Sei sulla pagina 1di 17

ARTICLE IN PRESS

Renewable Energy 32 (2007) 680696


www.elsevier.com/locate/renene

Techno-economic analysis of the integration


of hydrogen energy technologies in renewable
energy-based stand-alone power systems
E.I. Zoulias, N. Lymberopoulos
Centre for Renewable Energy Sources (CRES), 19th km Marathon Avenue, GR 19009 Pikermi, Greece
Received 7 March 2005; accepted 9 February 2006
Available online 29 March 2006

Abstract
A large number of stand-alone power systems that are based on fossil fuel or renewable energy
(RE) based, are installed all over Europe. Such systems, often comprising photovoltaics (PV) and/or
diesel generators provide power to communities or technical installations, which do not have access
to the local or national electricity grid. The replacement of conventional technologies such as diesel
generators and/or batteries with hydrogen technologies, including fuel cells in an existing PV-diesel
stand-alone power system providing electricity to a remote community was simulated and optimised,
using the hybrid optimisation model for electric renewables (HOMER) simulation tool. A technoeconomic analysis of the existing hybrid stand-alone power system and the optimised hydrogenbased system was also conducted. The results of the analyses showed that the replacement of fossil
fuel based gensets with hydrogen technologies is technically feasible, but still not economically viable,
unless signicant reductions in the cost of hydrogen technologies are made in the future.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Hydrogen; Electrolyser; Fuel cells; Stand-alone power systems; Simulation; Photovoltaics

1. Introduction
Numerous stand-alone power systems (SAPS) have been installed around Europe. These
systems provide power to technical installations and communities in areas that do not have
access to the regional or national power grid. An increasing number of SAPS include
Corresponding author. Tel.: +30 2106603327; fax: +30 2106603301.

E-mail addresses: mzoulias@cres.gr (E.I. Zoulias), nlymber@cres.gr (N. Lymberopoulos).


0960-1481/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.renene.2006.02.005

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

681

renewable energy (RE) technologies (solar or wind power), often in combination with
diesel generators and/or batteries for back-up power, but the majority of larger SAPS are
still based on fossil fuel power production. Replacing diesel generators and batteries by
fuel cells (FC) running on locally produced hydrogen would diminish fossil fuel
dependence, reduce environmental impact and potentially reduce operation and
maintenance costs. The SAPS market is believed to be a market niche where fuel cell
technology can be competitive in the short term, due to the high operation and
maintenance costs [1,2].
The storage of renewable energy (mostly solar energy) in the form of electrolytically
produced hydrogen in stationary stand-alone power systems and the re-electrication of
hydrogen in fuel cells has been investigated in many demonstration systems [37] and is
relevant to non grid-connected communities and remote areas. Various researchers have
investigated such systems, as described in the following paragraphs.
Besides the development of energy technology components, the study of systems and the
optimisation of their operation have also received some limited attention [8,9] and is
certainly an area requiring further attention.
Agbossou et al. [10] described a renewable energy SAPS system, comprising a small wind
turbine and photovoltaic (PV) panels, which supply excess energy to an electrolyser
producing hydrogen that is stored in order to be used in fuel cells. Development and
experimental testing of the best methods for design optimisation and control strategies, as
well as the integration of the fuel cell was examined during this study. Such a system was
proposed for remote communication stations.
Kelouwani et al. [11] developed a dynamic model in order to simulate power systems,
which are fed by renewable energy sources (mostly wind and solar energy), with batteries
and gaseous hydrogen as energy reservoirs, and electrolysers and fuel cells as converters of
energy between electricity and hydrogen. In this model, power interfaces (boost and buck
converters) are modelled for both their transient and steady state behaviours.
Wallmark and Alvfors [12] have studied the technical design and performed an
economic evaluation of a stand-alone proton exchange membrane (PEM) FC system,
which is installed to avoid replacing or upgrading an ageing, distant power grid connection
that only supplies a few buildings. The results of the study showed that the fuel cell system
installation is not economically viable for the present conditions; the importance of sizing
the system components is also stressed.
Stand-alone hydrogen energy systems with different renewable energy sources, namely
solar, hydraulic and wind energy were analysed by Santarelli et al. [13], in view of their
design and operation. The analysis showed that the preferred solution is case specic,
depending on meteorological resources and load prole. For the selected site (a remote
community in North Italy) it was concluded that a micro-hydro plant combined with an
electrolyser, a hydrogen storage tank and a PEM fuel cell was the optimum choice.
Khan and Iqbal [14] prepared a pre-feasibility study for stand-alone hybrid energy
systems with hydrogen as an energy carrier for applications in Newfoundland, Canada.
Different renewable and non-renewable energy solutions and various energy storage
methods were considered using hybrid optimisation model for electric renewables
(HOMER) as a simulation and optimisation tool. According to the results, at
present costs a wind-diesel-battery system was the most favourable solution, but a fuel
cell cost reduction to 15% of its current value would make a wind-fuel cell system a
superior choice.

