Next Article in Journal
Fuzzy Fault Detection Observer Design for Unmanned Marine Vehicles Based on Membership-Function-Dependent H/H_ Performance
Next Article in Special Issue
Methodological Solutions for Predicting Energy Efficiency of Organic Rankine Cycle Waste Heat Recovery Systems Considering Technological Constraints
Previous Article in Journal
Analytical Framework for Tension Characterization in Submerged Anchor Cables via Nonlinear In-Plane Free Vibrations
Previous Article in Special Issue
Method of Plotting Fundamental Diagrams of Waterway Traffic Flow—Shipping-Lane Subdivision
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermo-Economic Comparison between Three Different Electrolysis Technologies Powered by a Conventional Organic Rankine Cycle for the Green Hydrogen Production Onboard Liquefied Natural Gas Carriers

1
Energy Engineering Research Group, University Institute of Maritime Studies (ETSNM), Centre for Research in Naval and Industrial Technologies (CITENI), Ferrol Industrial Campus, University of A Coruna, Paseo de Ronda 51, 15011 A Coruna, Spain
2
Energy Engineering Research Group, University Institute of Maritime Studies (ETSNM), Nautical Sciences and Marine Engineering Department, University of A Coruna, Paseo de Ronda 51, 15011 A Coruna, Spain
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(8), 1287; https://doi.org/10.3390/jmse12081287
Submission received: 30 June 2024 / Revised: 22 July 2024 / Accepted: 26 July 2024 / Published: 31 July 2024

Abstract

:
The high demand for natural gas (NG) worldwide has led to an increase in the size of the LNG carrier fleet. However, the heat losses from this type of ship’s engines are not properly managed, nor is the excess boil-off gas (BOG) effectively utilised when generation exceeds the ship’s power demand, resulting in significant energy losses dissipated into the environment. This article suggests storing the lost energy into green H2 for subsequent use. This work compares three different electrolysis technologies: solid oxide (SOEC), proton exchange membrane (PEME), and alkaline (AE). The energy required by the electrolysis processes is supplied by both the LNG’s excess BOG and engine waste heat through an organic Rankine cycle (ORC). The results show that the SOEC consumes (743.53 kW) less energy while producing more gH2 (21.94 kg/h) compared to PEME (796.25 kW, 13.96 kg/h) and AE (797.69 kW, 10.74 kg/h). In addition, both the overall system and SOEC stack efficiencies are greater than those of PEME and AE, respectively. Although the investment cost required for AE (with and without H2 compression consideration) is cheaper than SOEC and PEME in both scenarios, the cost of the H2 produced by the SOEC is cheaper by more than 2 USD/kgH2 compared to both other technologies.
Keywords:
SOEC; PEME; AE; gH2; ORC; ICE

