Vous êtes sur la page 1sur 11

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

Available online at www.sciencedirect.com

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

Energy and exergy analyses of hydrogen production via


solar-boosted ocean thermal energy conversion and PEM
electrolysis
Pouria Ahmadi*, Ibrahim Dincer, Marc A. Rosen
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology (UOIT), Oshawa, Ontario, Canada L1G8B8

article info

abstract

Article history:

Energy and exergy analyses are reported of hydrogen production via an ocean thermal

Received 10 August 2012

energy conversion (OTEC) system coupled with a solar-enhanced proton exchange

Received in revised form

membrane (PEM) electrolyzer. This system is composed of a turbine, an evaporator,

29 October 2012

a condenser, a pump, a solar collector and a PEM electrolyzer. Electricity is generated in the

Accepted 4 November 2012

turbine, which is used by the PEM electrolyzer to produce hydrogen. A simulation program

Available online 17 December 2012

using Matlab software is developed to model the PEM electrolyzer and OTEC system. The
simulation model for the PEM electrolyzer used in this study is validated with experimental

Keywords:

data from the literature. The amount of hydrogen produced, the exergy destruction of each

Energy

component and the overall system, and the exergy efficiency of the system are calculated.

Exergy

To better understand the effect of various parameters on system performance, a para-

Hydrogen production

metric analysis is carried out. The energy and exergy efficiencies of the integrated OTEC

OTEC

system are 3.6% and 22.7% respectively, and the exergy efficiency of the PEM electrolyzer is

Water electrolysis

about 56.5% while the amount of hydrogen produced by it is 1.2 kg/h.

Solar energy

Crown Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All
rights reserved.

1.

Introduction

Hydrogen as an energy carrier can facilitate sustainable


energy systems. The development of sustainable carbonneutral energy sources has become one of the most significant issues in the world today. Hydrogen can be produced
from various energy sources using methods like biomass
conversion, steam methane reforming and water splitting.
Hydrogen can be produced in a relatively environmentally
benign manner (depending on the source of the input energy)
via splitting water by photocatalysis, thermochemical cycles
and electrolysis. Currently, both thermochemical and photocatalysis hydrogen production are not economically

competitive. Water electrolysis is a mature technology for


large scale hydrogen production. Hydrogen production by
proton exchange membrane (PEM) electrolysis has numerous
advantages, such as low environmental impact and easy
maintenance.
A large amount of solar energy is stored as heat in the
surface waters of the worlds oceans, providing a source of
renewable energy. Ocean thermal energy conversion (OTEC) is
a process for harnessing this renewable energy in which
a heat engine operates between the relatively warm ocean
surface, which is exposed to the sun, and the colder (about
5  C) water deeper in the ocean, to produce electricity. OTEC
usually incorporates a low-temperature Rankine cycle engine

* Corresponding author. Tel.: 1 289 600 9505.


E-mail addresses: pouryaahmadi81@gmail.com, Pouria.Ahmadi@uoit.ca (P. Ahmadi), Ibrahim.Dincer@uoit.ca (I. Dincer), Marc.
Rosen@uoit.ca (M.A. Rosen).
0360-3199/$ e see front matter Crown Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2012.11.025

1796

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

which boils a working fluid such as ammonia to generate


a vapour which turns the turbine to generate electricity, and
then is condensed back into a liquid in a continuous process.
Eighty percent of the energy the Earth receives from the sun is
stored in the Earths oceans [1,2], and many regions of the
world have access to this OTEC resource. OTEC can produce
fuels by using its product electricity to create hydrogen, which
can be used in hydrogen fueled cars as well as in the development of synthetic fuels. For a small city, millions of tons of
CO2 are generated annually through fossil fuel use while with
OTEC the value is zero, during the operation of devices. OTEC
has a potential to replace some fossil fuel use, perhaps via
OTEC ships travelling the seas of the world.
An OTEC system utilizes low-grade energy and has low
energy efficiency (about 3e5%). Therefore, achieving a high
electricity generating capacity with OTEC requires the use of
large quantities of seawater and a correspondingly large
amount of pumping energy. These factors impact negatively
on the cost-effectiveness of this technology and OTEC is not
commercially viable today. To improve the effectiveness and
economics of OTEC cycles, it is proposed that they being
integrated with industrial operations so that, apart from
generating electricity, they could be used for fresh water
production, air conditioning and refrigeration, cold water
agriculture, aquaculture and mariculture, and hydrogen
production [1]. Potential markets for OTEC have been identified, most of which are in Pacific Ocean, and about 50 countries are examining its implementation as a sustainable
source of energy and fresh water, including India, Korea,
Palau, Philippines, the U.S. and Papua New Guinea [3]. In 2001,
as a result of cooperation between Japan and India a 1-MW
OTEC plant was built in India [3], and others are planned to
be constructed in the near future [4].
Considerable research has been directed to the development of OTEC recently. Uehara [5e8] conducted numerous
theoretical and experimental studies on the major components of an OTEC plant, and showed that ammonia is a suitable working fluid for an OTEC plant employing a closed
organic Rankine cycle (ORC). The energy efficiency of the
Rankine cycle in an OTEC plant is usually limited to around 5%
due to the small temperature difference between surface
water and deep water of the ocean. So, to improve the efficiency of OTEC, other thermodynamic cycles like the Kalina
cycle and the Uehara cycle that use an ammoniaewater
mixture as the working fluid are being considered [9]; they are
reported to have better energy efficiencies rather than
a Rankine cycle at the same temperature difference [9].
Increasing in the temperature difference between the hot
heat source and the cold heat sink can improve the efficiency
of OTEC plants, as can the integration of OTEC with other
energy technologies. Saitoh and Yamada [10] proposed
a conceptual design of a multiple Rankine cycle system using
both solar thermal energy and ocean thermal energy in order
to improve the cycle efficiency. In addition, Owens [11] optimized OTEC plants using ammonia as the working fluid in
closed cycles. Uehara and Ikegamia [6] performed an optimization study using the power method for a closed-cycle OTEC
plant. Later, Uehara and Ikegamia [12] studied the performance of OTEC using the Kalina cycle and proposed a new
OTEC cycle with an absorption and extraction process [13].