ARTICLE IN PRESS
682

E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

A stand-alone power system comprising a diesel generator, photovoltaics and batteries


was installed in a remote area of Kythnos island, Greece to serve a small community of
holiday houses, with a daily peak power demand of approximately 8.3 kW, which are
inhabited mostly during summer, as described by Strauss et al. [15].
The current paper examines the techno-economic aspects of replacing diesel generators
and/or batteries of the Kythnos system by fuel cells, an electrolyser and a conventional
hydrogen storage tank and also presents sizing optimisation and simulation results of both
the conventional and hydrogen-based power systems. HOMER version 2.13 developed by
National Renewable Energy Laboratory (NREL) was used as the simulation and
optimisation tool. HOMER contains a number of energy component models, such as
photovoltaics, wind turbines, hydro, batteries, diesel and other fuel generators, electrolysis
units, fuel cells, etc. [16]. To perform analysis with HOMER, information on resources,
electrical loads, economic constraints, current and future equipment cost, and control
strategies is required. Different combinations of PV, batteries, electrolysers and fuel cells
sizes were selected in order to identify the optimal combination of the hydrogen based
system. Since costs of fuel cells are expected to reduce in the future [17], emphasis was
given in analysing the size, cost and operational characteristics of the PV-hydrogen system
in comparison to the already existing PV-diesel one.
2. Methods and system description
Kythnos is an island, which is part of the Cyclades in the Aegean Sea (371250 N, 241250 E).
In the settlement of Gaidouromantra, an area non-interconnected to the local electricity
grid, a small community of ten houses has been electried through two European Joule
projects PV-MODE and MORE. Distributed PV modules with a total nominal
capacity of 8.8 kW (covering an area of around 73 m2), a battery bank and a small diesel
engine generator set of 8 kW are all connected in an AC mini-grid [15]. The settlements are
only populated during summertime thus the batteries are fully charged at the beginning of
the load period and PV-power is being dumped for long periods of the year. It should be
noted that the users require no production of heat, since the settlements are not inhabited
during winter. The experience from the system operation showed that the RE penetration
is very low during the peak months and the diesel generator is operated frequently. These
problems could be overcome by introducing hydrogen technologies to the system, which
would replace conventional technologies, such as diesel generators and/or batteries. Thus
the excess PV power can be used to drive an electrolyser and produce hydrogen at low
demand periods. Electrolytically produced hydrogen can then be stored in conventional
tanks, under pressure, and re-electried, at peak power demand periods, through a PEM
fuel cell.
The objectives of this study were to: (1) simulate the existing PV-diesel system to record
its operational characteristics and calculate the cost of energy produced by the
conventional system, (2) optimise the component sizes of the envisaged PV-hydrogen
system with regards to its performance and cost of energy produced, (3) perform a technoeconomic analysis of the PV-hydrogen system in comparison to the already existing PVdiesel power system, taking also into account future cost scenarios and EU cost targets for
fuel cells and other hydrogen technologies. Simulation, optimisation and techno-economic
analysis of the hybrid power systems were performed with the help of HOMER tool
developed by NREL.

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

683

Actual load prole and meteorological data from the operation of the PV-diesel system
in Kythnos were used in this study. In Fig. 1 the load prole of the system for a typical day
of July is shown. The electrical load has a seasonal variation with June, July and August as
peak months and no load in the winter months. The typical daily load prole from the
same month shows a relatively high nighttime base load. The annual average of the
electrical load is 50 kWh day1 and the annual daily peak is 8.28 kW. Solar radiation data
from Kythnos, namely clearness index and daily radiation are given in Fig. 2. Hourly solar
radiation measurements for a period of 1 year (8760 points) were imported into HOMER
in order to calculate monthly average values of clearness index and daily radiation. As it
can be seen, the solar radiation is high, especially between May and August. For this
location, the average annual clearness index is 0.601 and the average daily radiation is
4.87 kWh m2. The same resources and load data were also used in optimisation process
and techno-economic analysis of the PV-hydrogen system.

Daily Load Profile


9

Electrical Load (kW)

8
7
6
5
4
3
2
0

10

12

14

16

18

20

22

24

Time (hr)
Fig. 1. Electrical load of Kythnos power system for a typical peak month day.

1.0

Daily Radiation
Clearness Index

0.8

0.6
4
0.4
2

0.2

Clearness Index

Daily Radiation (kWh/m2/d)

Solar Resource
8

0.0

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Fig. 2. Average daily solar radiation and clearness index of Kythnos for a 1-year period.

