network
scheme is considered in this study (Table 2.4). The number of daily destinations is set to
10 which corresponds to two return flights, one short-haul and one medium-haul, per aircraft
respectively. Since the airline hub is the Geneva International Airport (LSGG), the selected flight
range enables thus to target most of the Western European cities which are highly popular and
profitable (Figure 2.2). Furthermore, in order to assess the maximum fuel requirements, each
wing is assumed to fly with full seat capacity (i.e. with three capsules) although the demand
varies strongly during the day and between destinations.

Table 2.4: Scenario characteristics

Fleet properties
Number of Clip-air wings [-] 5
Number of destinations [day−1 ] 10
Consumption [mt day−1 ]
Sum of short-haul return flights [day−1 ] 5 116.1
Sum of medium-haul return flights [day−1 ] 5 187.7
Total Jet-A [day−1 ] 10 303.8

6
2 FUEL ASSESSMENT

2.1.4 Hydrogen assessment


In order to determine the daily utilization of LH2, an energy balance is performed by using the
defined Jet-A requirements for the given scenario (Table 2.4). Hence, the turbofan engines are
assumed to consume the same amount of energy for both fuels although this hypothesis might
be pretty disputable [12]. Indeed, the performance analogy between cryogenic fuel and standard
kerosene combustion jet engines has never been validated. Moreover, the lower density of liquid
hydrogen leads to lower total aircraft weight and thus, influences the consumption assessment.
However, due to lack of time and information, this assumption is constant throughout the study.
The fuel proprieties and hydrogen assessment are presented in Table 2.5.

Table 2.5: Hydrogen assessment

Fuel properties [12]


Jet-A LH2
Specific density ρ at boiling point [kg/m3 ] 790-808 71
Specific energy a [MJ/kg] 43.2 120
Fuel consumption Jet-A LH2
Short-haul return flights [mt day−1 ] 116.1 41.8
Medium-haul return flights [mt day−1 ] 187.7 67.6
Total weight [mt day−1 ] 303.8 109.4
Total volume [m3 day−1 ] 380.2 1540.8
Total energy [MWh day−1 ] 3646

2.2 Conclusion
This chapter presented the different elements and assumptions considered throughout the sce-
nario elaboration. Based on various economical simulations, specific parameters such as network
type, operation strategy and fleet size were selected. Consequently, the daily utilization of liquid
hydrogen for the resulting model amounts to 109.4 mt. A 10% margin is added to this value in
order to take in account losses due to leakage, transportation and boil-off, hence leading to a
total requirement of 120 mt day−1 (4000 MWh day−1 ).

7
3 PRODUCTION TECHNOLOGIES

3 Production technologies
Since hydrogen is an extremely abundant element, various feedstocks can be used for its gen-
eration. Nowadays, it is mostly generated by using processes involving fossil fuels such as coal
and natural gas. Nevertheless, considering both the ecological impact trough greenhouse gases
and the limited amount of available resources, the present production paths are highly unsus-
tainable. Consequently, alternative technologies and processes using renewable energy sources
must be considered in order to adress both economical and environmental issues.

3.1 Fossil fuels


Presently, most hydrogen is produced by using fossil fuels such as natural gas, coal and other
hydrocarbon fuels. The chemical processes required for hydrogen generation form these feed-
stocks are highly developed and well-known today. Due to the mature technology and the strong
market position of fossil fuel based hydrogen production, the resulting costs are extremely low
relative to other sources.

3.1.1 Natural gas


In order to convert natural gas (NG) to hydrogen, several different chemical processes can be
used [25] such as:
• Steam reforming
• Partial oxidation
• Autothermal reforming
In steam reforming, methane is combined in an endothermic process with steam in order to
generate a synthetic gas which is largely composed of carbon monoxide and hydrogen (3.1). The
heat required for the methane-steam reaction is often supplied through combustion of some of
the available natural gas. Finally, the CO present in the syngas is further converted to carbon
dioxide and hydrogen by using the CO shift reaction (3.2).
CH4 + H2 O −−→ CO + 3 H2 (3.1)
CO + H2 O −−→ CO2 + H2 (3.2)
Partial oxidation is an exothermic process where methane is partially combusted in a reformer,
generating a differently composed synthesis gas. Again, the carbon monoxide is further converted
to CO2 and H2 by using (3.2).
CH4 + 21 O2 −−→ CO + 2 H2 (3.3)
The last conversion method, autothermal reforming, is a combination of steam reforming and
partial oxidation. The complete process is exothermic and hence do not require any external
heat source. The efficiency of the whole process is nevertheless lower since the produced gases
(especially the H2 ) need purifying and thus, involves higher costs [25].

Cost Since methane steam reforming and partial methane oxidation are well-developed pro-
cesses, the cost of hydrogen production is highly depend on the cost of natural gas. The following
equation proposed by [26] presents this relationship as follows:
H2 [$/kg] = 0.286 × NG price[$/MMBtu] + 0.15 (3.4)
Given the NG price of 13.2$/MMBtu (45$/MWh) for 2013 in Switzerland [29], the hydrogen
production cost reaches 3.93$/kg.

8
3 PRODUCTION TECHNOLOGIES

3.1.2 Coal
Coal can be converted into hydrogen by a variety of endothermic gasification processes; fixed
bed, fluidized bed and entrained flow (Figure 3.1). The selection of the appropriated technology
depends on various parameters like the reactor size, the desired gas utilization and quality.
Large scale applications such as an IGCC (Integrated Gasification Combined Cycle) power plant
require a high and stable conversion rate and hence, fluidized bed or entrained flow gasifiers
are the most suitable solutions. Furthermore, entrained flow gasification, compared to fixed
and fluidized bed gasification, takes place at temperatures above the ash fusion point [24] and
consequently, generates a superior gas quality.

Figure 3.1: Gasification processes for coal [24]

In a typical gasification reaction, the injected fuel and gasification agent are converted into a
mixture mainly composed of carbon monoxide and hydrogen (3.5). Similarly to steam reforming,
the CO can be further processed by using the water-gas shift reaction (3.2) in order to form
CO2 and H2 .
C + H2 O −−→ CO + H2 (3.5)

Cost Hydrogen generation from coal gasification is a mature technology and commercially
available. However, the cost of the different plants varies widely. While sophisticated designs
such as combined cycles (CC) and advanced facilities using solid oxide fuel cells (SOFC) pro-
vide the lowest hydrogen cost (especially due to electricity co-production), smaller cycles using
heat recovery steam generators (HRSG) to power the plant require a lower investments costs
[3]. In order to reduce greenhouse gas emissions, these plants can be equipped with a carbon
sequestration system, increasing further the production costs.
Rutkowski et al. [18] evaluated the hydrogen production cost from coal gasification. In-
cluding a sequestration system, the plant investment expenses reach US$ 612.3 millions and
produces 255400 kg/day of hydrogen at a cost of 1.83$/kg.

9
3 PRODUCTION TECHNOLOGIES

3.2 Renewable energies


As the global fossil fuel demand increases, alternatives to the conventional path of hydrogen
production must be considered in order to face both economical and environmental issues. In-
deed, fossil fuel costs are continuously growing while recent technological improvement and
developments have decreased the cost of using renewable energy sources. Hence, this section
presents and compares the different processes able to use alternative energy sources for hydrogen
generation in order to select the most suitable solution.

3.2.1 Water electrolysis


The electrochemical reaction of water electrolysis (3.6) is a well-established technique which
consists of splitting water into its single components by using electrical energy. Therefore, if the
energy source produces no carbon compounds for the required power generation, then the entire
production process is carbon free.

H2 O −−→ H2 + 21 O (3.6)

Alkaline electrolysis Alkaline electrolysis is well-developed technology which uses an aqueous


solution (KOH or NaOH) as electrolyte. During the process (Figure 3.2), water at the cathode
consumes electrons to form hydrogen. Hydroxide ions are migrating through the solution towards
the anode at which they release electrons and consequently, close the cycle. A diaphragm which
separates both electrodes and their respective reaction enables to collect the generated gases
afterwards.
The half and overall reactions of the alkaline electrolysis can be written as:

Cathode 2 H2 O+ + 2 e− −−→ H2 + 2 OH− (3.7a)

Anode 2 OH− −−→ 1


2 O2 + H2 O + 2 e− (3.7b)
1
Sum H2 O −−→ 2 O2 + H2 (3.7c)

Alkaline electrolysis is a commercially mature technology and globally present on the hydro-
gen production market. The high gas production rate, the long stack lifetime and the relatively
high efficiency (Table 3.1) make it an interesting candidate for large scale and stationary appli-
cations.

Cathode - Anode +

H2 1
2 O2
OH−

2H+

Diaphragm

Figure 3.2: Operating principle of an alkaline electrolyzer

However, despite these attractive characteristics, some drawbacks are usually associated with
alkaline electrolysis [5]. Indeed, the diaphragm is not entirely impermeable to the generated gases

10
3 PRODUCTION TECHNOLOGIES

which highly influences the efficiency of the system since the oxygen will react with the hydrogen
present at the cathode to form water again. Especially at low operating points, cross-diffusion
represents a serious safety issue due to the increasing concentration of hydrogen on the anode
side which might reach dangerous levels. A second disadvantage of alkaline electrolysis are the
strong ohmic losses created by the separation membrane and the aqueous electrolytic solution,
therefore limiting the current density throughout the stack.

Table 3.1: State of the art for specifications of Alkaline and PEM electrolysis [5]

Alkaline electrolysis PEM electrolysis


Operating temperature [C] 40-90 20-100
Current density [mA/cm2 ] 0.2-0.4 0.6-2.0
Cell voltage [V] 1.8-2.4 1.8-2.2
Power density [mW/cm2 ] <1 < 4.4
Voltage efficiency (HHV) [-] 0.62-0.82 0.67-0.82
H2 production rate [Nm3 /hr] < 760 < 10
Lower partial load [-] 0.2-0.4 0-0.1
Lifetime stack [hr] < 90,000 < 20,000
Lifetime system [y] 20-30 10-20

Polymer electrolyte membrane electrolysis PEM electrolyzer use a solid polymer mem-
brane as electrolyte which highly improves the conductivity issues related to the utilization of a
liquid electrolyte. During operation (Figure 3.3), the water is decomposed at the anode where it
releases electrons in order to form oxygen and hydrogen ions. After crossing the separation mem-
brane, the protons are combined with the supplied electrons to produce the desired hydrogen
gas which can be easily collected afterwards.
The half and overall reactions of PEM electrolysis can be written as

Cathode 2 H+ + 2 e− −−→ H2 (3.8a)

Anode H2 O −−→ 1
2 O2 + H+ + 2 e− (3.8b)
1
Sum H2 O −−→ 2 O2 + H2 (3.8c)

The positive aspects of PEM electrolysis are mainly related to its solid polymer electrolyte
type and thickness [5]. Indeed, compared to alkaline electrolysis, the thin electrolyte provides
an excellent proton conductivity and thus, higher achievable current densities which lead to a
reduction in operating and equipment costs. Furthermore, the low cross-diffusion rate enables
PEM electrolyser to work under a wider range of operating points and to adapt to the actual
production demand.
However, since the membrane impermeability decreases with pressure, thicker and more
resistant electrolytes are required at higher operating conditions (above 100 bar [5]) which
therefore increases electrical resistance. Additionally, due to the harsh environment in a PEM,
only a few expensive materials can be used for the different cell components, leading to a higher
system cost [5]. With short membrane lifetime and low production capacity (Table 3.1), current
PEM electrolysers are not as developed as alkaline electrolyzers [25] and less attractive for large
scale applications.