Rabas et al. [14] studied a non-condensable gas removal


system for a particular 10-MW hybrid power plant. More
recently, Kazim [15] proposed a hydrogen production cycle
using ocean thermal energy, and Ikegamia et al. [7,16]
proposed desalination using an integrated hybrid cycle and
an open OTEC cycle.
Energy analysis, which is based on the first law of thermodynamics, often does not provide a clear picture of thermodynamic efficiency and losses. Exergy is not subject to
a conservation law, but is destroyed due to irreversibilities
during any process [17]. Exergy analysis overcomes such
deficiencies and can help improve efficiency and sustainability [18]. Exergy is a useful tool for determining the location,
type and true magnitude of exergy losses, which appear in the
form of either exergy destruction or waste exergy emission
[19]. Therefore, exergy methods can assist in developing
strategies and guidelines for more effective use of energy
resources and technologies. Sun et al. [20] report an optimization design and exergy analysis of an organic Rankine cycle
in ocean thermal energy conversion, and show that ammonia
is a good working fluid for an ORC cycle in OTEC from a net
power output point of view. Also, from an exergy viewpoint, it
has been shown that a larger scale ORC in OTEC can justify the
use of a heat exchanger with a higher effectiveness to
decrease the exhaust loss, which accounts for a large
proportion of the exergy loss of an ORC in an OTEC system [20].
In this paper, a novel OTEC system equipped with a solar
collector and PEM electrolysis to produce hydrogen is thermodynamically modelled and assessed with energy and
exergy analyses. The primary objective is to improve understanding of this system. The following specific tasks are
performed:
 Model and simulate (using Matlab software) the integrated
system.
 Validate each part of the model and simulation.
 Perform energy and exergy analyses of the integrated
system to determine the exergy destruction rate and energy
and exergy efficiencies of each component and the entire
system.
 Determine the sustainability index for the system and
quantify its variation with exergy destruction, and examine
the relation between exergy efficiency, exergy destruction
and sustainability index.
 Conduct a comprehensive parametric study to determine
the effect of major design parameters on system
performance.

2.

Modelling and energy analysis

For thermodynamic modelling purposes, the integrated OTEC


system for hydrogen production considered here (Fig. 1) is
divided into three parts: flat plate solar collector, ocean
thermal energy conversion (OTEC) unit and PEM electrolyzer.
Fig. 1 shows a schematic diagram of an integrated OTEC
system equipped with a flat plate solar collector and PEM
electrolyzer. This integrated system uses the warm surface
seawater to evaporate a working fluid like ammonia or a Freon

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

1797

Fig. 1 e Schematic of ocean thermal energy conversion (OTEC) with a flat plate solar collector and PEM electrolyzer for
hydrogen production.

refrigerant, which drives a turbine to produce electricity,


which in turn is used to drive a PEM electrolyzer to produce
hydrogen. After passing through the turbine, the vapour is
condensed in a heat exchanger that is cooled by cold deep
seawater. The working fluid is then pumped back through the
warm seawater heat exchanger, and the cycle is repeated
continuously.
Energy and exergy analyses are used to determine the
temperature profile in the plant, input and output enthalpy
and exergy flows, exergy destructions rates and energy and
exergy efficiencies. The relevant energy balances and governing equations for the main sections of the plant shown in
Fig. 1 are described in the following subsections.

2.1.

Energy analysis

2.1.1.

Flat plate solar collector

S saI

(2)

where T0 is the ambient temperature and the FR is heat


removal factor which is defined as:

(3)

(4)

where sa is optical efficiency and I is solar radiation intensity. The energy efficiency of the solar flat plate collector is
expressed as:
Q_ u
IAP

(5)

More information about solar flat plate collector modeling is


detailed in Ref. [22].

2.1.2.
(1)

_ are the water outlet temperature, inlet


where T3, T2, Cp and m
temperature, specific heat at constant pressure and mass flow
rate. The HotteleWhillier equation for the heat gained by the
flat plate collector considering heat losses from the collector is
calculated as [21]:
Q_ u AP FR S  Ul Tin  T0 

_ p
mC
0
_ pg
1  efF Ul AP =mC
Ul AP

here, F0 is collector efficiency factor which is around 0.914 for


this case [21] and Ul is the overall collector loss coefficient
obtained from [21]. In Eq. (2), radiation flux absorbed by the
absorber is calculated as:

As shown in Fig. 1, water enters the solar collector at point 2


and is heated by the collector. The useful heat gained by the
working fluid can be written as:
_ p T3  T2
Q_ u mC

FR

Organic rankine cycle (ORC)

As shown in Fig. 1, the electricity production unit is based on


an organic Rankine cycle which is suitable for low-grade heat.
Fig. 2 shows the corresponding temperatureeentropy (TeS )
diagram of the ORC.
The net power output of the system is expressible as:


_G W
_ WS W
_ CS W
_ WF
_ net W
W

(6)

_ WS and W
_ CS are
_ G is the turbine generator power, W
where W
_ WF is
the warm and cold seawater pumping power, and W
working fluid working power.