ARTICLE IN PRESS
684

E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

3. Results and discussion


As described earlier, the PV-diesel power system, was simulated by HOMER, in order to
evaluate its operational and economic characteristics. Dimensioning of RES and hydrogen
technology components of the PV-hydrogen stand-alone power system was then optimised.
Finally, a techno-economic analysis of the optimised hydrogen based system and a
comparison to the existing conventional one were performed. The results of simulation,
optimisation and analyses for both systems are described in the following sections.
3.1. PV-diesel power system
The PV-diesel power system is schematically presented in Fig. 3. The main energy
components of the system are PV panels, a diesel generator, batteries and a power
converter. Number of units to be used, capital, replacement and O&M costs, operating
hours, etc. have to be dened in HOMER in order to simulate the system. Information on
autonomous power systems components costs were mainly drawn by RETScreen database
in Canada [18] and through personal contacts with equipment manufacturers [1].
Component descriptions are given in the following paragraphs.
3.1.1. Photovoltaics
The installation costs for a stand-alone photovoltaic array range from 6000 to
9000 h kW1 [18]. In the present case, a cost of 6750 h kW1 was used, resulting in a total
capital cost of 59,400 h for an 8.8 kW PV array. Photovoltaics O&M cost is considered
practically zero and their lifetime is 30 years.
3.1.2. Diesel generator
Capital costs for a commercially available diesel generator range from 200 to
500 h kW1 [18], the cost dropping with increasing capacity. In the Kythnos case a

Fig. 3. HOMER implementation of the PV-diesel power system [16].

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

685

relatively small generator (8 kW) is installed therefore a capital cost of 500 h kW1 was
used in the calculations. In stand-alone power systems a diesel fuel tank should also be
introduced to the system. The cost of a diesel tank with a capacity of 3 m3 was
approximately 3300 h. Therefore, the total capital cost of the diesel system was ca. 7300 h.
3.1.3. Batteries
Batteries are considered as a major cost factor in small-scale stand-alone power systems.
A battery bank of commercially available units with a capacity of 1000 Ah each was
installed. The total capacity of batteries installed was 420 kWh and the estimated lifetime
was 5 years. The total capital cost of the battery bank was 42,000 h [18].
3.1.4. Power converter
An ACDC power converter unit has been installed in the PV-diesel system of Kythnos.
Power conditioning capital cost is around 600800 h kW1 [14,18]. A cost of 670 h kW1 was
chosen for Kythnos case. A 10 kW converter was integrated to the system and its total capital
cost was 6700 h. A lifetime of 16 years was assumed and a converter efciency of 95%.
3.1.5. PV-diesel system simulation
The already existing system was then simulated in order to evaluate its operational
characteristics, namely annual electrical energy production, annual electrical loads served,
excess electricity, renewable energy fraction, capacity shortage, unmet load etc. Some
environmental impact parameters of the system (carbon emissions, annual diesel
consumption) were also calculated. A load-following control strategy was followed in
the simulation. Under this strategy, whenever a power generator is needed it produces only
enough power to meet the demand. Load following strategy tends to be optimal in systems
with a surplus of RE compared to the load, which is the case for the PV-diesel system. In
addition, the simultaneous operation of multiple generators in the system was allowed, as
well as the integration of generators with capacities less than peak load.
Moreover, a cost analysis of the power system and a sensitivity analysis of the effect of
diesel price on the cost of energy produced by the PV-diesel system were also conducted.
Three discrete values (0.4, 0.6 and 0.8 h L1) were used in the sensitivity analysis. At
present the average diesel price in Greece is 0.42 h L1 and for a very remote location,
especially on an island, this could increase up to 0.8 h L1 [19]. An annual interest rate of
6% and a project lifetime of 16 years were used in the economic calculations.
The results of the simulation showed that the PV-diesel system had a total annual
electrical energy production of 24,147 kWh, the renewable energy fraction of which was
0.65 (i.e. 15,591 kWh were produced by the PV array). All results related to the electricity
production and load served by the system are summarised in Table 1. As it can be derived
from Table 1, a large amount of electricity, approximately 20% (mainly produced by PV)
is dumped, therefore RE penetration is low during the peak months and the diesel
generator is operated frequently. The latter is attributed to the high nighttime load, which
enables the operation of diesel because PV energy stored in batteries is not adequate to
serve the load overnight. To increase RES penetration excess energy can be stored in the
form of compressed hydrogen and drive a PEM fuel cell as described under paragraph 3.2.
The monthly average production of electricity from Photovoltaics and the diesel genset
for the existing system is demonstrated in Fig. 4. The diesel generator serves the electrical
load mainly during the summer months and according to the simulation results, it operates

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

686

Table 1
Electrical production and demand for the PV-diesel system
Annual electricity production
PV-array
Diesel generator
Renewable fraction
Total production

15,591 kWh (65%)


8556 kWh (35%)
0.646
24,147 kWh

Annual electrical load served


AC primary load served

18,247 kWh

Other
Excess electricity
Capacity shortage
CO2 emissions

4717 kWh
76.4 kWh
8764 kg year1

Monthly Average Electrical Production


7

PV
Diesel generator

Power (kW)

6
5
4
3
2
1
0

Jan

Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Fig. 4. Monthly average production of electricity for the PV-diesel system.