1. Introduction

Liquefied natural gas (LNG) is one of the preferred options for ships in replacing conventional marine fuels, such as marine diesel oil (MDO) and heavy fuel oil (HFO), thanks to its environmentally friendly aspects [1,2]. Moreover, LNG carriers are a significant contributor to the maritime industry, transporting natural gas (NG) over the globe [3,4,5,6,7]. The propulsion systems of LNG carriers are classified according boil-off gas (BOG), which is mainly used for mechanical or electric propulsion [8,9], with two-stroke engines being the most popular thanks to their high efficiency compared to steam turbines and four stroke engines [10]. In order to comply with the International Maritime Organization (IMO)’s regulations and restrictions concerning the decarbonisation of the maritime sector, ship owners are required to adopt different measures related to the fuel used as well as the propulsion system, reducing the ships’ emissions while improving their efficiency and flexibility [5,8,11]. The BOG generated is generally consumed by the ship’s engine; however, in cases of low power demand and in order to maintain the stable pressure of the cargo tanks, excess BOG is either sent to a reliquefication plant or burned wastefully in a gas combustion unit (GCU) [1,3,8,12]. Although reliquefication plants are still adopted onboard LNG carriers, they present disadvantages, such as their expensive cost, onboard space requirements, and the intensive energy demand [1,8,13,14].
Fernández et al. [5] suggest employing excess BOG to produce hydrogen (H2) fuel through a steam methane reforming plant (SMR), improving the LNG carrier’s efficiency, abandoning the GCU, and reducing emissions. SMR produces around three-quarters of the available H2 thanks to both technological maturity and H2 production costs. However, the drawbacks of this method cannot be overlooked as it relies on fossil fuel (NG), resulting in emissions, as well as onboard space limitations [1,5,15,16]. An eco-friendly alternative is to produce H2 via electrolysis processes as oxygen (O2) is the only by product of the water splitting reaction. Sebbahi et al. [17] presented a comparative analysis of alkaline electrolysis (AE), proton exchange membrane electrolysis (PEME), and solid oxide electrolysis (SOEC) for H2 production using renewable sources. They concluded that although AE is the most mature electrolysis technology, it is still less efficient than PEME and SOEC, respectively. Nejadian et al. [18] conducted a comparative analysis together with techno-economic optimisation of AE, PEME, and SOEC integrated in a multi-generation energy system for power generation, water desalination, and H2 production. They concluded that, thanks to the cooperation of both thermal and electric energy, SOEC shows a better performance in terms of H2 production, as well as energy and exergy efficiencies, followed by PEME and AE, in this order. According to the results of the exergy–economic optimisation using Pareto frontier, the authors concluded that SOEC registers the highest system cost rate, while PEM registers the highest H2 cost rate. Zaccara et al. [19] compared the H2 production achieved through different renewable energy processes: PEME, SOEC, and biomass gasification, coupled with methanol and methane synthesis included in the steel industry. The results demonstrated that the H2 produced through biomass is less pure than that produced through both PEME and SOEC, in addition to carbon dioxide (CO2) production. SOEC consumes less water and energy compared to PEME and is more attractive if high temperatures and waste heat sources are available. Nasser and Hassan [20] compared two different systems that included SOEC and PEME powered by waste heat obtained via steam and organic Rankine cycles. They concluded that the steam Rankine cycle (SRC) shows a better performance than the organic Rankine cycle (ORC) and the SOEC is more efficient, operates more effectively, and has a lower H2 production cost than PEME. Ferrero et al. [21] conducted a comparative analysis between high-temperature (SOEC) and low-temperature (PEME) electrolysis for H2 production. They concluded that, at the same H2 production rate and pressure, SOEC is more efficient and shows a better performance than PEME. Dere et al. [22] studied H2 production onboard ships through PEME powered by the waste heat of the engine’s excess exhaust gases. The results of this study show that the fuel consumption is reduced by 0.5% and achieved a USD 42,740 annual saving thanks to waste heat recovery. Wang et al. [23] conducted a comprehensive thermodynamic analysis of an SOEC powered by a marine engine’s waste heat through an ORC for H2 production and power load adjustment. The results show that the proposed system recovers 44.13% of the engine’s waste heat, producing 0.431 kg/s of H2.
According to the aforementioned literature review, there exists no article that thermo-economically compares the three electrolysis technologies for maritime transport. Hence, the novelty of this article lies in the comparison of the H2 production onboard LNG carriers through different electrolysis processes coupled separately with an ORC system recovering the waste heat (WH) from both the ship’s engine and the BOG excess. The H2 produced is compressed for use when required.
The article is divided as follows: first, Section 2 compares the main characteristics of the three studied electrolysis processes, then Section 3 is dedicated to the overall systems description (ORC, electrolysis processes, H2 compression plant). The electrochemical modelling of the different electrolysers’ stacks is presented in Section 4, while Section 5 outlines the economic analysis of the overall systems. Finally, the simulations results are presented in Section 6 for a better understanding and comparison of the overall H2 production systems.

2. Comparison between the Different Electrolysis Technologies

Table 1 compares the main characteristics of the low- and high-temperature electrolysis technologies studied in this article, namely AE, PEME, and SOEC electrolysis. AE is a cheap technology operating at low pressures and temperatures; however, it is hindered using corrosive liquid electrolytes, has low current densities, and has a low H2 purity requiring an additional H2 purification stage [24,25,26]. Unlike AE, PEME is characterised by high current densities and a non-corrosive solid electrolyte. The main drawback of this technology is the need for ultra-pure feed water and expensive equipment such as the membrane and noble metal electrodes [24,25,26]. SOEC is a high-temperature electrolysis that is still under investigation; high efficiency and low energy consumption play in favour of this technology. The high temperature and thermal cycles of SOEC are limiting, respectively, its application and the materials’ lifetime [24,25,26].

3. Systems Description

This section describes the possible H2 production chain (energy recovery, H2 production and its compression) onboard LNG carriers. As depicted in Figure 1, an organic Rankine cycle is used to recover all the available onboard waste heat and deliver it to the electrolysis stack (AE, PEME, or SOEC) for the H2 production. The H2 produced is then compressed to be used as a clean fuel when needed.