11
3 PRODUCTION TECHNOLOGIES

Cathode - Anode +

H2 1
2 O2
H+

H2 O

Membrane

Figure 3.3: Operating principle of a PEM electrolyzer

Solid oxide electrolysis Based on the solid oxide fuel cell principle, solid oxide electrolyzer
cells (SOEC) operate at high temperature, typically between 973 and 1273 K where the reaction
becomes more reversible [25]. Since the electrical energy required to split water decreases with
temperature (Figure 3.4), part of the heat available for electricity generation can directly be used
by the system. Indeed, the molar Gibbs energy of reaction ∆G falls from 1.23 V at ambient
conditions to about 0.96 V at 1123 K [4] which represents a drop of 22%. This important amount
of energy can be provided by an available heat source (nuclear, solar, geothermal), avoiding thus
a penalizing heat-to-electricity conversion loss

Figure 3.4: Energy of reaction (∆ H, red) and Gibbs energy of reaction (∆ G, gray) in function of the
steam inlet temperature

During the process (Figure 3.5), superheated steam is injected into the cathode side where
it is decomposed into hydrogen and oxide ions. After passing through the electrolyte, generally
yttria-stabilized zirconia [4], the oxide ions release electrons at the anode in order to form oxygen.
The half and overall reactions of solid oxide electrolysis can be written as

Cathode H2 O + 2 e− −−→ H2 + O2− (3.9a)

Anode O2− −−→ 1


2 O2 + 2 e− (3.9b)
1
Sum H2 O −−→ 2 O2 + H2 (3.9c)

Since their introduction in the 1980s, solid oxide electrolyzers have gathered much interest
around the world, especially due to their huge potential for hydrogen mass production. Prelim-

12
3 PRODUCTION TECHNOLOGIES

inary lab-scale studies have shown the strong advantages of high temperature water-splitting
[5]:

• High efficiency - the electrical-to-hydrogen conversion efficiency can exceed 100% depend-
ing on the system operation mode (exothermal, thermoneutral and endothermal) [4]
• High pressure operation
• No noble materials required

Despite the fact that SOECs are currently available on the market, further research and devel-
opment are necessary in order to meet the required performances for sustainable and profitable
hydrogen generation. Indeed, issues regarding durability, manufacturing and production costs of
the different ceramic components still need to be solved [5]. However, considering that the tech-
nology is not commercially mature and economies of scale have not been achieved yet, SOECs
remain an extremely attractive conversion process in the long term.

Cathode - Anode +

2H2 O2
O−2

2H2 O

Solid Oxide

Figure 3.5: Operating principle of a SOEC

13
3 PRODUCTION TECHNOLOGIES

3.2.2 Thermochemical water splitting


Thermochemical water splitting is a thermally driven process which converts water into hydrogen
(and oxygen) through different chemical reactions. Since no electrical energy is required, the high
temperature heat source (solar, nuclear) can be directly used for hydrogen generation and hence,
improve the overall conversion efficiency of the cycle. Indeed, considering typical electrolyzer
and heat-to-electricity efficiency values of 80% and 35% respectively, the conversion heat-to-
hydrogen ratio only reaches 28%. Over the last decades, about 200 thermochemical cycles have
been identified from which three have been the most studied [14]:
• Sulfur - Iodine
• Sulfur hybrid
• Hybrid copper-chloride

Sulfur - Iodine The sulfur-iodine cycle is composed of three distinct sections and was first
presented by General Atomics in the 1970s [13]. Based on the HHV of hydrogen, the heat-to-
hydrogen efficiency of sulfur-iodine cycles is predicted to reach over 40% (34% if considering
the hydrogen LHV), depending on peak temperature of decomposition reaction [14]. The main
reactions can be written as:
Section I Bunsen reaction

I2 + SO2 + 2 H2 O −−→ 2 HI + H2 SO4 ∆HT =373K = −126 kJ/mol (3.10a)

Section II Sulfuric acid decomposition

H2 SO4 −−→ SO2 + H2 O + 12 O2 ∆HT =1123K = 373 kJ/mol (3.10b)

Section III Hydrogen iodine decomposition

2 HI −−→ I2 + H2 ∆HT =723K = 12 kJ/mol (3.10c)

H2 O

Section I Section II
I2
I2 + SO2 + 2 H2 O −−→ 2 HI + H2 SO4 2 HI −−→ I2 + H2
H2 O

Separator
H2 O, I2 , HI H2
SO2 , H2 O

Section III H2 SO4

H2 SO4 −−→ SO2 + H2 O + 21 O2


O2

Figure 3.6: Operating principle of a sulfur-iodine cycle

14
3 PRODUCTION TECHNOLOGIES

The water input is injected into the Bunsen reactor which operates at 100◦ C and releases
heat which can be recovered and used elsewhere in the cycle (Figure 3.6). The output stream is
then separated into two streams, the aqueous sulfuric acid flow and the hydrogen iodine solution.
In the second section, the sulfuric acid is recycled at 850◦ C by being decomposed into sulfur
dioxide and water which are then fed into the Bunsen reactor again. Finally, the hydrogen iodine
is split into the desired hydrogen and iodine at 450◦ C in the last section of the cycle.

Sulfur Hybrid The sulfur acid hybrid thermochemical cycle (SAHT) is composed of only two
distinct sections and was first developed by Westinghouse [14]. The SAHT cycle is ”hybrid” due
to the second production step (3.11b) which consists of an electrochemical reaction. Similarly
to the sulfur-iodine cycle, the SAHT cycle is expected to be very efficient (over 45% based on
the HHV or 38% if considering the LHV of hydrogen [14]) which makes it extremely attractive
for large scale hydrogen production. The main reactions can be written as follows:

Section I Sulfuric acid decomposition

H2 SO4 −−→ SO2 + H2 O + 12 O2 ∆HT =1123K = 373 kJ/mol (3.11a)

Section II Electrochemical hydrogen production

2 H2 O + SO2 −−→ H2 + H2 SO4 (3.11b)

The water input is injected into the sulfur dioxide electrolyzer which operates at about 373
[K] at an overall cell voltage of 0.67 [V] [14]. The required electrical energy is indeed much lower
than for conventional water splitting through electrolysis (see Section 3.2.1) increasing hence,
the hydrogen production efficiency. The generated by-product, sulfuric acid, is then recycled by
using the decomposition reaction in order to form sulfuric dioxide again.

Hybrid Copper Chloride The hybrid copper-chloride cycle is a recent thermochemical cycle
which exists in various types, depending on the different intermediate reaction steps. The
number of steps, which can vary from two to five, highly influences large-scale design challenges
and the overall efficiency of the setup. Indeed, multiple studies have shown major economical
and technological advantages and drawbacks related to the increase/decrease of process steps.
Based on an idealized process assumption, Lewis et al [16] predicted a five step cycle efficiency
of 50% (LHV). The main reactions occurring in a five step Cu-Cl cycle can be written as:

Step I Chlorination

2 Cu(s) + HCl(g) −−→ 2 CuCl(molten) + H2 (g) at 723 K (3.12a)

Step II Disproportionation (electrolysis)

4 CuCl(aq) + 2 Cu(s) −−→ 2 CuCl2 (aq) at 303-353 K (3.12b)

Step III Drying

CuCl2 (aq) + nf H2 O(l) −−→ CuC2 l(s) · nh H2 O(s) + (nf −nh )H2 O at 303-353 K (3.12c)

Step IV Hydrolysis

2 CuCl2 · nf H2 O(s) + H2 O(g) −−→ Cu2 OCl2 (s) + 2 HCl(g) + nh H2 O at 648 K (3.12d)

Step V Decomposition

Cu2 OCl2 (s) −−→ 2 CuCl(molten) + 12 O at 803 K (3.12e)

15
3 PRODUCTION TECHNOLOGIES

where nf ≥ 7.5 is the amount of water in the aqueous solution of CuCl2 and nh = 0-4 the water
quantity in the dry CuCl2 hydrate.

3.3 Technology selection


The different hydrogen generation processes being presented, the most suited production type
need to be selected in order to design and evaluate both the thermal and electrical requirements
of the plant. Table 3.2 synthesizes the different benefits and drawbacks of each conversion type
described throughout this chapter. Since the production should be entirely driven by renewable
energy sources, conventional conversion processes from fossil fuel based feedstocks, such as coal
and natural gas, are not considered in this study.

Table 3.2: Major characteristics for each hydrogen production process

Alkaline electrolysis PEM SOEC


Commercialized Commercialized Lab-scale/Commercialization
Advantage Long stack lifetime High current densities High efficiency
High production rate Wide operating range High operating pressures
High safety No noble materials
Drawback Operating range High material costs High graded heat
Low current densities Short stack lifetime Production costs
Low production rate Ceramic manufacturing
Stack lifetime

Sulfur-iodine cycle SAHT cycle Cu-Cl cycle


Small pilot plant Well developed In development
Advantage No electrical input3 High efficiency Medium graded heat
High efficiency Safe
Lab-scale
demonstrated
Drawback High graded heat High graded heat Many technological
Large flow rates Safety issues challenges
Safety issues Electrolyzer development
(Crossover issues)

Given the objective of this paper, following parameters need to be considered during the
technology selection:

• Large scale production (see Section 2.2)


• High efficiency
• Long system lifetime
• No important safety issues
• Low operating and capital costs

Regarding the different characteristics of each conversion process, the copper-chloride ther-
mochemical cycle is selected. Indeed, both the sulfur-iodine and the SAHT cycle are rejected
due to safety issues (large scale handling of hazardous chemicals such as H2 SO4 , SO4 , HI, I2 ) and
3
Power to drive pumps, compressors and other utilities not included

16
3 PRODUCTION TECHNOLOGIES

the high graded heat requirements for the sulfuric acid decomposition. Since only concentrated
solar energy is able to generated these high temperatures (nuclear power does not fall into the
category of renewable energy sources), a thermal storage utility need to be designed in order to
enable ”around-the-clock” operation. However the required temperature exceeds the operation
limit of the nitrate salt (about 570◦ C) which are currently used as heat storage medium. This
issue also applies to the solid oxide electrolysis which requires the preparation of superheated
steam and hence, SOECs are eliminated from the technology list.
Finally, both alkaline and PEM electrolysis are rejected due to their lower conversion effi-
ciency and short stack lifetime respectively. Consequently, solely the Cu-Cl cycle is considered
for hydrogen production throughout the following chapters. It is important to note that recent
technologies such as biophotolysis, photolysis, high temperature decomposition and biomass con-
version have not being included in the down-selection due to lack of time or/and low potential
for large scale hydrogen production.