1858 h year1 and has a fuel consumption of around 3300 L year1, which is in good
agreement with experience from the real system.
Fig. 5 shows the operational characteristics of the PV-diesel system (namely AC primary
load served, power produced by PV panels, and power produced by diesel generator) for a
typical day of month July. According to this gure, during daytime (from 9:00 to 17:00)
when solar radiation is highthe PV panels serve the best part of AC load and the diesel
generator operates at its minimum load ratio (30%) producing only 2.4 kW. Moreover, a
signicant amount of excess electricity is produced during the same period of the day. On
the contrary, during nighttime the diesel generator mainly serves the electrical demand.
The results of the economic analysis performed with HOMER were of great interest. As
can be seen in Table 2, where total annualised costs for each component of the PV-diesel
system are given, the battery bank is the main cost factor of such a hybrid power system,
followed by the diesel generator and PV panels. This is attributed to the fact that the
lifetime of batteries is only 5 years, and the system lifetime is 16 years. Therefore the
battery bank needs to be replaced several times during the project, resulting in a highannualised cost. O&M and diesel fuel costs are also signicant (3616 h year1) for this
system.
As shown in Fig. 6, where the results of sensitivity analysis on diesel fuel price
are presented, the cost of diesel affects the total cost of electrical energy produced by a

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

687

9
8
AC load

Power (kW)

7
6

PV power

5
4

Diesel
generator

Excess
electricity

2
1
0
1

11

13

15

17

19

21

23

Time (hr)
Fig. 5. PV-diesel system characteristics for a typical day of month July.

Table 2
Distribution of annualised costs for the main components of the PV-diesel system
Component

PV array
Diesel
generator
Battery
Converter
Total

Annualised
replacement
(h year1)

Annual O&M
(h year1)

Annual fuel
(h year1)

5878
720

1080
903

0
528

0
2662

4798
4814

4156
663
11,417

5852
0
5675

420
6
954

0
0
2662

10,428
669
20,708

Annualised
capital
(h year1)

Total
annualised
(h year1)

stand-alone hybrid system, but the impact is not very signicant due to the fact that energy
components installed in this system have high capital costs, therefore O&M and fuel cost is
only a small percentage of the total cost of energy produced. The cost of energy for the
PV-diesel system is estimated at around 1.135 h kWh1, for a diesel price of 0.8 h L1.
The operational characteristics derived from simulations performed with HOMER as
well as the techno-economic analysis of the PV-diesel system were used as a basis for
comparison to the respective results of the hydrogen-based power system in order to
evaluate its feasibility and economic viability.
3.2. PV-hydrogen power system
Hydrogen production from photovoltaics received much attention in the late 1980s and
early to mid 1990s. Another general observation is that stand-alone operation or partly
stand-alone operation has become more of an issue over the last ten years and a signicant
number of PV-hydrogen demonstration projects have been realized [37].
Diesel generators and/or batteries can be replaced by integrating to the system an
electrolysis unit driven by excess electricity produced from RES. In our case, excess

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

688

Cost of Energy ( /kWh)

1.2
1.15
1.1
1.05
1
0.95
0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Diesel price ( / L)
Fig. 6. Cost of energy produced by the PV-diesel system as a function of diesel price.