1798

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

Fig. 2 e TeS diagram of the ocean thermal energy conversion (OTEC) for hydrogen production.

2.1.2.1. Turbine generator power. The turbine power output is


calculated by writing the energy balance for a control volume
around a turbine which is the product of working fluid mass
_ F and the adiabatic heat loss between the evapoflow rate, m
rator and the condenser, as follows:
_Gm
_ F hT hG h5  h6
W

(7)

here, hT and hG are the turbine isentropic efficiency and


generator mechanical efficiency.

2.1.2.2. Warm seawater pumping power. The warm seawater


pumping power can be written as:
_ WS DHWS g
_ WS m
W
hWSP

(8)

where DHWS is the total pump head difference of the warm


seawater piping:
DHWS DHWS P DHWS E

(9)

here, DHWS P is the pump head of the warm seawater pipe,


which can be given as [23]:
DHWS P DHWS SP DHWS B

(10)

where DHWS SP is the friction loss of the straight pipe and


DHWS B is the bending loss on the warm seawater pipe, which
are given as [23]:
DHWS SP

1:85

LWS
VWS
6:82 1:17 
;
dWS
CWS

DHWS B

lm

2
VWS
2g

CWS 100

(11)

(12)

here, LWS is the length of the warm seawater pipe, dWS is the
warm seawater inner pipe diameter and VWS is the velocity of
warm seawater inside the pipe. Also, DHWS E is the pressure
difference of warm seawater in the evaporator, expressible as:
DHWS E lE

2
VWS
L
 E
2g Deq W

(13)

here, LE is the length of the evaporator plate and lE is taken


from Ref. [23]. Also, Deq is the equivalent diameter, which is
calculated as:
Deq 2d

(14)

where d is the clearance.

2.1.2.3. Cold seawater pumping power. The cold seawater


pumping power is expressed as:
_
_ CS mCS DHCW g
W
hCSP

(15)

where hCSP is the cold seawater pump efficiency and DHCW is


the total pump head of the cold seawater piping, given as:
DHCS DHCS P DHCS C DHCS d

(16)

here DHCS P is the pump head of the cold seawater pipe [23]:
DHCS P DHCS SP DHCS B

(17)

These two terms are similar to terms in Eqs. (11) and (12). Also,
DHCS C is the cold seawater pressure difference in the
condenser, defined as:
DHCS C lC

2
VCS
L
 EC
2g Deq C

(18)

where LC is the length of the evaporator plate and lC is taken


from reference [23]. Also, Deq is the equivalent diameter given
in Eq. (14) and DHCS d is the pressure difference caused by the
density difference between the warm seawater surface and
cold deeper seawater, calculated as [23]:
DHCS d LCS 

2.1.3.



1 1
rWS rCS LCS
rCS 2

(19)

PEM electrolyzer

The schematic diagram of the PEM electrolyzer for H2


production is shown on the right side of Fig. 1. During electrolysis, the required electricity and heat are both supplied to
the electrolyzer to drive the electrochemical reactions. As

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

shown in Fig. 1, liquid water is fed to the PEM electrolyzer at


ambient temperature. Liquid water enters a heat exchanger
that brings it to the PEM electrolyzer temperature and it then
enters the electrolyzer. Leaving the cathode, the H2 produced
dissipates heat to the environment and cools to the reference
environment temperature. The oxygen gas produced at the
anode is separated from the water and oxygen mixture and
then cooled to the reference environment temperature. The
remaining water is then returned to the water supply stream
for the next hydrogen production cycle. The overall PEM
electrolysis reaction is simply water splitting, i.e., electricity
and heat are used to separate water into produce hydrogen
and oxygen. Hydrogen is stored in a tank for later usage.

2.1.3.1. Thermochemical modeling of a PEM electrolyzer.


Energy and exergy analyses of a PEM electrolyzer can be performed in conjunction the electrochemical modeling. The
total energy needed by the electrolyzer can be obtained
noting:
DH DG TDS

(20)

where DG is Gibbs free energy and TDS is the thermal energy


requirement. The values of G, S, and H for hydrogen, oxygen
and water can be obtained from thermodynamic tables [24].
The total energy need is the theoretical energy required for
H2O electrolysis without any losses. The catalyst used in PEM
electrolysis provides an alternative path for the reaction with
lower activation energy. The mass flow rate of hydrogen is
determined by [25]:
J
N_ H2 ;out
N_ H2 O;reacted
2F

(21)

where J is the current density and F is the Faraday constant.