17
4 H2 PRODUCTION PLANT DESIGN

4 H2 Production plant design


The production technology being selected, the hydrogen conversion plant needs to be mod-
eled in order to evaluate its minimum energy requirements (MER) and capital expenses. This
chapter presents the different steps and assumptions made during the design process and cost
calculations.

4.1 Copper-chloride cycle step number


As presented previously (Section 3.2.2), the copper-chloride cycle exists in different types which
are mainly characterized by the number of major steps and the reaction taking place. The
five-step cycle (3.2.2) is considered to be the standard form of the Cu-Cl cycle family and often
used as a baseline for comparison agmong cycles with lower step numbers. The main advantages
and drawbacks of decreasing the step number are [30]:

• Advantages
– Fewer equipment material challenges
– Less steps, less equipment
– Reduced complexity
• Disadvantages
– Higher heat grade and intensity
– More equipment material challenges
– Lower reaction yield (undesirable side products)
– Lower efficiency

Regarding the multiple benefits and drawbacks of varying the step number, only the four-
step copper-chloride cycle will be studied throughout the following chapters since it represents
the best trade-off between thermal efficiency and viability. The main reactions occurring in a
four-step cycle are presented in Table 4.1.

Table 4.1: Chemical reactions in a four-step Cu-Cl cycle

Step Reaction Temperature [K]


I 2 CuCl(aq) + 2 HCl(aq) −−→ 2 CuCl2 (aq) + H2 (g) <373.15
II CuCl2 (aq) + nf H2 O(l) −−→ CuCl2 · nh H2 O(s) + (nf −nh )H2 O 303.15-353.15
III CuCl2 · nh H2 O(s) + H2 O(g) −−→ Cu2 OCl2 (s) + HCl(g) + nh H2 O(g) 673.15
1
IV Cu2 OCl2 (s) −−→ 2 CuCl(molten) + O(g)2 823.15
Note: nf > 7.5 and nh = 0-4

4.2 Four-step copper-chloride cycle


The four-step copper-chloride cycle is composed of one electrochemical and three thermochemical
reaction: (I) electrolysis, (II) crystallization, (III) hydrolysis and (IV) oxychloride decomposi-
tion. The following section will present each step of the conversion process and their respective
technological challenges. Figure 4.1 presents the operating principle of the selected Cu-Cl cycle
type.

18
4 H2 PRODUCTION PLANT DESIGN

Section I Electrolysis H2 (g)


CuCl(s)
2 CuCl(aq) + 2 HCl(aq) −−→ 2 CuCl2 (aq) + H2 (g)
H2 O HCl(aq)

Section II Crystallizer CuCl2 (aq)

CuCl2 (aq) + nf H2 O(l) −−→ CuCl2 · nh H2 O(s) + (nf −nh )H2 O

Section III Hydrolysis CuCl2 (s)

2 CuCl2 · nh H2 O(s) + H2 O(g) −−→ Cu2 OCl2 (s) + 2 HCl(g) + nh H2 O(g)

Section IV Decomposition Cu2 OCl2 (s) H2 O

Cu2 OCl2 (s) −−→ 2 CuCl(molten) + 21 O(g)

O2 (g)

Figure 4.1: Operating principle of a four-step copper chloride cycle

4.2.1 Electrolysis
In the first section, hydrogen is produced from an electrochemical reaction of hydrogen chloride
and copper(I) chloride. At the anode, the copper ion Cu(I) is oxidized to Cu(II) while at the
cathode, hydrogen ions H+ are reduced in order to form the desired product. The half and
overall reactions taking place in the electrolyzer stack (Figure 4.2) can be written as:

Anode 2 HCl(aq) + 2 CuCl(aq) −−→ 2 CuCl2 (aq) + 2 H+ (aq) + 2 e− (4.1a)

Cathode 2 H+ (aq) + 2 e− −−→ H2 (g) (4.1b)

Sum 2 HCl(aq) + 2 CuCl(aq) −−→ 2 CuCl2 (aq) + H2 (g) (4.1c)

The copper(I) chloride which is dissolved in a hydrogen chloride solution, is pumped to the
anode side of the electrolyzer, whereas an aqueous HCl solution is fed to the cathode section.
Hydrogen ions migrate through the membrane, from the anode to the cathode where they react
with the available electrons in order to form hydrogen gas. The produced aqueous CuCl2 is then
collected and send to the crystallization unit.

19
4 H2 PRODUCTION PLANT DESIGN

2e−

Anode + Cathode -

CuCl2 (aq) + H+ (aq) H2 (g)


H+

CuCl(aq) + HCl(aq) 2 H+ (aq)

Proton conducting membrane

Figure 4.2: Operating principle of the CuCl/HCl electrolysis

4.2.2 Crystallization
The anode outlet stream flows from the electrolyzer system to the next section which is the crys-
tallization unit. Wand et al. [30] presented the different benefits and drawbacks of two existing
drying methods which are considered for this process step; spray drying and crystallization. In
order to classify each solution, two important parameters have been identified.
The first variable, the amount of water in the aqueous solution of CuCl2 is represented by nf
(Table 4.1) which has a minimum value of 7.5 if the solution is saturated. The second variable
nh denominates the hydrate water quantity in the dry copper(II) chloride. It can vary from
0, when the CuCl2 exists in anhydrous form after drying, to 4. Obviously, the value of nh
highly depends on major parameters such the drying method and their operating conditions
(temperature, carrier gas velocity).

Table 4.2: Heat grade and quantity required for precipitate drying [30]

Parameter Value
Heat required ∆H [kJ/kmol H2 ] 122
Heat grade ∆T [◦ C] 30 → 100

Given the lower heat grade and heat amount required, the crystallization technology has
been selected for this study. The energy requirements of the desired crystallization unit are
reported in Table 4.2.

4.2.3 Hydrolysis
After exiting the crystallizer, the CuCl2 particles are heated up to 673 K and, in addition to
superheated steam at same temperature level, injected into the hydrolysis reactor (Figure 4.3).
Due to the large pressure drop of 23 bar, the CuCl2 stream forms a free jet at the top of the
spray reactor. Since the jet expands, it aspirates the superheated steam into the jet and hence,
highly improves the heat and mass transfer between both reactants [20]. The solid Cu2 OCl2
settles at the bottom of the hydrolysis reactor where it is removed afterwards. Finally, the
hydrogen chloride and the remaining amount of steam leave the reactor in form of gas before
being cooled for separation.

20
4 H2 PRODUCTION PLANT DESIGN

CuCl (24 bars)

Steam HCl, steam

Cu2 OCl2

Figure 4.3: Hydrolysis spray reactor [20]

Nevertheless, in order to achieve stoichiometric yield of HCl or Cu2 OCl2 , the amount of
H2 O injected into the hydrolysis reactor must largely exceed the stoichiometric steam quantity.
Indeed, past studies [30] have shown that an optimal steam-to-copper (S/Cu) ratio of 17 is
required in order to reach a conversion yield higher than 95%. Hence, the main objective of
CuCl2 hydrolysis step development is to maximize the copper oxychloride production rate while
reducing as much as possible the steam input. The energy requirements of the hydrolysis unit
are reported in Table 4.3.

Table 4.3: Heat grade and quantity required for the hydrolysis reaction [21, 20]

Parameter Value
Heat required ∆H [kJ/kmol H2 ] 116.6
Heat grade T [◦ C] 375

4.2.4 Decomposition
The generated copper oxychloride flows from the hydrolysis reactor outlet to the decomposition
reactor where it is decomposed into copper(I) chloride and oxygen (Figure 4.4). Before entering
the reactor, the Cu2 OCl2 is heated to 823 K, the temperature at which the decomposition
reaction occurs. Similarly to the hydrolysis step, the oxygen leaves the reactor as gas (at the
top) whereas the molten CuCl spills over a weir in order to be collected and send for recycling
afterwards. The energy requirements of the oxychloride decomposition unit are reported in
Table 4.4.
Table 4.4: Heat grade and quantity required for the decomposition reaction [21, 20]

Parameter Value
Heat required ∆H [kJ/kmol H2 ] 129.2
Heat grade T [◦ C] 550

21
4 H2 PRODUCTION PLANT DESIGN

O2

Weir

CuCl

Cu2 OCl2

Figure 4.4: Oxychloride decomposition reactor [20]

4.3 Complementary utilities


In addition to the copper chloride cycle, complementary utilities regarding the water preparation
and hydrogen liquefaction need to be included during the energy and cost assessment. Indeed,
assuming a free access to a large water source such as the Mediterranean Sea, the water is
pumped into a first unit which desalinates and purifies it before sending it to the production
plant (Figure 4.5). After conversion, the generated hydrogen is sent to the liquefaction utility
where the product is prepared for shipping.

Overall production plant

sea purified
water water H2 (g) H2 (l)
Desalination Cu-Cl cycle Liquefaction

E˙+ Q˙+

CSP plant

Figure 4.5: Plant functional sheet

Finally, a concentration solar power plant is designed next to the production facility in order
to satisfy the electrical and thermal energy requirements of the system. It is important to note
that both the desalination and the liquefaction utilities are not described as precisely as the
major production unit (Section 4.2) since only the related costs are relevant for this study.

Water preparation unit Desalination methods have been heavily studied and are well-known
technologies which are commercially available. However, the state of development and the
respective costs of each existing process are not similar. Indeed, Wade [19] presents and describes
the energy requirements of three major water preparation methods: multi-stage flash distillation,
multiple effect distillation and reverse osmosis. Since the reverse osmosis (with brine booster) has
the lowest energy consumption of all three processes, it is selected for this study and incorporated

22
4 H2 PRODUCTION PLANT DESIGN

into the overall production plant design. The additional energy requirements are presented in
Table 4.5
Table 4.5: Energy requirements for the water preparation unit

Parameter Value
Conversion energy [MJ/m3 ] [19] 24.2
Plant operation [h/day] 24
Daily requirements [m3 /day] 1072.5
Electric power required [kWe ] 300.4

Liquefaction unit In order to achieve proper hydrogen storage for large scale applications,
two major methods are considered; high compressed gas, generally up to 350-700 bar, or liquid
hydrogen (LH2). Both technologies have their respective benefits and drawbacks which have
been widely defined and described in [12]. Hence, following the selection made during this
previous study, a liquefaction utility is chosen.
Although the minimum theoretical energy needed for hydrogen liquefaction from almost
ambient conditions (300 K, 1 atm) is 3.3-3.9 kWh/kg LH2, the real conversion requirements
are much higher [17]. Indeed, actual liquefaction plants required about 10 to 13 kWh/kg LH2
depending on the system size. Nevertheless, recent technologies such as active magnetic regen-
erative liquefier are predicted to consume much less (Table 4.6) and are thus, highly interesting
for large scale applications. Since this paper evaluates hydrogen production for the long term,
this novel conversion method is selected, despite the fact that it is not commercially mature yet.
Table 4.6: Energy requirements for liquefaction unit

Parameter Value
Conversion energy [kWhe /kg LH2] [17] 7
Plant operation [h/day] 24
Production rate [mt/day] 120
Electric power required [MWe ] 35

4.4 Energy requirements


The different cycle steps being described, the energy requirements, thermal and electrical, need
to be determined in order to assess the MER afterwards. The following section presents the
different methods and assumptions used throughout the evaluation.