electricity (almost 20%, as mentioned in 3.1.) produced by the PV array, mostly during
wintertime (when there is no demand of electricity), can be used to electrolytically produce
hydrogen gas, which can be stored in conventional tanks under pressure, at 30 bar. It must
be mentioned that advanced electrolysers are capable of delivering hydrogen at 30 bar
without using a hydrogen compressor, which improves system overall efciency. Stored
hydrogen can then be re-electried in PEM fuel cells, providing electricity to the system at
periods of high electrical demand.
The experience from the operation of PV-hydrogen stand-alone power systems showed
that the alkaline electrolyser technology is mature enough for solar and wind applications.
However, the need for more long-term testing of deterioration of hydrogen yield when
applying intermittent energy was emphasised. The most frequent reason for plant
shutdowns in similar RES and hydrogen installations was failure of electrolyser auxiliary
components; therefore the need for more R&D work on the eld of electrolysers power
electronics was stressed.
Another general concern was the complexity and parasitic energy losses of the hydrogen
energy systems due to the need for gas control. Hence, controllers, compressors, converters
and gas cleaning equipment increased the complexity compared to conventional standalone power systems and also reduced the overall energy efciency [1].
The dimensioning of the system components such as PV arrays, diesel generators,
batteries etc. in a hybrid power system is a more general issue of design [2]. Appropriate
software tools and optimisation processes are used in order to determine the most reliable
solution to the problem of system dimensioning.
3.2.1. PV-hydrogen system optimisation
One of the most difcult problems in the design of stand-alone power systems is the
optimisation of different energy components sizes, with regards to the cost of energy and
overall system performance. The HOMER simulation tool was used to optimise the sizes
of different hardware components in the PV-hydrogen system, taking into account the
technical characteristics of system operation and minimising total net present cost of the
system. The economic assumptions were an important input to the optimisation. These
were identied from a search through available literature and direct contacts with

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

689

Table 3
Cost model assumptions used in the techno-economic optimisation
Component
description

PV
Electrolyser
H2 storage unit
FC
Battery
a

Capacity in

kW
Nm3 h1
Nm3
kW
kWh

Cost t parameters linear model Y A+BXa

O&M

% of capital cost

0
0
0
0
0

6750
8150
38
3000
100

0.0
2.0
0.5
2.5
1.0

Where Y: component capital cost and X: component capacity.

hydrogen and renewable energy manufacturers performed during a previous study [1,2].
The cost models used in the optimisation and the sensitivity analysis of the hydrogen
stand-alone power system are given in Table 3.
Different sizes for the PV array, batteries, fuel cells, electrolyser, and hydrogen tank
were selected in order to dene optimum combination of equipment dimensions. PVhydrogen system components are described in more detail in the following sections.
3.2.1.1. Photovoltaics. In the optimisation process, PV capital, replacement and O&M
costs, as well as component lifetime described under 3.1.1 were used. Previous experience
on modelling of RES-hydrogen stand-alone power systems has shown that the replacement
of diesel generator results in a need for a larger RE source in order to produce adequate
quantities of hydrogen for the fuel cell [1,2]. After some preliminary runs with HOMER it
was decided that the most suitable PV sizes to be considered were 8.8, 15.9 and 19 kW.
3.2.1.2. Batteries. Batteries considered for the PV-hydrogen system were of the same
type with the ones used in the PV-diesel system. Sizes of batteries considered in the
optimisation were: 420, 210 (50% reduction compared to the conventional system) and
0 kWh (no batteries).
3.2.1.3. Fuel cells. The cost of fuel cells varies with type of technology, size, auxiliaries
installed etc. PEM type fuel cells cost range between 3000 and 6000 h kW1, but it is
expected that it will be reduced in the near future [1]. The European Commission (EC)
long-term (2020) cost target for fuel cells in stationary applications is 300 h kW1. The
capital cost of PEM fuel cells used in the optimisation were 3000 h kW1. Fuel cells lifetime
used in the calculations was 15,000 operating hours. Two different PEM fuel cells sizes
(8 and 10 kW) were considered in the calculations performed with HOMER.
3.2.1.4. Electrolyser. Even though alkaline electrolysis is considered a well-established
technology since the 1940s, the cost of electrolysers is still high, due to the lack of massproduction. For larger units, capable of producing 100120 normal (N) m3 h1 the cost per m3
of hydrogen produced reduces and approaches 7300 kh per Nm3 h1 of hydrogen, which is
still high [1,2]. The cost of electrolysis integrated to the proposed PV-hydrogen cost was
8150 h per Nm3 h1 H2. According to water electrolyser manufacturers, mass-production
of electrolysis units is expected to result in at least a 50% cost reduction in the long term

ARTICLE IN PRESS
690

E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

[1]. The introduction of two different electrolyser sizes was investigated with HOMER:
3.2 Nm3 h1 H2 (16 kWe) and 4.2 Nm3 h1 H2 (21 kWe). The lifetime of these units was
considered as equal to 20 years.
3.2.1.5. Hydrogen tank. Hydrogen produced at 30 bar from the electrolyser will be stored
in conventional hydrogen storage tanks, the average cost of which is 38 h Nm3 of
hydrogen stored [1]. It is predicted a 40% cost reduction at 22.8 h Nm3 of hydrogen
stored. Three hydrogen storage tanks options were investigated in the optimisation
process, namely 3000 Nm3 (264 kg), 3400 Nm3 (300 kg) and 5110 Nm3 (450 kg) and the
lifetime was also considered 20 years.
3.2.1.6. PV-hydrogen system optimisation results. According to the results of the
optimisation process for the hydrogen stand-alone power system, which is schematically
shown in Fig. 7, the optimal system comprises a 15.9 kW PV panel, an 8 kW PEM-type fuel
cell, an electrolysis unit of 16 kW, capable of producing ca. 3.2 Nm3 h1 of hydrogen and a
hydrogen storage tank with a capacity of 5110 Nm3 (320 kg) of compressed H2 at 30 bar.
The battery bank is removed. Load following control strategy is optimal in this case, too.
Therefore, the PV array needs to be over dimensioned in relation to the one used in the
PV-diesel system (from 8.8 to 15.9 kW), in order to be able to fully replace the diesel
generator. In addition the battery bank was removed, according to HOMER optimisation
results. Nevertheless in real systems a small battery is always needed in order to take over
the load and serve auxiliaries during fuel cells start-up. According to the optimisation