The electric energy input rate to the electrolyzer can be
expressed as:
Eelectric JV

(22)

where Eelectric is the electric energy input and Exelectric the


electric exergy input. Also, V is given as:
V V0 Vact;a Vact;c Vohm

(24)

Also, Vact;a , Vact;c and Vohm are the activation overpotential of


the anode, the activation overpotential of the cathode, and the
ohmic overpotential of the electrolyte, respectively. Ohmic
overpotential in the proton exchange membrane is caused by
the resistance of the membrane to the hydrogen ions transporting through it. The ionic resistance of the membrane
depends on the degree of humidification and thickness of the
membrane as well as the membrane temperature [26]. The
local ionic conductivity sx of the proton exchange
membrane can be expressed as [25,26]:



1
1

sPEM lx 0:5139lx  0:326exp 1268
303 T

where x is the distance in the membrane measured from the


cathode-membrane interface and lx is the water content at
a location x in the membrane. The value of lx can be
calculated in terms of the water content at the membraneelectrode edges:
lx

la  lc
x lc
D

(26)

where D is the membrane thickness, and la and lc are the


water contents at the anode-membrane and the cathodemembrane interfaces, respectively. The overall ohmic resistance can thus be expressed as [25]:
ZD
RPEM
0

dx
sPEM lx

(27)

Based on Ohms law, the following equation can be written


for the ohmic overpotential:
Vohm;PEM JRPEM

(28)

The activation overpotential, Vact, caused by a deviation of


net current from its equilibrium, and also an electron transfer
reaction, must be differentiated from the concentration of the
oxidized and reduced species [25,26]. Then,
Vact;i



RT
J
1
; i a; c
sinh
F
2J0;i

(29)

here, J0 is the exchange current density, which is an important


parameter in calculating the activation overpotential. It
characterizes the electrodes capabilities in the electrochemical reaction. A high exchange current density implies
a high reactivity of the electrode, which results in a lower
overpotential. The exchange current density for electrolysis
can be expressed as [26]:


Eact;i
; i a; c
J0;i Jref
i exp 
RT

(30)

where Jref
is the pre-exponential factor and Eact;i is the actii
vation energy for the anode and cathode. More details about
PEM electrolysis modeling can be found elsewhere [26, 27].

(23)

where V0 is the reversible potential, which is related to the


difference in free energy between reactants and products, and
V0 can be obtained by the Nernst equation as follows:
V0 1:229  8:5  104 TPEM  298

1799

(25)

3.

Exergy analysis

Exergy analysis can help to develop strategies and guidelines


for more effective use of energy, and has been applied to
various thermal processes, including power generation, CHP
and trigeneration. Exergy can be divided into four components: physical, chemical, kinetic and potential. For the
processes involved in this study, the latter two are neglected
since changes in elevation and speed are negligible [28,29].
Physical exergy is defined as the maximum work obtainable as
a system interacts with a reference environment at an equilibrium state [21,22]. Chemical exergy is associated with the
departure of the chemical composition of a system from the
chemical equilibrium of a reference environment. Chemical
exergy is important in combustion evaluation. Applying the
first and the second laws of thermodynamics, the following
exergy balance for each component is obtained:

1800

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

Table 1 e Input data for the system simulation.


Parameter

Value

Parameter

Value

0.80
0.90
0.78
0.80
25
500
22
4

Warm seawater mass flow rate (kg/s)


Cold seawater mass flow rate (kg/s)
Cold seawater pipe length (m)
Cold seawater pipe inner diameter (m)
Warm seawater pipe length (m)
Warm seawater pipe length (m)
Solar collector effective area (m2)
Electrolyzer working temperature ( C)

150
150
1000
0.70
50
0.70
5000
80

Turbine isentropic efficiency, hT


Generator mechanical efficiency, hG
Working fluid pump isentropic efficiency, hWFP
Seawater pumps isentropic efficiency, hP
Ambient temperature ( C)
Solar radiation incident on collector surface, I (W/m2)
Warm seawater temperature, TWSI ( C)
Cold seawater temperature at depth of 1000 m, TCSI ( C)

ExQ

_ i exi
m

_ e exe ExW ExD


m

(31)

where the subscripts e and i denote the inlet and outlet of


$
a control volume, ExD is the exergy destruction and other
terms are as follows [28e31]:


$
T0 _
Qi
ExQ 1 
Ti

(32)

_
Exw W

(33)

ex exph exch

(34)

_ net
W
_
Q EVP

_ net is given in Eq. (6), and Q_ EVP is expressible as


where W
_ WF h1  h4
Q_ EVP m

3.1.2.

(35)

Here, ExQ and Exw respectively are the corresponding exergy


rates associated with heat transfer and work across the
boundary of a control volume, T is the absolute temperature
and the subscript o refers to the reference environment
conditions. The reference environment considered in our
study is T0 20  C and P0 1 bar. The specific chemical exergy
for gas mixtures is defined as follows [28,29]:
exch

xk exkch RT0

xk lnxk

1
DP

(37)

where DP is a depletion number, which is defined as the exergy


destruction/input exergy [21,29]. This relation demonstrates
how reducing a systems environmental impact can be achieved by reducing its exergy destruction.

3.1.

Efficiencies

3.1.1.

Energy efficiency

Exergy efficiency

_ net
W
$

Exin

_ net
W

(40)

Exin;WS Exin;CS

$




_ in;WS hWS;in  h0  T0 sWS;in  s0
Exin;WS m

(41)

$




_ in;CS hCS;in  h0  T0 sCS;in  s0
Exin;CS m

(42)

here, the reference environment state is taken to be


P0 1.01 bar and T0 298.15 K.

(36)

Further details about each term in this equation are given in


Refs. [28e31].
To improve environmental sustainability, it is necessary
not only to use sustainable or renewable sources of energy,
but also to utilize non-renewable sources like natural gas
more efficiently, while minimizing environmental damage. In
this way, society can reduce its use of limited resources and
extend their lifetimes. Here, a sustainability index SI is used to
relate exergy with environmental impact [30,32,33]:
SI

(39)

The exergy efficiency is defined as the product exergy output


divided by the exergy input [23]. According to Nag and Dai
[34,35], the exergy efficiency of the ORC power generation
cycle in an OTEC system is given as:
J

exph h  h0  T0 s  s0

(38)

The energy efficiency of the OTEC system is defined as the net


power output of the system divided by input energy at evaporator, which can be expressed as:

4.