Thermal requirements In order to assess the MER, a flowsheet is first elaborated which
models the main process steps occurring in the considered thermodynamical cycle. Due to lack
of information and time, following assumptions have been made:

• The Cu2 OCl2 has been modeled as two single compounds, copper oxide CuO and cop-
per(II) chloride CuCl2 . Obviously, this assumption is disputable. However, since no mul-
tiple objective optimization is performed in this study, all energy balances related to this
compound can be defined by an external code. As the mass balance is not altered, this
hypothesis can be hence validated regarding the precision level of the model required for
this study.

23
4 H2 PRODUCTION PLANT DESIGN

• All reactions are assumed to have a conversion yield of 1


• No pressure drops throughout the entire cycle dp = 0
• All turbines and compressors have an isentropic efficiency of ηc = ηt = 0.85
• Similarly, all pumps have a volumetric efficiency of ηp = 0.85

The corresponding composite curves (energy) and Carnot composite curves (exergy) are
presented in Figure 4.6 and 4.7 respectively. The resulting minimum energy requirements (MER)
are reported in Table 4.7.

Figure 4.6: Composite curves

24
4 H2 PRODUCTION PLANT DESIGN

Figure 4.7: Carnot composite curves

25
4

26
H2 PRODUCTION PLANT DESIGN

Figure 4.8: Four step Cu-Cl cycle


4 H2 PRODUCTION PLANT DESIGN

Table 4.7: Minimum thermal energy requirements of the production cycle

Parameter Value
High graded heat at 565 ◦C
[MWh ] 153.397

Low graded heat at 100 C [MWh ] 86.576
Cooling water at 10 ◦ C [m3 /s] 0.157

Electrical requirements The electrical requirements of the hydrogen production plant and
the complementary utilities are summarized in Table 4.8. It is important to note that the
electrolyzers consumption is not yet reported here since it is defined in the following chapter
(Section 4.4.1).

Table 4.8: Minimum electrical energy requirements of the production cycle

Parameter Value
Net cycle consumption [kWh ] 17’2324
Water preparation unit [kWe ] 300.4
Liquefaction unit [kWe ] 35’000
Subtotal [MWe ] 52.53

4.4.1 Electrolyzer requirements


Most of the electrical power required to drive the Cu-Cl cycle is consumed by the first production
step, the electrolysis. Hence, in order to minimize the energy necessary to run the electrolyzer
modules, the cell voltage Ucell must be as low as possible. Additionally, the current density
should be as high as possible to reduce the cell area and thus, decrease investment costs.
First, the reversible open circuit voltage Uoc across the cell can be expressed as follows
∆G
Uoc = [V ] (4.2)
z·F
where z is the number of electrons transfered, F represents Faraday’s constant and ∆G is the
change in free Gibbs energy of the reaction presented in (4.1), which is defined as

∆G = ∆H − T ∆S [kJ/mol] (4.3)

Finally, by adding the sum of all overpotentials required to overcome the different losses through-
out the system, the cell operating voltage can be determined (4.4).
X
Ucell = Uoc + Uloss,i [V ] (4.4)
i

Regarding first experimental results generated by [27] (Figure 4.9), the operating voltage
decreases linearly with increasing temperature. Moreover, the slope increases with HCl concen-
tration, leading to even lower operating voltages at higher temperatures. The extrapolated data
shows that in order to achieve a high current density of 0.5 A/cm2 and a low cell voltage of 0.7
V, operating conditions over 90◦ C are required.
4
Pumps, compressors and turbines power consumption/generation

27
4 H2 PRODUCTION PLANT DESIGN

Figure 4.9: Effect of temperature and HCl concentration on cell voltage [27]

Consequently, given the operating values of 0.7 V at 0.5 A/cm2 proposed by [22], the min-
imum power required to drive this first production step reaches 93.07 MWe for a total active
cell area of 26’600 m2 . It is important to note that this result does not include the power re-
quired by the different auxiliary units necessary to drive the electrolyzers (pumps, control, etc.).
The following table (Table 4.9) summarizes the total electrical requirements of the hydrogen
production plant.

Table 4.9: Minimum electrical energy requirements of the production plant

Parameter Value
Electrolyzer requirements [MWe ] 93.07
Total cell area [m2 ] 26’600
HCl concentration M [mol/l] 11
Total plant requirements [MWe ] 145.6

4.4.2 Efficiency
In order to evaluate the performance of the hydrogen production plant, the efficiency must be
defined. Hence, based on the first thermodynamic law, the efficiency compares the produced
amount of energy in form of hydrogen (based on the LHV) to the energy input of the system
(Table 4.7 and 4.9). The following equation describes the efficiency used in this study

ṁH LHVH2
= 1 P 2 P (4.5)
η Ėe + Q̇h

where η=0.36 is the heat-to-electricity efficiency of the power block (Section 5.3.2).

28
4 H2 PRODUCTION PLANT DESIGN

4.5 Production plant cost


4.5.1 Capital cost
The investment costs of each plant unit can be defined by using relationships based on reference
data 4.6 or statistical analysis 4.7 [6]. In both cases, the capital expenses are expressed in
function of an equipment attribute A and the cost index ratio It between the actual and reference
year.

A γ It
 
Cp = Cp,ref [$] (4.6)
Aref It,ref

It 2
Cp = Cp,ref 10k1 +k2 logA+k3 (logA) [$] (4.7)
It,ref
Due to the corrosive environment generated by the Cu-Cl process, the carbon steel equipment
is coated with a porcelain lining and hence, capital costs are assumed to be increased by 20%
[15].

Heat exchanger network Since the heat exchanger network (HEN) is not exactly deter-
mined, the related costs are computed by using a different relation than the one presented
above. Indeed, in order to assess the HEN capital expenses, the composite curves are split into
multiple vertical sections for which the corresponding heat capacity cp is constant. After deter-
mining the heat exchanger area for each section, the mean heat exchanger area can be defined
by summing up all computed areas and dividing them by the number of units. By assuming
an equal separation of the area, an overestimation of the HEN investment costs can be finally
generated.

Electrolyzers The capital expenses related to the electrolyzer modules can be approached by
assuming a specific cost factor cel which is multiplied by total active cell area Ael .

Ielectrolyzer = cel · Ael [$] (4.8)

Regarding the investment cost projections proposed by [15], a specific cost factor cel = 1140 $/m2
is considered in this study.

Complementary utilities Similarly to the HEN cost assessment method, the liquefaction
and water preparation plant investment expenses are defined by using a different approach than
presented above. Indeed, given the results presented by [19] and [2], the capital cost of each
utility is defined by only considering their respective output capacity cout . Consequently, the
related plant costs are solely multiplied by a multiplication factor Fmulti in order to meet the
desired utility size required for the hydrogen production. The different parameter values are
presented in Table 4.10.

Cutility = Fmulti · cout,cost [$] (4.9)

Annualized capital cost After defining the capital expenses Cp for each plant unit, the
investment is annualized. Considering a given interest rate i and plant lifetime n the annual
expenses I ? related to the plant equipment cost can be defined. Finally, in order to include

29
4 H2 PRODUCTION PLANT DESIGN

Table 4.10: Complementary utility investment costs

Parameter Value
Liquefaction plant output cout [LH2 kg/d] 30’000
Liquefaction plant output cout,cost [M$] 40
Liquefaction plant factor F [-] 4
Preparation plant output cout [m3 /d] 31’822
Preparation plant output cout,cost [M$] 37.7
Preparation plant factor F [-] 0.034

all additional installation expenditure, the equipment cost is multiplied by a bar module factor
FBM .
? i(1 + i)n
Idirect = FBM Cp [$/yr] (4.10)
(1 + i)n − 1

4.5.2 Fixed cost


The fixed expenses are upfront costs which include various indirect capital investments such
as contingency, construction permits, engineering and design. Considering the indirect cost
assumption established by Kromer et al. [15], fixed expenses can be expressed as function of
direct capital cost (4.11) where cf ixed,total is the sum of all contributing cost factors (Table 4.11).
Similarly to the direct capital expenses, the indirect capital investment is annualized afterwards
(4.12) in order to be included in the total annual plant cost.

Cf ixed = cf ixedcap.,total FBM Cp [$] (4.11)

? i(1 + i)n
Iindirect = Cf ixed [$/yr] (4.12)
(1 + i)n − 1

Table 4.11: Indirect capital cost factors [15]

Parameter Value
Project contingency [-] 0.18
Upfront Permit cost [-] 0.03
Total cf ixedcap.,total [-] 0.21

4.5.3 Operating cost


Since the solely plant input is assumed to be sea water (Section 4.3), the operating costs only
consist of operation and maintenance (O&M) expenses which incorporate labor, property tax,
insurance and replacement costs. Considering the O&M cost assumption defined by Kromer et
al. [15], the following expressions are used in this study

• Labor expenses are determined by multiplying the plant personnel number by the average
labor rate cl,plant . The required staff size can be approached by using [15]

Staff = (6.2 + 31.7P2 + 0.23N)0.5 [−] (4.13)

30
4 H2 PRODUCTION PLANT DESIGN

where P represents the number of steps involving solid particles handling whereas N de-
nominates the remaining process step number (Table 4.12). Since no specific process
diagram was generated for the water preparation unit, the labor cost is assumed to be
0.126$ per cubic meter of water desalinated [19].

• Property tax, insurance and replacement cost are expressed as a function of the direct
capital expenses [15]

Cf ixedop. = cf ixedop.,total FBM Cp [$/yr] (4.14)

where cf ixedop.,total is the sum of all contributing cost factors (Table 4.12).