Fig. 7. HOMER implementation of the PV-hydrogen power system.

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

691

results a DC conguration was economically favourable, and the PV-hydrogen system had
a total annual electrical energy production of 39,347 kWh, 72% of which was produced by
the PV array. HOMER results related to the electricity production and load served by the
optimal hydrogen-based system are given in Table 4.
As it is concluded from Table 4, the optimal hydrogen based system is operating 100%
on renewable energy; diesel fuel and carbon emissions are eliminated (0 kg year1).
Dumped electricity is signicantly reduced (from 20% to 13%) compared to the original
PV-diesel system. The analysis of the proposed PV-hydrogen system showed that the fuel
cell operated 2449 h year1 and consumed approximately 670 kg H2 year1. Its estimated
lifetime was 6.12 years and its average electrical output was 4.56 kW. In Table 5, a
comparison of the component sizes between the already existing PV-diesel and the
proposed hydrogen-based system is presented.
Figs. 8(a) and (b) present the operational characteristics of the PV-hydrogen system (as
calculated with HOMER) namely: electrical demand, power produced from PV, power
produced from fuel cells, electrical power consumption of the electrolysis unit, and excess
electricity for a typical day of month July (when houses served by the system are
inhabited).
Fig. 8(a) shows that the primary electrical load in the daytime is served almost
completely by PV panels, and the fuel cell driven by stored hydrogen gas takes over the
Table 4
Electrical production and demand for the PV-hydrogen System
Annual electricity production
PV-array
Fuel cell
Renewable fraction
Total production

28,177 kWh (72%)


11,170 kWh (28%)
1.000
39,347 kWh

Annual electrical load served


DC primary load served
Electrolyser load served
Total load served

18,248 kWh
16,009 kWh
34,258 kWhO&M

Other
Excess electricity
Capacity shortage
Carbon emissions

5090 kWh (13%)


77.5 kWh
0 kg year1

Table 5
Component sizes of the PV-diesel and the PV-hydrogen system
Component

PV-diesel system

PV-hydrogen system

Photovoltaics
Electrolyzer
H2 storage tank
Diesel generator
Fuel cells
Batteries

8.8 kW

15.9 kW
3.2 Nm3 h1 H2
450 kg

8 kW
8 kW
280 kWh

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

692

14
12

Power (kW)

10
DC load
8

PV
Fuel Cell

6
4
2
0
1

(a)

11 13 15
Time (hr)

17

19

21

23

14
12

Power (kW)

10
8

PV

Electrolyser
consumption

Excess
electricity

2
0
1
(b)

11
16
Time (hr)

21

Fig. 8. (a) Electrical demand and power produced by PV and fuel cells in the hydrogen-based system for a typical
day of month July. (b) Electrolysis unit consumption, power produced by PV and excess electricity in the
hydrogen-based system for a typical day of month July.

electrical load either during nighttime, or in case of cloudiness when there is not adequate
PV power production. The noiseless operation of fuel cells during the night is another
signicant advantage of the proposed system. As described before, excess electricity
produced by PV panels during daytime drives the electrolysis unit and is stored in the form
of hydrogen gas, which supplies PEM fuel cells upon demand. As one can see in Fig. 8(b),
excess electricity produced during summer time is minimised, thus hydrogen is mainly
produced during winter when the settlements are not inhabited and there is not electrical
demand.