3.1. Ship Model and Characteristics

The study adopts an LNG ship model propelled mechanically with two principal two-stroke engines and four auxiliary four-stroke dual fuel engines. The engines’ data are extracted from MAN-CEAS [35] and the project guide [36], respectively. The engine wastes heat through different stream sources, mainly from the jacket water (JW), scavenge air (SA), and exhaust gases (EG). For simplification and consistency in calculation, the data collected from the engines are correlated with the Engineering Equation Solver (EES), generating equations for the WH streams and the specific energy consumption as a function of the ship’s load (set to a value of 70% for the principal engines and 80% for the auxiliary engines). The main ship characteristics are summarized in Table 2.
The heat needed ( Q F W G ) to generate a given freshwater flow rate is calculated by Equation (1) with 15% of tolerance ( t F W G ) [37]:
Q F W G = m ˙ F W G 0.03 ( 1 t F W G )
The mass flow rate ( m ˙ B O G ) of the BOG extracted from the LNG tanks is calculated by the following equations:
m ˙ B O G = m ˙ B O G N ρ B O G ρ L N G ( m ˙ L N G + m ˙ B O G N )
m ˙ B O G N = B O R · V t a n k · ρ L N G
where m ˙ L N G and m ˙ B O G N are the mass flow rate of the LNG extracted from the tanks and the natural BOG, respectively; ρ B O G   and ρ L N G   are the BOG and LNG densities; V t a n k is the total cargo capacity; and B O R is the boil-off rate [38].

3.2. Waste Heat Recovery System (ORC)

Recovering the waste heat onboard ships is advantageous as it improves the ships’ energy efficiency and reduces the fuel consumption, which is turn results in lower emissions and reduced operating costs [39,40,41]. Among the different waste heat recovery (WHR) cycles, organic Rankine cycles (ORCs) are commonly studied in the literature and widely used for the WHR onboard marine vessels [39,40]. ORCs are an attractive system to recover the waste heat and convert it to a useful power [42,43,44]. They are characterised by their simplicity and the use of affordable and readily available components (similar to those in a refrigeration system), as well as their ability and flexibility to recover heat from both low- and medium-temperature sources, such as the scavenge air and jacket water [39,45,46]. Choosing the right ORC working fluid is tricky, as it involves considering several factors, including environmental impact, operating conditions (e.g., pressures and temperatures), and economic considerations [39]. This study adopts R245fa as the ORC’s working fluid (WF). Abdul Qyyum et al. [47] assessed various WFs and concluded that, among the 95 studied, R245fa is the most commonly used and optimal WF for ORC systems. In addition, despite the high global warming potential (GWP) of R245fa that reaches a value of 1030, it is a suitable WF for the WHR from marine engines [42,48,49], and it has many advantages such as the low pump power consumption, availability, nonflammability, high net power output, fire hazard reduction, and payback minimisation [42,47,50,51]. Table 3 summarised the key parameters of R245fa.
Figure 2 illustrates the ORC configuration adopted in this article. The organic fluid R245fa is pressurised (2) by the pump (P1), then preheated and vaporised through the different WHR streams (JW, SA, and EG) of the multi-streams heat exchanger (MSHEX). The resulting saturated vapor (3) is expanded in the turbine (T), producing mechanical power that is converted to electric energy through a generator (G). Hence, supplying the ship’s utility services with their power needs and providing the remaining energy to the electrolysis stacks. The remaining fluid (4) leaving the turbine is condensed back to a liquid state (1) and returned to the pump for a new cycle. The economiser (ECO) is used to preheat the steam required by the ship services, as well as for an additional water preheating process in the case of SOEC. Freshwater is secured onboard the ship through a freshwater generator (FWG) by using the available heat supplied by the JW.
The ORC energy efficiency ( ƞ O R C ) as well as the available ( ƞ a v ,   L H V ) and recovered ( ƞ r e c ,   L H V ) system efficiencies are determined as follows:
ƞ O R C = W T u r b W P u m p Q W H R i n
ƞ a v ,   L H V = m ˙ H 2 p r o d · L H V H 2 Q a v
ƞ r e c ,   L H V = m ˙ H 2 p r o d · L H V H 2 Q r e c
where Q W H R i n is the heat recovered from the engines WH streams, and Q a v   refers to all the available WH, while Q r e c is only the WH recovered and used. W T u r b and W P u m p are the mechanical works of the ORC turbine and pump. m ˙ H 2 p r o d is the mass flow rate of the H2 produced, and L H V H 2 is its lower heating value.

3.3. Hydrogen Production Systems

Although H2 production through electrolysis contributes only 4% of the total worldwide production [53], it is considered the cleanest process since O2 is the only by-product of the H2O splitting reaction. In addition, H2 production onboard ships through electrolysis is more advantageous than steam methane reforming (SMR) due to its ease of use and compactness. This subsection describes the different layouts of electrolysis technologies adopted in this article, namely AE, PEME, and SOEC.