Considering the high annual output of gaseous oxygen co-generated with the hydrogen pro-
duction, a storage system can be designed in order to sell it as a byproduct and hence strongly
decrease the operating expenses . Indeed, regarding the selling price of gaseous oxygen cO2 in
2002 of 0.109 $/kg O2 5 and the annual plant production of 346’896 mt O2 /yr, a $37.8 millions
income can be included during the operating cost assessment. It is nevertheless important to
mention that the equipment required for pressuring and preparing the oxygen for delivery was
not included in either the energy bill or the direct investment cost evaluation.

OP EX = Clabor + Cf ixedop. − cO2 ṁO2 [$/yr] (4.15)

Table 4.12: Operating cost factors [15]

Parameter Value
Labor rate cl,plant [$/hr] 50
Property tax [-] 0.02
Insurance [-] 0.06
Total cf ixedop.,total [-] 0.08
Cu-Cl Cycle Liquefaction unit
Solids handling P [-] 3 0
Remaining steps N [-] 12 8

4.5.4 Total annual cost


Finally, the total annual cost Ctotal can be determined by simply adding the annualized direct
and indirect capital expenses to the operating costs.
?
Ctotal = Idirect ?
+ Iindirect + OP EX [$/yr] (4.16)

5
http://ed.icheme.org/costchem.html

31
4 H2 PRODUCTION PLANT DESIGN

4.6 Conclusion
This chapter presented the different steps and assumptions used to make a preliminary design of
a hydrogen production plant based on a four-step hybrid copper-chloride themochemical cycle.
Table 4.13 summarizes the different results generated throughout this section.

Table 4.13: MER and investment costs of the overall production plant

Total annual cost MER and efficiency


Parameter Value Parameter Value
Preparation utility [M$] 1.3 Total electricity demand [MWe ] 145.6
Liquefaction utility [M$] 160 High graded heat demand [MWth ] 153.4
Production plant [M$] 215.5 Low graded heat demand [MWth ] 86.6
Total direct capital cost [M$] 376.8 Cooling water [m3 /s] 0.157
Fixed cost [M$] 79.1
Interest rate i [-] 0.08
Plant lifetime n [yr] 25
Operating cost (with O2 selling) [M$/yr] 2.27
Annualized capital cost [M$/yr] 42.7 Cycle efficiency [-] 0.305
Total annual cost [M$/yr] 45
Hydrogen production cost [$/kg LH2] 1.03 Overall plant efficiency [-] 0.26

It is interesting to note that, as presented above, the computed cycle efficiency is much lower
than the value stated in the previous chapter (Section 3.2.2). However, Lewis et al. [16] first
assumed an idealized process for the efficiency computation and hence, by increasing the model
complexity, his efficiency evaluation decreases to 30% which corresponds to the generated result.
Kromer et al. [15] confirmed this value during his assessment.

32
5 POWER SUPPLY PLANT DESIGN

5 Power supply plant design


Since the electrical and thermal energy requirements of the overall production plant have been
defined, the power plant must be designed in order to meet the necessary demands. As the high
grade heat level does not exceed the operating temperature of the thermal storage medium -
which was an important decision parameter - a concentration solar power plant is selected. The
following chapter describes the different steps and assumptions made during the design process.

5.1 Localization
Regarding the low direct normal irradiation and high weather factor related to Switzerland, the
solar power plant and the respective conversion plant are assumed to be located elsewhere in
order to achieve a higher efficiency. Therefore, the coastal region near Ben Gardane in Tunisia
is chosen. Indeed, the benefits of outsourcing the production to this location are:

• The average yearly solar irradiation is higher than any other location in Europe and hence
capital costs are slightly lower

• The coastal land is very cheap in this region compared to South European coasts which
are well developed and extremely expensive

• The coastal location enables a free access to the plant raw material, sea water. After
desalination and deionisation, the water is used for hydrogen generation or solar panel
cleaning and hence, the operating costs are much lower.

Obviously, the major drawback of selecting this location are the delivery costs which are
substantially higher considering the means of transport necessary in this case. It is important
to note that due to lack of time and information, the trade-off between capital and shipping
expenses has not been optimized in this study.

5.2 Solar irradiation


The daily solar irradiation ED at Ben Gardane can be defined by using different parameters
such as the orbital eccentricity
 r 2 D
= 1 + 0.033 cos(2π ) [−] (5.1)
r 365

the declination correction factor


 π  284 + D
δ = 23.45 sin(2π ) [rad] (5.2)
180 365

leading to
 r 2
ED = 10.45 sin(δ) sin(φ)[HD − tan(HD )] [kW h/m2 ] (5.3)
r
where φ is the latitude at which the power plant is located, D the day of the year and HD the
half day of sunlight expressed as an angle

HD = arccos(− tan(δ) tan(φ)) [rad] (5.4)

33
5 POWER SUPPLY PLANT DESIGN

Figure 5.1: Daily extraterrestrial irradiation ED (red), irradiation including reflection losses ηa (yellow),
the weather factor ηw (green) and ground albedo ηg (blue) over the year

Nevertheless, since ED represents the daily extraterrestrial irradiation, correction factors must
be added in order to take in account the different losses engendered by the atmosphere. Hence,
effects due to the atmosphere albedo ηa , the ground albedo ηg > 1 and the weather ηw < 1 are
included
EDcorrected = ED · ηa · ηg · ηw [kW h/m2 ] (5.5)

The previous graph (Figure 5.1) represents the daily irradiation (DI) and the corresponding
losses over an entire year. The different values used during the evaluation are presented in Table
5.1.
Table 5.1: Solar correction factors and parameters

Parameter Value
Latitude of Ben Gardane [◦ ] 33.171
Atmosphere albedo factor [-] 0.69
Weather factor [-] 0.9
Ground albedo factor [-] 1.1

Finally, in order to assess the mean yearly irradiation at Ben Gardane, the corrected daily
irradiations are summed for each day
365
X
EDcorrected = EYcorrected (5.6)
D=1

leading to a value of EY corrected = 2140.68 kWh/m2 · yr.

34
5 POWER SUPPLY PLANT DESIGN

5.3 Plant setup


Since the temperature potential of parabolic trough plants is limited by the HTF, a tower plant
is selected in order to satisfy the high heat grade required by the hydrogen production plant
(Section 4.4). Nevertheless, regarding the lower investment cost related to parabolic trough
plants, a combination of both technologies is studied in this chapter. Indeed, Augsburger [7]
analyzed the thermo-economic performance of a hybrid CSP plant and presented the resulting
benefits through a 100 MWe case study.

5.3.1 Hybrid CSP plant


By combining a parabolic trough and a heliostat field in the same power plant, both efficiency
and investment expenses can be improved [7]. First, medium level heat which is provided by
the trough system is used to satisfy the thermal energy required by the preheater and the first
evaporator. Since the specific cost related to parabolic through fields are lower than heliostat
fields, the plant capital expenses are strongly decreased.

Superheater
Heliostat field 565◦ C
Turbine &
Evaporator II Generator
390◦ C
Evaporator I Condenser

Trough field 290◦ C


Preheater
Pump

Figure 5.2: Hybrid tower/through CSP plant

In the second part, the tower system generates the heat required by the second evaporator, the
superheater and the reheater. Hence, by increasing the inlet temperature of the steam turbine,
the cycle efficiency is highly improved. Moreover, in this case, the heliostat field satisfies the high
graded heat requirements of the hydrogen production facility designed in the previous chapter.
Figure 5.2 illustrates the power plant layout presented above.

5.3.2 Rankine cycle


Regarding the electrical and thermal power requirements of the hydrogen production plant, the
Rankine cycle of the combined CSP plant is designed. Therefore, a flowsheet of the power cycle
is determined by using following assumptions
• No pressure drops throughout the entire cycle dp = 0 [bar]
• All turbines have an isentropic efficiency ηt = 0.85 [-]
• Similarly, all pumps have a volumetric efficiency of ηp = 0.85 [-]
The different performances and characteristics of the cycle considered in this study are presented
in Table 5.2. Since a substantial fraction of the production plant’s MER is low graded heat
(Table 4.7), an extraction system for the low pressure turbine is added to the system. Indeed,
steam is withdrawn at desired temperature and condensed in order to satisfy the low level heat
requirements.

35
5 POWER SUPPLY PLANT DESIGN

Figure 5.3: Rankine cycle with LPE flowsheet

Efficiencies The electrical efficiency of the power cycle can be expressed as

Ėe−
= (5.7)
Q̇+
high

where Q̇+ high is the required heat input delivered by the solar concentration field. Regarding
(5.7), the integration of a low pressure extraction (LPE) system obviously decreases the electrical
efficiency of the cycle. Nevertheless, since the low graded heat can be satisfied by a heat source
at lower temperature than the available solar high level heat, the exergetic efficiency is strongly
affected. Indeed, considering the same electrical and thermal output for both cases (5.8 and 5.9),
the overall exergetic heat input is higher when a cycle without LPE is used since the thermal
output is directly satisfied by the available high graded heat source. Consequently, the exergetic
efficiency of the power cycle is higher when a LPE system is included.

Ėe− + Ėq−low,required
ηwith LP E = (5.8)
Ėq+with LP E
Ėe− + Ėq−low,required
ηwithout LP E = (5.9)
Ėq+without LP E + Ėq+low,required

5.3.3 Cycle integration


The performance and characteristics of the Rankine cycle being presented, the different ther-
mal requirements can be determined in order to size both solar concentration fields. Indeed,
knowing the high level heat required by the production plant - the low level heat and electrical
requirements being already satisfied by the power cycle - and the thermal power necessary to

36
5 POWER SUPPLY PLANT DESIGN

Table 5.2: Power cycle performance and characteristics

Parameter Value
Electrical output Ėe+[MWe ] 145.6
Thermal output at 100◦ C [MWth ] 86.58
Turbine inlet conditions 500 [◦ C] at 175 [bar]
Condensation conditions 25 [◦ C] at 0.25 [bar]
with LPE without LPE
Cycle heat input Q̇+ high [MWth ] 409.79 373.67
Thermal output [MWth ] 86.58 -
Electrical efficiency [-] 0.36 0.39
Exergetic efficiency [-] 0.62 0.55

drive the Rankine cycle, the overall heat amount is defined. Table 5.3 summarizes the thermal
energy requirements of the power cycle.

Table 5.3: Power cycle requirements

Parameter Value
Preheater [MWth ] 180.98
Evaporator [MWth ] 85.07
Superheater [MWth ] 93.69
Reheater [MWth ] 50.05
High graded heat (production plant) [MWth ] 153.397

Figure 5.4: Temperature-enthalpy flow diagram of the molten salt streams (red) and vapor/water
streams (blue)

37
5 POWER SUPPLY PLANT DESIGN

By using energy integration, the composite curves of the molten salt streams and the wa-
ter/steam streams are plotted (Figure 5.4) and hence, the size of each solar concentration field
can be defined. Considering a yearly solar operation time of 2500 hr/yr, the thermal storage
size which enables ”around-the-clock” operation and the respective active mirror area of each
CSP plant are finally determined. The following table presents the different energy demands for
both the trough and tower plants (Table 5.4).