Results and discussion

4.1.

Modelling validation and simulation code results

To model the integrated OTEC system for hydrogen production, a simulation code using Matlab software is developed.

Table 2 e Input parameters used to model PEM


electrolysis [26,27].
Parameter
PO2 (atm)
PH2 (atm)
TPEM ( C)
Eact,a (kJ/mol)
Eact,c (kJ/mol)
la
lc
D (mm)
2
Jref
a (A/m )
2
Jref
(A/m
)
c
F (C/mol)

Value
1.0
1.0
80
76
18
14
10
100
1.7  105
4.6  103
96,486

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

1801

Table 3 e Parameter values resulting from energy and


exergy analyses of the system.
Parameters

Value

_ net (kW)
Net power output, W
Exergy efficiency, J (%)
Energy efficiency, h (%)
Sustainability index, SI
$
Total exergy destruction rate, ExD;tot (kW)
_ H2 (kg/h)
Hydrogen production rate, m

101.96
22.70
3.60
1.29
42.12
1.20
56.34
1.30
3.13
0.88

PEM electrolyzer exergy efficiency, JPEM (%)


_ WS (kW)
Warm surface pump power, W
_ CS (kW)
Cold surface pump power, W
_ WF (kW)
Working fluid pump power, W

Fig. 3 e Comparison of present model with experimental


data.

Three main parts are first individually modelled including the


each exergy flow rate. Engineering equation solver (EES) is
linked to Matlab to calculate the properties of the different
working fluids (i.e., water and ammonia) such as pressure,
temperature, enthalpy and entropy. Several simplifying
assumptions are made here to render the analysis more
tractable, while retaining adequate accuracy to illustrate the
principal points of the study
 All processes operate at steady state.
 The thermodynamic cycle of the integrated system in Fig. 1
is an ideal saturated Rankine cycle using pure ammonia as
the working fluid.
 All the components are adiabatic.
 Pressure drops in ORC cycle are negligible.
 State 5 is saturated vapour.
 Heat losses from piping and other auxiliary components are
negligible.
To conduct the simulation, input data are required. For
each subsystem some reliable data are input to the simulation
code in order to determine the outputs. Table 1 lists the input
parameters for the OTEC system simulation. Also, Table 2 lists
the parameter used to simulate the PEM electrolyzer.

To ensure the accuracy and validity of the developed


computer simulation code, the PEM electrolyzer is validated
with experimental data from the literature. Specifically, the
electrochemical model is used to simulate experiments published in the literature and the modeling results and experimental data are compared. The electrolyte used in the
experiments [36,37] is Nafion, a polymer widely used as
electrolyte in fuel cells and electrolyzers. The thicknesses of
the electrolytes tested by Ioroi et al. [36] and Millet et al. [37]
were 50 mm and 178 mm, respectively. Platinum was used as
the electrode catalyst. The simulation code for the JeV characteristics for PEM electrolysis are compared with experimental data of Ioroi et al. [36] as shown in Fig. 3. The modeling
results agree well with the experimental data, supporting the
accuracy of the present model. It is found that the cell
potential increases rapidly when current density is less than
300 A/m2. When J exceeds 300 A/m2, the cell potential
increases slightly with J. To enhance the understanding of the
electrochemical performance of the PEM electrolyzer, ohmic
and activation overpotentials are examined and shown
individually in Fig. 4. This figure shows that the ohmic overpotential is very small and increases slightly with current
density. This observation is attributable to the fact that the
membrane ionic conductivity (sPEM) is significantly high for
typical values of l as well as the operating temperature,
which leads to a lower RPEM (see Eqs. (25) and (27)), which in
turn, lowers the overall ohmic resistance and the ohmic
overpotential.

0.7

Overpotential(V)

0.6
Ohmic activation
Cathode activation
Anode activation

0.5
0.4
0.3
0.2
0.1
0
0

1000

2000

3000

4000

5000

6000

7000

8000

Current density, J (A/m2)

Fig. 4 e Variation of electrolyzer overpotentials at various


current densities.

Fig. 5 e Exergy destruction rate and dimensionless exergy


destruction ratio for each component of the ocean thermal
energy conversion (OTEC) system for hydrogen production.

1802

0.32
50
0.3
45
0.28
40

0.26
0.24
400

440

480

520

560

35
600

Sustainability index, SI

55

Total exergy destruction rate (kW)

0.34

Exergy efficiency

1.6

60

1.55

55

60

0.36

1.5
50
1.45
45
1.4

1.3
400

4.2.

Exergy analysis results

Results of the thermodynamic modelling and exergy analysis


are presented here, including assessments of the effects of
varying several design parameters on the cycle performance.
As already discussed, the inputs of the simulation program
are transferred to the developed code in order to calculate the
outputs. The results of the simulation program are listed in
Table 3. The net power output is seen to be around 101 kW,
which leads to a hydrogen production rate of about 1.2 kg/h. In
addition, the exergy efficiency of the integrated OTEC system
is much higher than the energy efficiency, mainly due to the
fact that the work is produced using a low-grade (in terms of
a temperature near to that of the reference environment) heat
at the ocean surface.
The exergy analysis results are presented in Fig. 5 and
show that the highest exergy destruction occurs in the
condenser, which is mainly due to the temperature difference
between two fluid streams passing through it, but also due to
the pressure drop across the device. Another important result
relates to the dimensionless exergy ratio, which is useful for
prioritizing exergy losses in an intuitive manner. Both the
exergy destruction rate and the dimensionless exergy

440

480

520

560

35
600

Solar radiation intensity (W/m2)

Solar radiation intensity (W/m2)

Fig. 6 e Variation with solar radiation intensity of the OTEC


exergy efficiency and total exergy destruction rate.