3.3.1. AE Layout

The configuration of the AE adopted in this article is depicted in Figure 3. A molar fraction of 80% H2O is mixed with 20% of an alkaline electrolyte (9), KOH in this study. The H2O needed for the electrolysis process (10) is pumped (11) and mixed (13) with the O2 derived from the anode electrode (12). The O2 produced (14) and the H2 gas leaving the cathode electrode (16) are separated from the electrolyte (22, 23) through separators (S1 and S2), respectively, at a pressure drop of 0.3 bar. The H2 gas produced (18) is separated from the remaining H2O (21) and undergoes further separation through (S4) to ensure its purity (19). The electrolyte residues (24) are pumped to the initial pressure (7 bar) and temperature (75 °C), then returned back to the stack (25).

3.3.2. PEME Layout

The PEME configuration, adopted from [19], is illustrated in Figure 4. H2 and O2 migrate the cathode (16) and anode (10) electrodes, respectively. Part of the H2 produced gas (17) permeates through the membrane to the anode electrode (18), while part of the O2 gas permeates to the cathode electrode (11). The permeation coefficient is calculated according to Equations (7) and (8). On one hand, the permeated O2 is mixed (20) with the H2 (19) derived from the cathode at 15 bar, cooled to 30 °C, then separated (22) from any H2O residues (25). The H2 gas is further purified (23) through S3. On the other hand, the permeated H2 is mixed with O2 (13) at 10% of the cathode pressure (1.5 bar), and the O2 produced (15) is also separated from H2O residues (27). The pressure of the remaining H2O from O2 (27) and H2 (25) separation is reduced to atmospheric pressure, mixed with the recycled H2O (29), and then returned back to the stack (31) at 15 bar and 30 °C.
H 2 p e r m = 0.0009 · e 0.025 · T · Δ P · A c l
O 2 p e r m = H 2 p e r m 2

3.3.3. SOEC Layout

Figure 5 illustrates the SOEC configuration used in this article. A molar fraction of 90% H2O and 10% H2 are mixed (14) in the SOEC stack. H2 and O2 ions migrate to the cathode (15) and anode (10) electrodes, respectively. In order to simplify the calculations, the O2 sweep gas flow (9) is neglected in this study. On one hand, H2 leaves the cathode at 800 °C and then exchanges heat gradually with H2O through a series of heat exchangers (HE1, HE2, and HE3), decreasing its temperature to 30 °C (18). The H2 is separated (19) and purified (20) from any residues. The remaining H2O (23, 24) and H2 (22) are recycled to be used in the H2 cooling process (26, 30), along with the water makeup (25). The H2 produced (21) is compressed for subsequent use as required. On the other hand, O2 leaving the anode at 800 °C is separated (12) from any O2 sweep gas (11), then cooled through (HE4) and released to the environment (13). The slightly heated H2O (27) is reheated (28) by the steam coming from the ORC (33) and used for the O2 cooling process in HE4. The resulting steam (29) is mixed with the recycled H2 (22), reforming the initial composition (90% H2O + 10% H2), while the temperature is raised to 800 °C.

3.4. Hydrogen Compression System

Among the different possible H2 storage methods, compressed H2 is the most commonly used due to its simplicity and technological maturity [54,55]. Figure 6 illustrates the H2 compression stages, the H2 produced (37) through the different studied electrolysis processes (AE, PEME, or SOEC) is separated (38) from any possible remaining H2O (49), then compressed through the compressors (Ci) at a constant pressure ratio, achieving a compression pressure of 150 bar. After each compression, the H2 is cooled down to 30 °C through HE6, HE7, and HE8, then separated from H2O residues. The compressed H2 (47) undergoes another purification process (S9), ensuring the resulting H2 is pure (48).

4. Electrochemical Modelling of the Electrolysis Stacks

Developing an electrochemical model of the electrolysis stacks is crucial for assessing the water splitting process. The stacks are modelled using Aspen Custom Modeler V12.1 (ACM), then exported to Aspen Hysys V12.1 for simulation with the rest of the systems’ components. An in-depth and comprehensive study of the electrochemical equations related to the different water electrolysis (AE, SOEC, and PEME) is already presented in the open literature. The following subsections will briefly state the necessary equations used in the electrochemical modelling according to each electrolysis stack type.