Table 5.4: Hybrid CSP plant size

Trough field size Heliostat field size


Production plant [MWth ] - 153.397
Power cycle [MWth ] 243.519 167.818
Total net output [MWth ] 243.519 321.215
Field efficiency [-] [7, 11] 0.5374 0.492
Tower/receiver efficiency [11] [-] - 0.95
Storage [MWh] 4176.52 5509.05
Mirror area [km2 ] 1.854 2.813
Spacing factor [-] [8] 3.4 5
Field area [km2 ] 6.3 14.06

38
5 POWER SUPPLY PLANT DESIGN

5.4 Cost evaluation


Since both CSP plants have been designed, the respective investment and operation expenses
of each power facility can finally be defined. Similarly to the previous chapter, the capital costs
are broken down into different equipment categories and finally evaluated by using reference
values form literature. The following section presents the different processes and assumptions
used during the cost assessment.

5.4.1 Investment cost


Land In order to assess the land expenses, the same method used by Ausgburger [7] is imple-
mented in this study. Hence, the corresponding investment expenses are split into purchasing
costs Ipurchase and land improvements Iimprov which include levelling, roads and other required
civil works.
Iland = Ipurchase + Iimprov [$] (5.10)
Considering a specific land cost cpurchase , the purchasing expenses can be defined as follows
Ipurchase = cpurchase · Aland [$] (5.11)
where Aland is the total field area (Table 5.4). Finally, the land improvement costs are approxi-
mated by
Aland sland
 
0
Iimprov = Iimprov [$] (5.12)
A0land
0
where A0land and Iimprov are the respective area and investment cost reference values and sland
a scaling factor. All baseline parameter values used during the capital expense assessment are
reported in Table 5.5.
Table 5.5: Land cost reference values [7]

Parameter Value
0
Reference improvement cost Iimprov [M$] 1.1
0 2
Reference field area Aland [km ] 2.8
Reference land specific cost cpurchase [$/m2 ] 0.5
Scaling factor sland [-] 0.3687

Heliostats Similarly to the land cost assessment, the heliostat expenses are computed by
using the correlation defined by Ausburger [7]. Therefore, the heliostat cost can be broken down
into direct purchasing cost Ipurchase , overhead costs Ioverhead which take in account contingency
and corporate oversight expenses and finally indirect costs Iindirect which include engineering
and manufacturing facilities.
Iheliostat = Ipurchase + Ioverhead + Iindirect [$] (5.13)
Regarding the purchasing cost for a single heliostat unit cheliostat , the direct investment expenses
can be expressed as
Ipurchase = cheliostat · Nheliostat [$] (5.14)
where Nheliostat represents the number of heliostat units required for the power plant. Based on
reference case values, the purchasing costs can be evaluated as follows
V
Ahelio shelio,i log2 helio
 
V0
X
0
cpurchase = cdirect,i · 0 · pri helio
· pii [$] (5.15)
i
A helio

39
5 POWER SUPPLY PLANT DESIGN

where A0helio , c0direct,i and Vhelio


0 are the cost reference values for the heliostat area, purchasing
cost and production volume respectively. Finally, pr characterizes the progress ratio of the
volume effect, pi is the price index and shelio defines a scaling factor. All baseline parameter
values used during the purchasing cost evaluation are summarized in Table 5.6 and Appendix
B.1.1.
Table 5.6: Direct heliostat cost reference values [7] and design parameters

Parameter Value
Reference heliostat area A0helio [m2 ] 148
0
Reference volume Vhelio [u] 1625
Heliostat area Ahelio [m2 ] 120
Heliostat Volume Vhelio [u] 23436

The second specific heliostat expenses, the overhead costs Ioverhead , are determined as a
function of the direct investment costs Ipurchase , a progress ration prover and an overhead ratio
or. The different reference and ratio values used during the overhead expense assessment are
reported in Table 5.7.
Vhelio
log2
V0
Ioverhead = Ipurchase · or · (prover ) helio [$] (5.16)

Table 5.7: Overhead cost reference values [7]

Parameter Value
Progress ratio prover [-] 0.96
Overhead ratio or [-] 0.2

Similarly to the direct investment definition (5.15), the last sub-cost related to the heliostats,
the indirect expenses Iindirect , are expressed as follows
shelio,i Vhelio
 log2
X
0 Ahelio V0
Iindirect = Iindirect,i · · pri helio
· pii [$] (5.17)
i
A0helio

0
where Iindirect,i represents the different indirect cost reference values which are presented in
Appendix B.1.1 with their corresponding scaling ratios.

Trough modules The parabolic trough modules expenses are assessed by using the same
correlations determined during the heliostats cost evaluation. The corresponding reference case
values required for the direct investment estimation are stated in Table 5.8 and Appendix B.1.2.
The remaining sub-cost are computed with the same baseline parameter values as for the he-
liostats assessment.

Tower/receiver The capital cost related to the tower/receiver of the heliostat field can be
approached by assuming a scaling factor stower which is multiplied by the field thermal output.

Itower = stower · Pth,heliostat [$] (5.18)

Regarding the investment projections proposed by Turchi et al. [28], a specific cost factor stower =
200 $/kWth is considered in this study.

40
5 POWER SUPPLY PLANT DESIGN

Table 5.8: Direct trough module cost reference values [7] and design parameters

Parameter Value
Reference Trough module area A0trough [m2 ] 3270
0
Reference volume Vtrough [u] 156
Trough module area Atrough [m2 ] 3270
Trough Volume Vtrough [u] 567

Thermal storage Similarly to the tower and receiver cost assessment, the storage facility
expenses can be determined by expressing the capital cost as a function of the overall storage
capacity Estorage and a specific cost factor sstorage .
Istorage = sstorage · Estorage [$] (5.19)
Regarding the investment projections proposed by Turchi et al. [28], a specific cost factor
sstorage = 20 $/kWhth is considered in this study.

Power block The last direct investment expense, the power block, can be evaluated by using
the following relation
Ipower = spower · Pel [$] (5.20)
where spower is the power block specific cost factor. Regarding the investment projections
proposed by Turchi et al. [28], a specific cost factor spower = 1140 $/kWel is considered in this
study.

5.4.2 Operating cost


Since the hybrid CSP does not require any fuel input, the operating costs only consist of oper-
ation and maintenance (O&M) expenses which incorporate labor, property tax, insurance and
replacement costs. Considering the O&M cost assumption defined by Kromer et al. [15], the
following expression are used in this study
• Labor expenses are determined by multiply the plant personnel number by the average
labor rate cl,power . The required staff size can be approached by using
Staff = 1.6 · 10−5 · Af ield [−] (5.21)
where Af ield represents the total field area of the power plant (Table 5.9)
• Property tax, insurance and replacement cost are expressed as a function of the direct
capital expenses
Cf ixedop. = cf ixedop.,total Icapital [$/yr] (5.22)
where cf ixedop.,total is the sum of all contributing cost factors (Table 5.9).
Finally, the total operating cost can be expressed as the sum of the fixed operating expenses
and the labor costs.
OP EX = Clabor + Cf ixedop. [$/yr] (5.23)

5.4.3 Total annual cost


Finally, the total annual cost Ctotal can be determined by simply adding the annualized capital
expenses to the operating costs.
?
Ctotal = Icapital + OP EX [$/yr] (5.24)

41
5 POWER SUPPLY PLANT DESIGN

Table 5.9: Operating cost factors [15]

Parameter Value
Replacement cost [-] 0.005
Insurance and property tax [-] 0.01
Total cf ixedop.,total [-] 0.015
Labor rate cl,power [$/hr] 31
Total field area Af ield [km2 ] 20.36
Staff size [-] 326

5.5 Conclusion
This chapter presented the different steps and assumption used to design a hybrid CSP plant
based on the minimum energy requirements defined in the previous chapter. Table 5.10 sum-
marizes the different results generated throughout this section.

Table 5.10: Investment costs and characteristics of the hybrid CSP plant

Parameter Tower field Through filed


Land expenses Iland [M$] 8.8 4.5
Modules expenses Iheliostats/through [M$] 472.1 275.2
Tower/receiver expenses Itower [M$] 64.2 -
Storage expenses Istorage [M$] 110.2 83.5
Subtotal [M$] 655.3 363.2
Power block Ipower [M$] 178.9
Interest rate i [-] 0.08
Plant lifetime n [yr] 25
Operating cost [M$/yr] 106.5
Annualized capital cost [M$/yr] 112.2
Total annual cost [M$/yr] 218.7
Hydrogen production cost [$/kg LH2] 4.99

42
6 DELIVERY COSTS

6 Delivery costs
The production and power plants being designed, the shipping methods and related costs must
finally be evaluated in order to assess the overall cost of hydrogen production for the given
scenario. Since the fleet hub is located in Geneva - Switzerland, three different transportation
means are analyzed in order to select the less expensive solution afterwards. The following
chapter describes the different steps and assumptions made during the selection and assessment
process.

6.1 Delivery methods


6.1.1 Pipelines
Pipelines are a well-developed delivery mode which is widely used for gaseous product distri-
bution, especially for natural gas. Nevertheless, since the investments costs are extremely high,
this means of transportation is solely economically attractive for very large scale delivery. In-
deed, pipeline capital expenses can be separated into two distinct groups; material costs which
directly depend on the pipeline size (i.e. diameter) and installation costs which highly vary with
various parameters (environment, national policies). Yang et al. [31] assessed these expenses in
the following table (Table 6.1)

Table 6.1: Pipeline operating and capital expenses [31]

Parameter Value
Installation cost - rural [$/km] 300’000
Installation cost - urban [$/km] 600’000
Pipeline capital cost [$/km] 1869 (d6 pipe )2
Fixed operating costs 5% of total capital

6.1.2 Trucking
A second hydrogen transportation type considered in this study is truck transport. The cryogenic
fuel is filled into a large cylindrical vessel which is fixed on an undercarriage of a standard size
semi-truck. The capital and operating costs related to liquid hydrogen trucks are defined by
Yang et al. [31] and presented in Table 6.2.

Table 6.2: LH2 Truck operating and capital expenses [31]

Capital expenses Operating expenses


Parameter Value Parameter Value
Truck capital cost [$] 800’000 Driver hours [hr/d] 8
Truck capacity [kg H2 ] 4000 Driver wage [$/hr] 28.75
Daily boil-off [kg/d] 0.3% Truck lifetime [yr] 20
Truck fuel [LH2 kg/100 km] 8.2
Fixed operating costs [$/yr] 5% of total capital

6
Diameter in inches

43
6 DELIVERY COSTS

6.1.3 Shipping
Since LH2 shipping has not being commercialized yet and hence, no information about its
costs are currently available, analogies with LNG carriers have been made. Similarly to truck
transport, shipping expenses can be split into two parts; capital and operating costs.