40

1.35

Total exergy destruction rate (kW)

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

Fig. 8 e Variation with solar radiation intensity of the


sustainability index and total exergy destruction rate.

destruction ratio are higher in the condenser than in other


components, suggesting that it would likely be worthwhile to
focus improvement efforts on this component.
Since the solar radiation intensity changes during the day,
the variation of system performance is investigated with solar
radiation. Fig. 6 shows the variation of exergy efficiency and
exergy destruction rate of the OTEC system for various values
of solar radiation intensity. Increasing solar radiation intensity results in an increase in the exergy efficiency of the OTEC
system, suggesting that during such a day, the solar intensity
increases the inlet temperature entering the evaporator,
which leads to an increase in enthalpy at point 3 entering the
turbine to produce work. The higher the temperature difference between cold and hot surfaces the more the work. Since
the cold surface temperature is almost constant at the depth
of the ocean, an increase in evaporator inlet temperature
results in an increase in the output power. Also, Fig. 6 shows
that an increase in solar radiation leads to a decrease in total
exergy destruction.
The effect of solar radiation intensity on exergy destruction
rate for each component is shown in Fig. 7. As shown in this
figure, an increase in solar radiation intensity has major
effects on the exergy destruction rates of turbine and
condenser while the changes in exergy destruction rate for the

130

2
mH2(kg/hr)

Wnet(kW)

1.9

120

110

1.7
1.6

100

mH2 (kg/hr)

Wnet (kW)

1.8

1.5
90
1.4
80
400

440

480

520

560

1.3
600

Solar radiation intensity (W/m2)

Fig. 7 e Variation with solar radiation intensity of the


exergy destruction rate of components.

Fig. 9 e Variation with solar radiation intensity of the OTEC


net power output and hydrogen production rate.

1803

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

1.4

ex

1.36

SI
Exergy efficiency

0.24

1.32

0.22
1.28
0.2
1.24
0.18
1.2

0.16

1.16

0.14
0.12
280

284

288

292

50

Sustainability index (SI)

0.26

Total exergy destruction rate (kW)

0.28

1.12
300

296

46

42

38

34

30
280

284

288

T 0 (K)

292

296

300

T0 (K)

Fig. 10 e Variation with ambient temperature of the exergy


efficiency and sustainability index.

Fig. 12 e Variation with ambient temperature of the total


exergy destruction rate.

evaporator and the pump are not drastic. So, since the
decrease in the exergy destruction rate of the condenser is
much higher than that for the turbine, the total exergy
destruction rate decreases with an increase in solar radiation
intensity. Fig. 8 shows the effect of solar radiation intensity on
both sustainability index (SI) and total exergy destruction rate.
The results are similar to those in Fig. 6, i.e., the overall exergy
destruction of the cycle decreases and the sustainability index
increases with increasing solar radiation intensity. The exergy
efficiency, exergy destruction and sustainability are thus
observed to be linked in such systems, supporting the utility of
exergy and environmental analyses.
Fig. 9 shows the effect of solar radiation intensity on
hydrogen production mass flow rate. As the solar intensity
increases the hydrogen production increases since the turbine
work increases and correspondingly the electrical input to
PEM electrolysis increases.
Another major parameter in an OTEC plant is ambient
temperature because it affects the surface temperature and
also the value for inlet exergy at each point of the plant. Hence,
this temperature is an important parameter. Fig. 10 shows the
variation of exergy efficiency and sustainability index with
ambient temperature. As can be seen, an increase in ambient
temperature first leads to an increase in exergy efficiency and

sustainability index, due to an increase in the fluid temperature entering the evaporator which increases the work of the
turbine. Above 285 K the exergy efficiency decreases because
the exergy input to the system, which is the denominator of
Eq. (33), increases, so two different effects occur and the effect
of an increase in inlet exergy values dominates so the exergy
efficiency of the OTEC system decreases. The variation of
exergy destruction rate of each component of the OTEC with
ambient temperature is shown in Fig. 11, which demonstrates
the same trend as Fig. 10. An increase in T0 has major effects on
the exergy destruction rate for the turbine and the condenser,
while for the evaporator and working fluid pump it is almost
constant. An increase in temperature for the condenser first
leads to an increase in exergy destruction rate and then above
285 K the exergy destruction rate decreases. However, the
exergy destruction rate always increases for the turbine. The
summation of these two effects results in an increase in the
total exergy destruction rate first then it leads to a decrease in
the total exergy destruction as shown in Fig. 12.
To enhance the analysis, the effect of condenser temperature on the exergy efficiency of the OTEC system is also
investigated (see Fig. 13). An increase in condenser temperature is seen to reduce the exergy efficiency of the system.
Increasing condenser temperature leads to an increase in
enthalpy at point 2 and results in a reduction in turbine work,

Condenser
Working fluid pump

60

0.32

50

0.28

20
15
10

30
0.2
20
0.16

5
0
280

40
0.24

284

288

292

296

300

T0 (K)
Fig. 11 e Variation with ambient temperature of the exergy
destruction rate of each component.