4.1. Standard Electrochemical Equations

Both heat and electricity sources are mandatory inputs to drive the splitting reactions in the different stacks [23,56]. The electric power input ( W e l ) is calculated as follows [18,57]:
W e l = V s t · I s t
V s t = V c l · N c l
I s t = J · A c l
where V s t and V c l are, respectively, the stack and cell voltages, I s t is the stack current, J the current density, A c l the cell area, and N c l   the stack number of cells.
The heat required/generated ( Q s t ) by the electrolysis stack is determined by Equation (12) [24,58], deciding if the electrolysis process operates in an endothermic ( Q s t < 0 ), exothermic ( Q s t > 0 ), or thermoneutral ( Q s t = 0 )   mode:
Q s t = I s t · N c l · V c l V t h
V t h = Δ H 2 F
where V t h is the thermoneutral voltage, Δ H is the total energy consumption, and F the Faraday constant.
The heat losses ( Q l o s s e s ) are calculated in this study as 10% of the electrolysis thermal heat (Equation (14)), while the heat excess ( Q e x c e s s ), if any, is determined by applying the global energy balance (Equation (15)):
Q l o s s e s = 10 % · Q s t
Q e x c e s s = Q s t Q l o s s e s
The energy ( ƞ e n L H V ) and exergy ( η e x ) efficiencies of the electrolysis stacks are assessed by Equations (16) and (17) as follows:
ƞ e n L H V = 100 · m ˙ H 2 p r o d · L H V H 2 W e l + Q i n h e a t
η e x = n ˙ H 2 p r o d · E H 2 E e l + E h e a t
E e l = W e l
E h e a t = Q i n h e a t 1 T 0 T
where Q i n h e a t is the sum of all heat inputs required by the electrolysis stack, L H V H 2 is the H2 lower heating value, m ˙ H 2 p r o d and n ˙ H 2 p r o d are the mass and molar flow rates of the H2 produced, respectively, E H 2 is the H2 standard chemical exergy, E e l   and E h e a t   are the rate of the electric and thermal exergy inputs, T is the stack temperature, and T 0 the reference environment temperature.

4.2. AE Stack Modelling

The electrochemical modelling of the AE stack is based on the equation used by [24,59,60]. In general, the cell voltage required for an electrolysis is the sum of the reversible voltage and the voltages generated by irreversible losses. Sánchez et al. [59] have developed comprehensive equations (Equations (20) and (22)) to calculate the cell ( V c l ) and reversible ( V r e v ) voltages using empirical correlations as functions of the stack temperature ( T ) and pressure ( P ) as follows [60]:
V c l = V r e v + r 1 + d 1 + r 2 T + d 2 P · J + s · l o g t 1 + t 2 T + t 3 T 2 · J + 1
V r e v = V r e v 0 + R T z F l n H 2 O 2 0.5 H 2 O
V r e v = a 1 a 2 T + a 3 T l n ( T ) + a 4 T 2 + a 5 T l n ( P ) + a 6 P a 7 P T a 8 P 2 T + a 9 P 2 T 3 / 2 + a 10 P 2 T 2 a 11 P 2 T 3
where ri, di, ti, ai, and s are constant parameters obtained by [24] through experiment (Table 4), and J   is the current density traversing the AE stack.
Faraday efficiency ( η f a r a d a y ) is calculated to measure the AE process effectiveness [57]. In general, η f a r a d a y is the ratio (Equation (23)) between the actual ( m ˙ H 2 p r o d )   and theoretical ( m ˙ H 2 t h e o r )   H2 production rate based on the consumed intensity. η f a r a d a y is also known as the “current efficiency” due the effect of the parasitic current losses throughout the gas conduits [61], these later are affected by the temperature while the pressure has a slight influence [57]; hence, for a given temperature, η f a r a d a y can be expressed by an empirical equation based on the four related parameters fij presented in Table 5 (Equation (24)) [57,59,61]:
η f a r a d a y = m ˙ H 2 p r o d m ˙ H 2 t h e o r
η f a r a d a y = 100 · J 2 f 11 + f 12 T + J 2 · f 21 + f 22 T
The molar flow rates of the H2O consumed ( n ˙ H 2 O c o n s ) during the AE process as well as the H2 ( n ˙ H 2 p r o d ) and O2 ( n ˙ O 2 p r o d ) produced are calculated by the following equations:
n ˙ H 2 p r o d = η f a r a d a y · J · A c l · N c l 2 F
n ˙ H 2 O c o n s = n ˙ H 2 p r o d
n ˙ O 2 p r o d = n ˙ H 2 p r o d 2
According to [24,57], during the AE process, part of the H2 flow ( n ˙ H T O ) is diffused into the O2 channel (from cathode to anode) through AE diaphragms. The amount of n ˙ H T O is expressed by Equation (28) using the hydrogen to oxygen (HTO) diffusion coefficient Equation (29). The counter-diffusion (oxygen to hydrogen (OTH)) is neglected due to the small amount (0.1 to 0.5%) of the O2 is diffused to the H2 channel [57]:
n ˙ H T O = H T O · n ˙ O 2 p r o d 1 H T O
H T O = C 1 + C 2 T + C 3 T 2 + C 4 + C 5 T + C 6 T 2   e x p C 7 + C 8 T + C 9 T 2 J + E 1 + E 2 P + E 3 P 2 + E 4 + E 5 P + E 6 P 2   e x p E 7 + E 8 P + E 9 P 2 J
where Ci and Ei are the gas purity parameters related to the temperature (T) and pressure (P), respectively, their values are summarised in Table 5.
The AE electrodes are submerged in a KOH electrolyte. Hence, the molar flow rates of H2, O2, H2O, and KOH at the anode and cathode are calculated as follows:
Anode:
n ˙ H 2 a n = n ˙ H T O
n ˙ O 2 a n = n ˙ O 2 p r o d
n ˙ H 2 O a n = n ˙ H 2 O i n n ˙ H 2 O c o n s 2
n ˙ K O H a n = n ˙ K O H i n 2
Cathode:
n ˙ H 2 c a t = n ˙ H 2 p r o d
n ˙ O 2 c a t = 0
n ˙ H 2 O c a t = n ˙ H 2 O a n
n ˙ K O H c a t = n ˙ K O H a n
n ˙ K O H i n and n ˙ H 2 O i n are known molar inlet flows of KOH and H2O.