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

693

3.2.2. PV-hydrogen system techno-economic analysis


In the techno-economic analysis of the hydrogen stand-alone power system, two
different cost scenarios were used. In the rst one, current costs of hydrogen energy
technologies were used, while in the second one, long-term (2020) costs for hydrogen
technologies components were introduced. Future costs of hydrogen technologies were
based on EU target costs and personal contacts to hydrogen technologies manufacturers
[1,2]. PV-hydrogen system techno-economic analysis was conducted in order to estimate
the future competitiveness of similar hydrogen-based stand-alone power systems. The
present and future cost assumptions are summarised in Table 6. In the long-term scenario,
as mentioned before, a 50% reduction in the cost of electrolysis units and a 40% reduction
in the cost of hydrogen storage tanks, were assumed. The EC target cost for fuel cells in
order to become competitive in stationary applications was considered in the long-term
scenario.
The results of the techno-economic analysis indicated that the replacement of diesel
generators and the battery bank was not economically viable at current costs of hydrogen
energy technology components. Table 7 shows the distribution of annualised capital and
replacement costs for the main components of the hydrogen-based system, using current
costs of hydrogen technologies. PV panels are still an important cost factor of the
proposed system, since they are over dimensioned in order to replace the diesel generator
set, but hydrogen storage tank is the major cost factor. This happens in stand-alone power
systems with seasonal energy storage, because it is needed to install a large hydrogen tank
in order to store the appropriate quantities of hydrogen fuel that will drive the fuel cells.
The total net present cost of the PV-hydrogen system is currently around 360,000 h, which
is 1.7 times higher compared to the total net present cost of the already existing PV-diesel
system. Current cost of energy of the proposed hydrogen system is also high
(1.72 h kWh1).
The analysis of the optimal hydrogen system, using future costs of hydrogen
technologies components, demonstrated that in the long-term (2020) the total net cost
of the power system will be reduced to 236,000 h (which is a 35% reduction compared to
the present cost) and the cost of energy will be approximately 1.129 h kWh1 (i.e. a 34%
reduction). Therefore, the future cost of energy produced by the PV-hydrogen system is
Table 6
Assumed lifetime, operations and maintenance cost, and estimated specic costs for the hydrogen energy system
components today and in 2020 [1]
Hydrogen technology Type
component

Current (20032005)
Lifetime (years)

O&M (% of
inv. costs)

Cost

Electrolyser
Fuel cell
H2 storage

Alkaline
PEM-type
Compressed gas

20
10
20

2.0
2.5
0.5

8150 h per Nm3 h1


3000 h kW1
38 h Nm3

Electrolyser
Fuel cell
H2storage

Alkaline
PEM-type
Compressed gas

Long-term (2020)
20
20
20

1.0
1.0
0.5

4075 h per Nm3 h1


300 h kW1
22.8 h Nm3

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

694

Table 7
Distribution of annualised costs for the main components of the PV-hydrogen system, at current costs of
hydrogen technologies
Component

Annualised
capital
(h year1)

Annualised
replacement
(h year1)

Annual O&M
(h year1)

Annual fuel
(h year1)

Total annualised
(h year1)

PV array
Fuel cell
Electrolyser
Hydrogen tank
Total

9357
2092
1123
16,193
28,766

973
2727
0
0
1755

0
7
258
639
903

0
0
0
0
0

8385
4827
1381
16,832
31,424

Cost of Energy ( / kWh)


1.0

1.60
1.55

H2 Tank Capital Cost Factor

1.50
1.45
1.40
1.35
1.30

0.9

1.25
1.20
1.15

0.8

1.10

0.7

0.6
0.2

0.4

0.6

0.8

1.0

FC Captial Cost Factor


Fig. 9. Cost of Energy of the proposed PV-hydrogen power system as a function of the costs of hydrogen storage
tanks and fuel cells. Hydrogen storage tank and fuel cell capital costs are expressed as a percentage of their current
costs.

lower than the current cost of energy produced by the PV-diesel system (which is
1.135 h kWh1) and the proposed hydrogen based system becomes economically
favourable. Sensitivity analysis demonstrated that the capital costs of hydrogen storage
devices and PEM fuel cells signicantly affect the nal cost of energy produced by a
hydrogen stand-alone power system. The cost of energy produced by the PV-hydrogen
system as a function of hydrogen storage tank costs and PEM fuel cells costs (for a xed
cost of electrolyser at 50% of its current cost) is given in Fig. 9.
The results of the sensitivity analysis showed that the PV-hydrogen power system will be
economically viable in the long-term (2020), when the above mentioned reductions in cost
of hydrogen technologies (hydrogen storage tanks, electrolysers and PEM fuel cells) will
take place. In the analysis it was assumed that the diesel fuel price and the PV capital costs