0.12
5

10

11

13

Total exergy destruction rate (kW)

25

Turbine
Evaporator

Exergy efficiency

Exergy destruction rate (kW)

30

0
15

Condenser temperature (0C)

Fig. 13 e Variation with condenser temperature of the total


exergy efficiency and exergy destruction rate.

1804

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

which lowers the exergy efficiency. Since the exergy


destruction rate in the condenser is much higher than in other
components, any changes in the input exergy flow to the
condenser leads to an increase in its exergy destruction, so an
increase in condenser temperature leads to an increase in
exergy flow at point 2 keeping other parameters constant.

5.

Conclusions

The comprehensive thermodynamic modeling and exergy


analysis of an integrated OTEC system for hydrogen production has provided useful insights. A complete thermodynamic
simulation is performed for an OTEC cycle equipped with a flat
plate solar collector and PEM electrolysis for hydrogen
production. Also, an exergy analysis is conducted and the
results show that the exergy efficiency of the integrated OTEC
system is about 22%. Other significant conclusions follow:
 An increase in solar radiation intensity decreases the total
exergy destruction rate of the system and raises the exergy
efficiency of the cycle and the hydrogen production rate.
 An increase in solar intensity results in a drastic reduction
in condenser exergy destruction rate, which ultimately
leads to a reduction in total exergy destruction rate of the
cycle.
 An increase in ambient temperature has two different
effects on exergy efficiency: an increase in both exergy
efficiency and sustainability index below T0 292 K, and
a reduction of both exergy efficiency and SI above T0 292 K.
 Increasing the condenser temperature reduces the exergy
efficiency of the cycle because this increase raises the
enthalpy of the turbine outlet which reduces the net power
output.

Acknowledgement
The authors acknowledge the support provided by the Natural
Sciences and Engineering Research Council of Canada.

Nomenclature

A
CP
d
Deq
DP
Eact;i
$
Ex
$
ExD
$
Exelectric
F
g
G
h
I
J

surface area (m2)


specific heat at constant pressure (kJ/kg K)
diameter (m)
equivalent diameter (m)
depletion number
activation energy in cathode or anode (kJ)
exergy flow rate (kW)
exergy destruction rate (kW)
electric exergy input rate (kW)
Faraday constant (C/mol)
gravitational acceleration (m/s2)
Gibbs free energy (J/mol)
specific enthalpy (kJ/kg)
solar radiation intensity (W/m2)
current density (A/m2)

J0
Jref
i
L
_
m
N_
P
Q_
R
RPEM
s
S
T
V
V0
Vact
Vact;a
Vact;c
_
W
_G
W
_ CS
W
_ WF
W
_ WS
W

exchange current density (A/m2)


pre-exponential factor (A/m2)
length (m)
mass flow rate (kg/s)
molar mass flow rate (mol/s)
pressure (kPa)
heat transfer rate (kW)
gas constant (kJ/kg K)
proton exchange membrane resistance (U)
specific entropy (kJ/kg K)
radiation flux (W/m2)
temperature ( C or K)
velocity (m/s)
reversible potential (V)
activation overpotential (V)
anode activation overpotential (V)
cathode activation overpotential (V)
work rate (kW)
turbine generator power (kW)
cold seawater pumping power (kW)
working fluid power (kW)
warm seawater pumping power (kW)

Greek letters
h
energy efficiency
turbine efficiency
hT
generator efficiency
hG
working fluid pump efficiency
hWFP
cold surface pump efficiency
hCSP
warm surface pump efficiency
hWSP
water content at the anode-membrane interface
la
(U1)
water content at the cathode-membrane interface
lc
(U1)
lx
water content at location x in the membrane (U1)
r
density (kg/m3)
proton conductivity in PEM (s/m)
sPEM
sx
local ionic PEM conductivity (s/m)
J
exergy efficiency
Subscripts
act
activation
G
generator
CS
cold surface
C
condenser
D
destruction
E
evaporator
eq
equivalent
in
inlet condition
net
net power
ohm
ohmic
out
outlet condition
PEM
polymer exchange membrane
T
turbine
p
pump
WF
working fluid
WS
warm surface
Superscripts
$
rate

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 7 9 5 e1 8 0 5

references

[1] Tchanche BF, Lambrinos GR, Frangoudakis A, Papadakis G.