AE Stack Validation

In order to validate the AE stack modelling, the cell voltage–current density curve is compared with the curve obtained by Sánchez et al. [57] and the Aspentech modified modelling version [60] under the conditions presented in Table 6. According to Figure 7, the curve displays a significant concordance with the existing findings.

4.3. PEME Stack Modelling

The PEME stack is modelled according to the equations used in research studies such as [62,63,64]. The cell voltage ( V c l ) is the sum of the reversible voltage ( V r e v ) calculated by the Nernst equation, activation overpotentials (Vact) at both anode and cathode electrodes, and the ohmic overpotential of the electrolyte ( V o h m ). The concentration overpotentials (Vconc) are negligible for high current densities not exceeding 10,000 A/m2 [64].
V c l = V r e v + V a c t + V o h m
V r e v = 1.229 8.5 × 10 4 T 298
V a c t = V a c t , a + V a c t , c
V a c t , i = R T F l n J 2 J 0 , i + J 2 J 0 , i 2 + 1 ,   ( i = a , c )
J 0 , i = J e x p , i e x p E a c t , i R T
V o h m = J R e l
R e l = 0 L d x σ λ x
σ λ x = 0.5139 λ x 0.326 · e x p 1268 1 303 1 T
λ x = λ a λ c L x + λ c
where T(K) is the stack temperature, J0,i is the exchange current density of anode (a) and cathode (c), Jexp,i is the pre-exponential factor, Eact,i is the activation energy, Rel is the overall ohmic resistance, L is the membrane thickness, σ[λ(x)] refers to the local ionic conductivity of the PEME, λ(x) is the water content at a location x in the membrane, and λa and λc are the water contents at anode and cathode membrane interfaces, while x is the distance calculated from the cathode–membrane interface. The parameters used in this study are summarised in Table 7.
The molar flow rates of the H2O reacted in the electrolysis process ( n ˙ H 2 O r e a c t ), the H2 produced ( n ˙ H 2 p r o d ), and the remaining H2O ( n ˙ H 2 O o u t ) are calculated by the following equations:
n ˙ H 2 O r e a c t = I · N c l 2 F
n ˙ H 2 p r o d = n ˙ H 2 O r e a c t
n ˙ H 2 O o u t = n ˙ H 2 O i n n ˙ H 2 O r e a c t
The molar flow rates of H2, O2 and H2O at cathode and anode are calculated as stated below:
Anode:
n ˙ H 2 a n = 0
n ˙ O 2 a n = n ˙ H 2 p r o d 2
n ˙ H 2 O a n = n ˙ H 2 O i n n ˙ H 2 O r e a c t
Cathode:
n ˙ H 2 c a t = n ˙ H 2 p r o d
n ˙ O 2 c a t = 0
n ˙ H 2 O a n = 0

PEME Stack Validation

The cell voltage–current density curve of the PEME stack modelling is validated with the finding reported by Mohtaram et al. [65] using the data of Zaccara et al. [19] as outlined in Table 8. The curve results shown in Figure 8 are in good agreement with the findings of [65].