Capital costs During the last years, LNG tanker prices have highly fluctuated, especially due
to the financial crisis which lead to a decreased demand for new vessels. Indeed, while 170,000
m3 vessel prices reached a peak value of $250 millions in 2008, mainly caused by high steel prices
and shipyard’s undercapacity, new vessel costs have sharply fallen to $203 millions by the end
of 20107 (Table 6.3).

Operating costs Since LNG tankers are very sophisticated ships, operating costs (Table 6.3)
which include manning, insurances and maintenance are much higher than traditional tanker
expenses, reaching $14 millions per year in 20107 . In addition to these fixed operating expenses,
travel costs which take in account the marine fuel and harbor fees must be added. Nevertheless,
in the case of LH2 carriers, the boil-off losses generated during the trip may directly be used to
power the vessel engines and hence, strongly decrease the fuel costs.

Table 6.3: LH2 carrier operating and capital expenses

Capital expenses Operating expenses


Parameter Value Parameter Value
15,000 cbm Vessel cost [M$] [23] 41 Operating costs [M$/yr] 14
170,000 cbm Vessel cost [M$] 203 Travel costs [M$/yr] 12

6.2 Methods selection


Regarding the distance between the production and consumption sites (Table 6.4), it is obvious
that the pipeline solution is much more expensive than truck and ship transport. Indeed, the
different drawbacks of using such delivery system are:

• The liquefaction plant does not need to be move to Geneva since the pipeline operates
which gaseous hydrogen. Hence, the objective regarding the production from renewable
energy sources is not altered as the plant is not connected to the national grid.

• Undersea pipelines create additional construction costs which are extremely difficult to
assess and therefore, investment costs might be inaccurate.

In light of these results, a combination of truck transport and shipping is selected as delivery
method. Hence, the produced liquid hydrogen is assumed to be loaded on a LH2 tanker and
shipped across the Mediterranean Sea to a European harbor in North Italy, Genoa. Once
arrived, the cryogenic fuel is filled into specific trucks and routed to its final destination at the
International airport of Geneva.

7
http://www.petroleum-economist.com/Article/2801286/LNG-shipping-economics-on-the-rebound.
html

44
6 DELIVERY COSTS

Table 6.4: Transportation data

Parameter Value
Sea distance (Ben Gardane - Genoa) [km] 1300
Mean sea travel speed [km/hr] 30
Land distance (Genoa - Geneva) [km] 400
Mean land travel speed [km/hr] 80

6.3 Delivery cost assessment


6.3.1 Shipping
Given the distance between the production plant and the destination harbor (Table 6.4), one
round trip requires about 5 days, including 33 hours for loading/unloading, maintenance and
maneuvering. Hence, the LH2 tanker capacity should reach 8450 m3 in order to satisfy the
hydrogen demand during the voyage. Since LH2 delivery would require additional costs such as
thicker insulation and different tank materials, a 15’000 m3 vessel price is assumed to be relevant
for this scenario.
Since the delivered amount exceeds the daily fuel requirements of the Clip-Air fleet, two stor-
age facilities, at the production plant and the transfer harbor respectively, must be included in
the investment costs. The capital expenses can be evaluated by using the relationship presented
by [10]:
CAP EXstorage = 1500 · V [$] (6.1)
where V is the net volume in cubic meter of the storage unit.

Table 6.5: Ship capital and operating expenses

Parameter Value
Vessel Capital expenses [M$] 41
Storage capital expenses [M$] 25.36
Operating expense [M$/yr] 26

6.3.2 Truck transport


Regarding the technical and geographical information presented in Table 6.2 and 6.4, thirty
LH2 trucks are required in order to delivery the daily hydrogen amount. The loaded trucks
are leaving the docks around midday and should arrive at destination in the evening. The next
day, the fleet returns to its starting point in the morning in order to reach the filling station at
the harbor before noon. Similarly to the shipping costs, part of the transported hydrogen may
be used to satisfy the overall energy requirements of the truck during the trip (∼82 kg LH2/
round-trip). The total capital and operating expenses are summarized in Table 6.6.

Table 6.6: Truck capital and operating expenses

Parameter Value
Capital expenses [M$] 24
Operating expense [$/yr] 4’348’125

45
6 DELIVERY COSTS

6.4 Conclusion
The total investment and operation costs for the hydrogen delivery are summarized in the
following table (Table 6.7). As presented, the transportation expenses account for almost 0.89
$/ LH2 kg of the overall hydrogen cost.

Table 6.7: Overall capital and operating expenses for hydrogen delivery

Capital expenses Operating expenses


Total capital costs [M$] 90.36
Interest rate [-] 0.08
Project lifetime [yr] 25
Annualized capital costs [M$/yr] 8.46
Total operating costs [M$/yr] 30.35
Total annual costs [M$/yr] 38.81
Hydrogen delivery costs [$/LH2 kg] 0.89
Hydrogen delivery costs [$/kWh LH2] 0.028

46
7 CONCLUSION

7 Conclusion
This study highlighted the different processes and their respective state of development able
to produce hydrogen at large scale, solely by using renewable energy sources. Their analysis
led to the conclusion that a hybrid copper-chloride thermochemical cycle is the best tradeoff
between high efficiency, long system lifetime and technological constraints - especially related
to high temperature thermal storage - and hence, it was selected for a preliminary design.
After determining the MER of the production plant and its auxiliary facilities, a hybrid solar
concentration power plant was designed in order to satisfy the energy demand.

7.1 Hydrogen cost


Regarding the total annual expenses related to the different investment and operating costs,
the overall hydrogen price can finally be assessed: $6.91 per kg of LH2 (or $207.3 per MWh).
Excluding the transportation costs, the hydrogen production price, $6.02 per LH2 kg, is still up
to three times higher than the cost generated when using conventional feedstocks such as coal.
The cost breakdown represented below (Figure 7.1) reveals that total annual expenses related
to the concentration solar power plant (CSPP) are the major contributing cost factor. Since
the production plant (PP) costs solely account for 15% of the global hydrogen price - 17% when
excluding the shipping and trucking expenses (S&T) - the generated result shows that the use
of renewable energy sources for large scale LH2 production is still highly expensive and remains
less attractive than conventional conversion methods.

Figure 7.1: Hydrogen price (left) and investment cost (right) breakdown

Indeed, the European Jet-A selling price reached $988.88 per metric ton in December 2013
which corresponds to an energetic cost of $82.4 per MWh. Thus, the equivalent hydrogen price
presented previously is 152% higher than the standard aviation fuel. Moreover, since no profit
margin has been assumed during the hydrogen cost assessment, the real price gap between both
fuels should be even stronger. Indeed, if an additional rate of return of 20% is considered, the
LH2 price drives up to $8.29 per kg ($7.22 per kg when excluding the transportation expenses)
increasing hence the cost difference up to 202%.
In light of these results, LH2 produced from renewable energy sources remains an unattrac-
tive fuel replacement for the airline industry in the near term. With growing oil prices and
decreasing specific costs of solar power plants, alternative fuel generation could become increas-
ingly interesting in the long term. However, this study gave a first approach of the different
challenges and impediments related to clean hydrogen production for a small aircraft fleet. Re-
garding the total infrastructure and area required to satisfy the daily fuel consumption of solely
five Clip-Air planes, it is obvious that the current air traffic density is highly unsustainable. In
8
http://www.iata.org/publications/economics/fuel-monitor/Pages/price-analysis.aspx

47
7 CONCLUSION

order to achieve environmental friendly air travel in the long term, both mobility and alternative
fuels generation cost must decrease.

7.2 Further work


Since this study was a preliminary hydrogen cost assessment, further key elements should be
analyzed

• Regarding the high level heat required for the selected conversion process, a single CSP
plant has been designed to satisfy the different energy requirements of the production
facility. Nevertheless, depending on the plant location, other renewable energy sources
can be considered during the power plant assessment in order to minimize the total annual
expenses. Indeed, as presented previously, the power generation represents the major cost
factor and hence, additional technologies such as photovoltaic systems, wind and/or sea
turbines could be analyzed with regard to find the most attractive solutions.

• As mentioned in the previous chapter, the cost tradeoff resulting from outsourcing the
LH2 production has not been minimized in this study. A detailed expense analysis should
indeed be performed in order to define the most attractive location leading to low delivery
cost and has high annual solar irradiation.

• A strong limiting factor during the process selection was the high level heat required by
most thermochemical cycles which exceeds the working temperature of current thermal
storage units. Hence, with the introduction of novel storage methods allowing higher heat
grades, a wider range of conversion processes could be considered for continuous LH2
production. Further studies related to high temperature storage medium research and
development could then be analyzed.

48
8 REFERENCES

8 References
[1] Birelaire M Leonardi C Atasoy B, Salani M. Impact analysis of a flexible air transportation
system. EJTIR, 13:123–146, 2013.

[2] Ferguson L Pouresfandiary J Cousins A Hampton T Rosenthal P Chumbley H Ralph H


Random J Barclay J, Oseen-Senda K. Active magnetic regenerative liquefier. DOE hydrogen
programm, annual progress report, February 2010.

[3] Olson N. K Bartels J. R, Pate M. B. An economic survey of hydrogen production from con-
ventional and alternative energy sources. International journal of hydrogen energy, 35:8371–
8384, 2010.

[4] Zahid M Brisse A, Schefold J. High temperature water electrolysis in solid oxide cells.
International journal of hydrogen energy, 33:5375–5382, 2008.

[5] Mergel J Stolten D Carmo M, Fritz D. L. A comprehensive review on pem water electrolysis.
International journal of hydrogen energy, 38:4901–4934, 2013.

[6] Marechal F. Advanced energetics. EPFL semester course, 2012.

[7] Augsburger G. Thermo-economic optimisation of large solar tower power plants. PhD
thesis, EPFL, 2013.

[8] Sargent & Lundy LLC Consulting Group. Assessment of parabolic trough and power tower
solar technology cost and performance forecasts. National Renewable energy laboratory
NREL, October 2003.

[9] Dr. Berster P Bischoff G Prof. Dr. Ehmer H Dr. Gelhausen M ... Dr. Scheelhaase J
Grunewald E, Ayazkhani A. Annual analyses of the European air transport market.
Deutsches Zentrum für Luft- und Raumfahrt (DLR), December 2008.

[10] Werner H. The flexible solution - i.m. skaugen’s fleet of small scale lng carriers. IMS -
Innovative Maritime Solutions, June 2007.

[11] Curtin B Wonhas A Boyd R Grima C Tadros A Hall R Naicker K Hinkley T. J, Hayward
A. J. An analysis of the costs and opportunities for concentrating solar power in australia.
Renewable energy, 57:653–661, 2013.