Low-grade heat conversion into power using organic rankine
cycles: a review of various applications. Renewable and
Sustainable Energy Reviews 2011;15:3963e79.
[2] Faizal M, Rafiuddin Ahmed M. On the ocean heat budget and
ocean thermal energy conversion. International Journal of
Energy Research 2011;35:1119e44.
[3] Meegahapola L, Udawatta L, Witharana S. The ocean thermal
energy conversion strategies and analysis of current
challenges. In: Proceeding of second international
conference on industrial and information systems (ICIIS
2007), Sri Lanka; 2007.
[4] Esteban M, Leary D. Current developments and future
prospects of offshore wind and ocean energy. Applied Energy
2012;90:128e36.
[5] Uehara H, Nakaoka T. OTEC using plate-type heat exchanger
(using ammonia as working fluid). Transactions of JSME
1984;50(453):1325e33.
[6] Uehara H, Ikegami Y. Optimization of a closed-cycle OTEC
system. Journal of Solar Energy Engineering 1990;112:247e56.
[7] Uehara H, Miyara A, Ikegami Y, Nakaoka T. Performance
analysis of an OTEC plant and a desalination plant using an
integrated hybrid cycle. Journal of Solar Energy Engineering
1996;118:115e22.
[8] Uehara H, Dilao CO, Nakaoka T. Conceptual design of ocean
thermal energy conversion power plants in the Philippines.
Solar Energy 1998;41(5):431e41.
[9] Yamada N, Hoshi A, Ikegami Y. Performance simulation of
ocean thermal energy conversion plant. Renewable Energy
2009:1752e8.
[10] Saitoh TS, Yamada N. Advanced multiple Rankine cycle
system including Uehara cycle for solar and ocean energy
utilization. In: Proceedings of forum on desalination using
renewable energy, Republic of Palau; 2003. p. 167e75.
[11] Owens WL. Optimization of closed-cycle OTEC plants.
ASMEeJSME Thermal Engineering Joint Conference 1980;2:
227e39.
[12] Uehara H, Ikegami Y, Parametric performance analysis of
OTEC using Kalina cycle. ASME joint solar engineering
conference; 1993. p. 203e7.
[13] Uehara U, Ikegami Y, Nishida T. Performance analysis of
OTEC system using a cycle with absorption and extraction
processes. Transactions of JSME 1998;64(624):384e9.
[14] Rabas TJ, Panchal CB, Stevens HC. Integration and
optimization of the gas removal system for hybrid-cycle
OTEC power plants. Journal of Solar Energy Engineering 1990;
112:19e28.
[15] Kazim A. Hydrogen production through an ocean thermal
energy conversion system operating at an optimum
temperature drop. Applied Thermal Engineering 2005;25:
2236e46.
[16] Ikegamia Y, Sasakib H, Goudab T, Uehara H. Experimental
study on a spray flash desalination. Desalination 2006;194:
81e9.
[17] Rosen MA, Dincer I. On exergy and environmental impact.
International Journal of Energy Research 1997;21:643e54.

1805

[18] Kanoglu M, Dincer I, Rosen MA. Understanding energy and


exergy efficiencies for improved energy management in
power plants. Energy Policy 2007;35:3967e78.
[19] Dincer I, Rosen MA. Exergy, energy, environment and
sustainable development. UK: Elsevier; 2007.
[20] Sun F, Ikegami Y, Jia B, Arima H. Optimization design and
exergy analysis of organic rankine cycle in ocean thermal
energy conversion. Applied Ocean Research 2012;35:38e46.
[21] Farahat S, Sarhaddi F, Ajam H. Exergetic optimization of flat
plate solar collectors. Renewable Energy 2009;34:1169e74.
[22] Sukhatme SP. Solar Energy. New York: McGraw-Hill; 1993.
[23] Nihous GC, Vega LA. Design of a 100 MW OTEC-hydrogen
plant ship. Marine Structure 1993;6:207e21.
[24] Cengel YA, Boles MA. In: Thermodynamics: an engineering
approach. 7th ed. McGraw-Hill; 2011.
[25] Esmaili P, Dincer I, Naterer GF. Energy and exergy analyses of
electrolytic hydrogen production with molybdenum-oxo
catalysts. International Journal of Hydrogen Energy 2012;
37(9):7365e72.
[26] Ni M, Leung MK, Leung DYC. Energy and exergy analysis of
hydrogen production by a proton exchange membrane (PEM)
electrolyzer plant. Energy Conversion and Management
2008;49:2748e56.
[27] Nieminem J, Dincer I, Naterer G. Comparative performance
analysis of PEM and solid oxide steam electrolysers.
International Journal of Hydrogen Energy
2010:10,842e10,850.
[28] Ahmadi P, Dincer I. Thermodynamic and
exergoenvironmental analyses, and multi-objective
optimization of a gas turbine power plant. Applied Thermal
Engineering 2011;31:2529e40.
[29] Ahmadi P, Dincer I, Rosen MA. Exergo-environmental analysis
of an integrated organic Rankine cycle for trigeneration.
Energy Conversion and Management 2012;64:447e53.
[30] Dincer I, Hussain MM, Al-Zaharnah I. Energy and exergy use
in the industrial sector of Saudi Arabia. Proceedings of the
Institution of Mechanical Engineers, Part A: Journal of Power
and Energy 2003;217(5):481e92.
[31] Bejan A, Tsatsaronis G, Moran M. Thermal design and
optimization. New York: Wiley; 1996.
[32] Ahmadi P, Rosen MA, Dincer I. Multi-objective exergy-based
optimization of a polygeneration energy system using an
evolutionary algorithm. Energy 2012;46:21e31.
[33] Ahmadi P, Rosen MA, Dincer I. Greenhouse gas emission and
exergo-environmental analyses of a trigeneration energy
system. International Journal of Greenhouse Gas Control
2011;5:1540e9.
[34] Nag PK, Gupta A. Exergy analysis of the Kalina cycle. Applied
Thermal Engineering 1998;18(6):427e39.
[35] Wang JF, Dai YP, Gao L. Exergy analyses and parametric
optimizations for different cogeneration power plants in
cement industry. Applied Thermal Engineering 2009;86(6):
941e8.
[36] Ioroi T, Yasuda K, Siroma Z, Fujiwara N, Miyazaki Y. Thin
film electrocatalyst layer for utilized regenerative polymer
electrolyte fuel cells. Journal of Power Sources 2002;112(2):
583e7.
[37] Millet P, Andolfatto F, Durand R. Design and performance of
a solid polymer electrolyte water electrolyzer. International
Journal of Hydrogen Energy 1996;21(2):87e93.

Vous aimerez peut-être aussi