4.4. SOEC Stack Modelling

The required equations for the SOEC electrochemical modelling are extracted from research studies such as [18,23,56,66,67,68,69,70,71,72]. The cell voltage (Vcl) of the SOEC stack is calculated as the sum of the reversible voltage (Vrev), ohmic overpotential (Vohm), activation overpotentials (Vact), and concentration overpotentials (Vconc) presented below:
V c l = V r e v + V o h m + V a c t + V c o n c
V r e v = V 0 + R T 2 F · l n P H 2 0 · P O 2 0 1 / 2 P H 2 O 0
V 0 = 1.253 2.4516 · 10 4 T
P H 2 0 = y H 2 · P
P O 2 0 = y O 2 · P
P H 2 O 0 = y H 2 O · P
V o h m = 2.99 · 10 5 · e x p 10300 T · J · L
where P H 2 O 0 , P H 2 0 , and P O 2 0 are the partial pressures of H2O, H2, and O2, respectively, while y H 2 O , y H 2 , and y O 2 are the corresponding input molar fractions, V0 the standard potential, L the electrolyte layer thickness (Table 9), J the current density, and T(K) the stack temperature.
Vact is calculated in the same way as in the case of PEME with the parameters summarised in Table 9. (Check Equations (40)–(42))
V c o n c = V c o n c , a + V c o n c , c
V c o n c , a = R T 4 F · l n P O 2 0 2 + R · T · J · µ · d a 2 F · B g P O 2 0
V c o n c , c = R T 2 F · l n 1 + J · R · T · d c 2 F · D H 2 O e f f · P H 2 0 1 J · R · T · d c 2 F · D H 2 O e f f · P H 2 O 0
where da and dc are the anode and cathode thickness, respectively, µ is the dynamic viscosity (Equation (66)), Bg is the flow permeability (Equation (67)), and D H 2 O e f f is the effective diffusion coefficient (Equation (68)). Note that P H 2 O 0 , P H 2 0 , and P O 2 0 are in Pascal.
μ = 1.692 + 889.75 T 1000 892.79 T 1000 2 + 905.98 T 1000 3 598.36 T 1000 4 + 221.64 T 1000 5 34.75 T 1000 6
B g = ε 3 72 ξ 1 ε 2 2 r a d 2
1 D H 2 O e f f = ξ ε · 1 D H 2 H 2 O + 1 D H 2 O K
where rad is the average pore radius, while ε and ξ are the electrode porosity and tortuosity, respectively, and their values are presented in Table 9. The Knudsen ( D H 2 O K ) and the molecular ( D H 2 H 2 O ) diffusions are calculated as follows:
D H 2 O K = 2 3 r a d 8 R · T π · M H 2 O
D H 2 H 2 O = 0.00133 1 M H 2 + 1 M H 2 O 1 2 T 3 2 P · ( σ H 2 H 2 O ) 2 · Ω D
Ω D = 1.06036 T 0.1561 + 0.193 e x p ( 0.47635 T ) + 1.03587 e x p ( 1.52996 T ) + 1.76474 e x p ( 3.89411 T )
T = T ε H 2 H 2 O k
ε H 2 H 2 O k = ε H 2 k ε H 2 O k
σ H 2 H 2 O = σ H 2 + σ H 2 O 2
Mj is the molecular mass of species j (j = H2, O2, H2O), ΩD is the dimensionless diffusion collision integral, and σH2O and σH2 the collision diameters of steam and H2, while εH2O/k and εH2/k are the Lennard-Jones potentials. T* is the dimensionless temperature.
The molar flow rates of the H2O reacted during the electrolysis process and the H2 and O2 produced are calculated by the following equations:
n ˙ H 2 O r e a c t = I · N c l 2 F
n ˙ H 2 p r o d = n ˙ H 2 O r e a c t
n ˙ O 2 p r o d = n ˙ H 2 p r o d 2
On the other hand, the molar flow rates of H2, O2, and H2O at the anode and cathode electrodes are determined as follows:
Anode:
n ˙ H 2 a n = 0
n ˙ O 2 a n = n ˙ O 2 p r o d
n ˙ H 2 O a n = 0
Cathode:
n ˙ H 2 c a t = n ˙ H 2 p r o d
n ˙ O 2 c a t = 0
n ˙ H 2 O a n = n ˙ H 2 O i n n ˙ H 2 O r e a c t

SOEC Stack Validation

The cell voltage–current density curve of the SOEC stack modelling is validated with the findings of [66] using the input data of [23] (Table 10). The cell voltage–current density curve depicted in Figure 9 shows a good agreement with the findings of [66].