[12] Amouyal Y Stadler P Ingram J, Nahon J. Cryogenic fuel tanks for the clip-air plane.
Bachelor project, EPFL, June 2012.

[13] Revankar T. S Kane C. Suflur-iodine thermochemical cycle: Hi decomoposition flow sheet


analysis. International journal of hydrogen energy, 33:5996–6005, 2008.

[14] Siegel N Kolb G. J, Diver R. B. Central-station solar hydrogen power plant. American
society of mechanical engineering, 129:179–183, 2007.

[15] Takata R Chin P Kromer M, Roth K. Support for cost analyses on solar-driven high
temperature thermochemical water-splitting cycles. Final report to the DOE, prepared by
TIAX LLC, February 2011.

[16] Masin G. J Lewis M. A. The evaluation of alternative thermochemical cycles - part ii: The
down-selection. International journal of hydrogen energy, 34:4125–4135, 2009.

49
8 REFERENCES

[17] Gardiner M. Energy requirements for hydrogen gas compression and liquefaction as related
to vehicle storage needs. Technical report, Department of energy (DOE), October 2009.

[18] Rutkowski M. Current (2005) hydrogen from coal with CO2 capture and sequestration.
National Renewable Energy Laboratory, Jun 2005.

[19] Wade N. M. Distillation plant development and cost update. Desalination, 136:3–12, July
2000.

[20] David F. T Paul M Michele A. L, Magali S. F. Evaluation of alternative thermochemical


cycles - part iiifurther development of the cu-cl cycle. International journal of hydrogen
energy, 34:4136– 4145, 2009.

[21] Wang Z. L Daggupati V.N Gravelsins R. Naterer G. F, Gabriel K. Thermochemical hy-


drogen production with a copper-chlorine cycle. i: oxygen release from copper oxychloride
decomposition. International journal of hydrogen energy, 33:5439–5450, 2008.

[22] Stolberg L Lewis M Wangd Z Dincer I Rosen M.A Gabriel K Secnik E Easton E.B Pioro I
Lvov S Jiang J Mostaghimi J Ikeda B.M Rizvi G Luf L Odukoya A Spekkens P Fowler M
Avsec J Naterer G.F, Suppiah S. Progress of international hydrogen production network
for the thermochemical cuecl cycle. International journal of hydrogen energy, 38:740 –759,
2013.

[23] United Nations. Review of maritime transport 2011. United Nations conference on trade
and development, pages 64–84, 2011.

[24] BINE projecktinfo. Power plant with coal gasification. RWE project description, 2006.

[25] Vie P. J. S Ulleberg O Riis T, Hagen E. F. Hydrogen production and storage. International
energy agency IEA, Paris, 2006.

[26] Penner S. Steps toward the hydrogen economy. Energy, 31:33–43, 2006.

[27] Suppiah S Stolberg L. N, Shkarupin A. Recent advances towards the development of


the electrolyzer for hydrogen production by the copper-chlorine thermochemical cycle.
WHEC2012, Toronto, Ontario, June 2012.

[28] Clifford K. H Kolb J. G Turchi C, Mehos M. Current and future costs for parabolic trough
and power tower dystems in the us market. National Renewable energy laboratory NREL,
October 2010.

[29] Beuret V. Evolution des marchés en énergies fossiles 2/2013. Office fédérale de l’énergie
(OFEN), Juin 2013.

[30] Gabriel K.S Gravelsins R Daggupati V.N Wang Z.L, Naterer G.F. Comparison of different
copperchlorine thermochemical cycles for hydrogen production. International journal of
hydrogen energy, 34:3267 – 3276, 2009.

[31] Ogden J Yang C. Determining the lowest-cost hydrogen delivery mode. International
journal of hydrogen energy, (32):268–286, 2007.

50
LIST OF FIGURES

List of Figures
1.1 The Clip-Air aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1 A320-100/200 fuel consumption (FC) graph fitting for the respective flight dis-
tance (FD) assuming a landing weight of 57 [mt] . . . . . . . . . . . . . . . . . . 5
2.2 Short-haul and medium-haul flight ranges from LSGG . . . . . . . . . . . . . . . 6
3.1 Gasification processes for coal [24] . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Operating principle of an alkaline electrolyzer . . . . . . . . . . . . . . . . . . . . 10
3.3 Operating principle of a PEM electrolyzer . . . . . . . . . . . . . . . . . . . . . . 12
3.4 Energy of reaction (∆ H, red) and Gibbs energy of reaction (∆ G, gray) in function
of the steam inlet temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.5 Operating principle of a SOEC . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.6 Operating principle of a sulfur-iodine cycle . . . . . . . . . . . . . . . . . . . . . 14
4.1 Operating principle of a four-step copper chloride cycle . . . . . . . . . . . . . . . 19
4.2 Operating principle of the CuCl/HCl electrolysis . . . . . . . . . . . . . . . . . . 20
4.3 Hydrolysis spray reactor [20] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4 Oxychloride decomposition reactor [20] . . . . . . . . . . . . . . . . . . . . . . . . 22
4.5 Plant functional sheet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.6 Composite curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.7 Carnot composite curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.8 Four step Cu-Cl cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.9 Effect of temperature and HCl concentration on cell voltage [27] . . . . . . . . . 28
5.1 Daily extraterrestrial irradiation ED (red), irradiation including reflection losses
ηa (yellow), the weather factor ηw (green) and ground albedo ηg (blue) over the
year . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
5.2 Hybrid tower/through CSP plant . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 Rankine cycle with LPE flowsheet . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.4 Temperature-enthalpy flow diagram of the molten salt streams (red) and va-
por/water streams (blue) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.1 Hydrogen price (left) and investment cost (right) breakdown . . . . . . . . . . . . 47
A1 Fuel consumption sheet of a A320 100/200 powered by two CFM56-5C4 . . . . . i

51
LIST OF TABLES

List of Tables
2.1 Clip-Air and A320 specific weight [1] . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Fitted function properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Fuel consumption of Clip-Air and A320-100/200 . . . . . . . . . . . . . . . . . . 5
2.4 Scenario characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.5 Hydrogen assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1 State of the art for specifications of Alkaline and PEM electrolysis [5] . . . . . . 11
3.2 Major characteristics for each hydrogen production process . . . . . . . . . . . . 16
4.1 Chemical reactions in a four-step Cu-Cl cycle . . . . . . . . . . . . . . . . . . . . 18
4.2 Heat grade and quantity required for precipitate drying [30] . . . . . . . . . . . . 20
4.3 Heat grade and quantity required for the hydrolysis reaction [21, 20] . . . . . . . 21
4.4 Heat grade and quantity required for the decomposition reaction [21, 20] . . . . . 21
4.5 Energy requirements for the water preparation unit . . . . . . . . . . . . . . . . . 23
4.6 Energy requirements for liquefaction unit . . . . . . . . . . . . . . . . . . . . . . 23
4.7 Minimum thermal energy requirements of the production cycle . . . . . . . . . . 27
4.8 Minimum electrical energy requirements of the production cycle . . . . . . . . . . 27
4.9 Minimum electrical energy requirements of the production plant . . . . . . . . . 28
4.10 Complementary utility investment costs . . . . . . . . . . . . . . . . . . . . . . . 30
4.11 Indirect capital cost factors [15] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.12 Operating cost factors [15] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.13 MER and investment costs of the overall production plant . . . . . . . . . . . . . 32
5.1 Solar correction factors and parameters . . . . . . . . . . . . . . . . . . . . . . . 34
5.2 Power cycle performance and characteristics . . . . . . . . . . . . . . . . . . . . 37
5.3 Power cycle requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.4 Hybrid CSP plant size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.5 Land cost reference values [7] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.6 Direct heliostat cost reference values [7] and design parameters . . . . . . . . . . 40
5.7 Overhead cost reference values [7] . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.8 Direct trough module cost reference values [7] and design parameters . . . . . . . 41
5.9 Operating cost factors [15] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.10 Investment costs and characteristics of the hybrid CSP plant . . . . . . . . . . . 42
6.1 Pipeline operating and capital expenses [31] . . . . . . . . . . . . . . . . . . . . . 43
6.2 LH2 Truck operating and capital expenses [31] . . . . . . . . . . . . . . . . . . . 43
6.3 LH2 carrier operating and capital expenses . . . . . . . . . . . . . . . . . . . . . 44
6.4 Transportation data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.5 Ship capital and operating expenses . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.6 Truck capital and operating expenses . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.7 Overall capital and operating expenses for hydrogen delivery . . . . . . . . . . . 46
B1 Direct heliostat cost reference values [7] . . . . . . . . . . . . . . . . . . . . . . . ii
B2 Indirect heliostat cost reference values [7] . . . . . . . . . . . . . . . . . . . . . . ii
B3 Direct trough module cost reference values [7] . . . . . . . . . . . . . . . . . . . . ii

52
A FUEL ASSESSMENT

Appendices
A Fuel assessment
A.1 A320 fuel consumption sheet

Figure A1: Fuel consumption sheet of a A320 100/200 powered by two CFM56-5C4

i
B POWER PLANT DESIGN

B Power plant design


B.1 Capital cost reference values
B.1.1 Heliostat reference values
The reference case values used for the direct and indirect capital cost assessment of the heliostats
are exposed in the following tables (Tables B1 and B2).

Table B1: Direct heliostat cost reference values [7]

Direct cost parameters c0direct [$/u] s [-] pr [-] pi [-]


Foundation 200 0.2274 0.9806 1.0816
Pedestal & structure 3777 1.47 0.99 1.807
Drives 6000 0.6 0.94 1.3702
Mirrors 4996 1.042 0.97 1.0861
Control & communications 875 0.2311 0.96 1.2841
Wiring 877 0.4479 1 1.0302
Shop fabrication 480 0.4264 0.98 1
Installation & checkout 450 0.2610 1 1

Table B2: Indirect heliostat cost reference values [7]

indirect cost parameters c0direct [$] s [-] pr [-] pi [-]


Engineering 250’000 0.9551 0.96 1.2623
Manufacturing facilities & tooling 800’000 0.9551 0.86 1.146
Equipment lease & tooling 200’000 0.9551 0.86 1.146

B.1.2 Trough modules reference values


The reference case values used for the direct capital cost assessment of parabolic trough modules
are exposed in the following table (Table B3).

Table B3: Direct trough module cost reference values [7]

Direct cost parameters c0direct [$/m2 ] pr [-] pi [-]


Foundation 21 0.8993 1.0958
Concentration structure 50 0.9408 1.6869
Drives 14 0.8059 1.328
Mirrors 40 0.8839 1.0505
Receivers 43 0.8753 1.0505
Electronics & control 16 0.7169 1.1894
Interconnection pipping 11 0.7319 1.2677
Header pipping 8 0.8313 1.2677
Installation 17 0.8933 1
Other 17 0.8665 1

ii

You might also like