ARTICLE IN PRESS
E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

695

will remain constant. Moreover external costs were not introduced. Taking into account a
further increase on the fossil fuels price, a reduction on the PV capital costs and
introducing external costs (emission taxes etc.) as well, we can safely conclude that such a
RES-hydrogen stand-alone power system will become economically viable even earlier
than 2020.
4. Conclusions
The replacement of conventional technologies, namely diesel generators and/or batteries
by hydrogen technologies including fuel cells in RE based stand-alone power systems is
technologically feasible, reduces emissions, noise and fossil fuel dependence and increases
RE penetration. Dumped excess energy is also reduced in such power systems. The
optimisation results of the envisaged hydrogen based system showed that the replacement
of diesel generator sets and battery bank in such a system can be accomplished by over
dimensioning the RE technology (PV array in the presented case). On the other hand, the
techno-economic analysis of the hydrogen-based system revealed that this system will be
economically favourable compared to the PV-diesel system as long as a 50% reduction on
the cost of electrolysers, a 40% reduction on the cost of hydrogen tanks are made and the
EC target of 300 h kW1 for fuel cells in stationary applications is achieved. This is in
agreement to the predictions of the European roadmap for Hydrogen (prepared by the
high level group (HLG) for hydrogen and fuel cells) that one of the rst commercial
applications of stationary fuel cells will be niche markets, in which non interconnected
remote applications are included [20].
References
[1] Zoulias EI, Glockner R, Lymberopoulos N, Tsoutsos T, Vosseler I, Gavalda O, Mydske HJ, Taylor P.
Integration of hydrogen energy technologies in stand-alone power systems analysis of the current potential
for applications. Ren Sust Energy Rev 2006;10(5):43262.
[2] Lymberopoulos N, Zoulias EI, Taylor P, Little P, Brodin M, Glockner R, et al. Market potential for the
introduction of hydrogen in stand-alone power systems. In: Proceedings of rst European hydrogen energy
conference, Grenoble, France, 2003.
[3] Schucan TH. International energy agency hydrogen implementing agreement task 11integrated systems
rep. IEA/H2/T11/FR1-2000, 1999. p. 2340, 63101.
[4] Vanhanen J. Operating experience on a self sufcient solar-H2-fuel cell system. In: Proceedings of second
NsonH and fuel cells for hydrogen storage conference, 1995. p.4654.
[5] Voss K, Goetzberger A, Bopp G, Haberle A, Heinzel A, Lehmberg H. The self-sufcient solar house in
Freiburg, results of 3 years operation. Sol Energy 1996;58:1723.
[6] Miland H. PV/hydrogen stand-alone power system. PhD. dissertation, NTNU, Norway, 2004.
[7] Varkaraki E, Lymberopoulos N, Zachariou A. Hydrogen based emergency back-up system for
telecommunication applications. J Power Sources 2003;118:1422.
[8] Ulleberg O. The importance of control strategies in PV-hydrogen systems. Sol Energy 2004;76:3239.
[9] Vosen SR, Keller JO. Hybrid energy storage systems for stand-alone electric power systems: optimization of
system performance and cost through control strategies. Int J Hydrogen Energy 1999;24:113956.
[10] Agbossou K, Chahine R, Hamelin J, Laurencelle F, Anouar A, St-Arnaud JM, et al. Renewable energy
systems based on hydrogen for remote applications. J Power Sources 2001;96:16872.
[11] Kelouwani S, Agbossou K, Chahine R. Model for energy conversion in renewable energy system with
hydrogen storage. J Power Sources 2005;140:3929.
[12] Wallmark C, Alvfors P. Technical design and economic evaluation of a stand-alone PEFC system for
buildings in Sweden. J Power Sources 2003;118:35866.

ARTICLE IN PRESS
696

E.I. Zoulias, N. Lymberopoulos / Renewable Energy 32 (2007) 680696

[13] Santarelli M, Cali M, Macagno S. Design and analysis of stand-alone hydrogen energy systems with different
renewable sources. Int J Hydrogen Energy 2004;29:157186.
[14] Khan MJ, Iqbal MT. Pre-feasibility study of stand-alone hybrid energy systems for applications in
Newfoundland. Renew Energy 2005;30:83554.
[15] Strauss P, Wurtz RP, Haas O, Ibrahim M, Reekers J, Tselepis S, et al. Stand-alone AC PV systems and
microgrids with new standard power componentsrst results of two European Joule projects PV-MODE
and MORE. In: Proceedings of sixteenth European photovoltaic solar energy conference, Glasgow, UK,
2000.
[16] HOMER National Renewable Energy Laboratory (NREL), 617 Cole Boulevard, Golden, CO 80401-3393,
URL: http://www.nrel.gov/homer
[17] Ballard Power Systems Inc., 4343, North Fraser Way, Burnaby, BC V5J 5J9, Canada, URL: http://
www.ballard.com
[18] RETScreenTM database, URL: http://www.retscreen.net
[19] Greek Ministry of Development, Average diesel prices, http://www.ypan.gr/ash_fuel/kafsima/MESES_
TIMES.htm
[20] European Hydrogen and Fuel Cell Technology Platform, http://www.hfpeurope.org/

Potrebbero piacerti anche