5. Economic Analysis

The equipment cost of the three H2 production overall systems is determined by the Aspen Process Economic Analyzer V12.1 (APEA). However, the electrolysers’ cost functions are calculated separately by using the following equations [18,74]:
Z S O E C = 2285   ( U S D / k W ) · W e l . S O E C
Z P E M E = 2068   ( U S D / k W ) · W e l . P E M E
Z A E = 1268   ( U S D / k W ) · W e l . A E
Note that the electrolysers’ cost functions are calculated for the year 2020, while the equipment cost is given for the year 2019. For accurate and significant results, the total equipment costs as well as the electrolysers’ cost are updated to the same year (2022 in this study), using the chemical plant cost index (CEPCI) as shown in the following equation:
Z 2022 = C E P C I 2022 C E P C I e q u i p m e n t   y e a r · Z e q u i p m e n t   y e a r
The cost of the H2 produced by each overall system is determined through the calculation of the levelised cost of H2 (LCOH) as follows [66]:
L C O H = Z ˙ c a p i t a l   c o s t + Z ˙ O & M + Z ˙ e l e c t r i c i t y + Z ˙ f u e l m ˙ H 2 p r
Z ˙ c a p i t a l   c o s t = Z t o t a l · C R F T W . h r
C R F = i · ( 1 + i ) N ( 1 + i ) N 1
Z ˙ O & M = α · Z t o t a l
where Z ˙ c a p i t a l   c o s t is the investment cost rate, Z ˙ O & M is the operating and maintenance annual cost, and Z ˙ e l e c t r i c i t y and Z ˙ f u e l are the electricity and fuel costs which are excluded from consideration in this study since all the energy used is that recovered from the WH without any additional energy. Z t o t a l is the total equipment cost, CRF refers to the capital recovery factor, and T W . h r is the uptime per year, while i, N, and α are the interest rate, the plant lifetime, and the operating and maintenance factor, respectively. The parameters adopted in this study are summarised in the following Table 11:

6. Results and Discussion

This section is dedicated to compare the three studied electrolysis in this article (AE, PEME, and SOEC). First, the electrolysis stacks are compared, separately, between each other in Section 6.1, then an overall comparison of the onboard H2 production system is discussed in Section 6.2. Finally, Section 6.3 presents an economic analysis of the overall systems.

6.1. Electrolysis Stacks’ Comparison

The AE, PEME, and SOEC stacks are compared in this subsection using the same possible input data. The results summarised in Table 12 as well as in Figure 10 and Figure 11 demonstrate the following:
  • SOEC (3.761 kg/h) and PEME (3.760 kg/h) stacks produce more H2 than the AE (3.652 kg/h) stack for the same current density J;
  • There is a slight difference between PEME and SOEC H2 production; however, PEME as well as AE consume more than double the energy (Wel) of SOEC;
  • The cell voltage required for the SOEC process is lower than for PEME and AE,
  • SOEC stack is more efficient than PEME and AE, respectively.

6.2. Overall Systems’ Comparison

After comparing the different electrolysis stacks separately, this section compares the overall H2 production systems from energy harvesting to the H2 compression stage. The thermodynamic properties of the different systems are presented in the Supplementary Materials, while Table 13 outlines the input parameters of the WHR sources as a function of the engine’s load which is set at a fixed value of 70%.
The input data together with the simulation results of the different H2 production chain systems are summarised in Table 14. The results demonstrate that SOEC consumes less energy (743.53 kW) than both PEME (797.69 kW) and AE (796.25 kW) while producing more H2 (21.94 kg/h) compared to 13.96 kg/h by PEME and 10.74 kg/h by AE. The difference in the power output between the ORC coupled with SOEC and the ORC coupled with PEME or AE is attributed to the slightly elevated temperature of the EG. This increase in temperature results from the additional heating process of the water needed by SOEC to reach the high operating temperature (800 °C).
Considering only the recovered WH, the overall system efficiency employing SOEC electrolysis is 10.59% with a stack efficiency of 64.34%LHV and an ORC efficiency of 12.80%, exceeding those of PEME and AE. The PEME and AE overall system efficiencies are 6.76% and 5.20%, respectively, while the stacks’ efficiencies register 58.33%LHV and 44.95%LHV. The ORC efficiency is 12.62% for both PEME and AE. Whether considering all the available onboard WH or solely the recovered one, the SOEC system outperforms both PEME and AE in terms of energy efficiencies, particularly the overall system efficiency, which is almost double, as well as H2 production thanks to the high operating temperatures.

6.3. Economic Analysis of the Overall H2 Production Systems

This subsection economically compares the overall systems using SOEC, PEME, or AE as the H2 production process in order to decide the most viable technology to be adopted onboard maritime vessels. Table 15, Table 16 and Table 17 summarise the equipment’ cost functions of the three studied systems: SOEC, PEME, and AE, respectively, considering two different scenarios: (with (SC-1) or without (SC-2) a H2 compression plant. As mentioned before, the cost function of the electrolysers is calculated separately, the results are shown in Table 18.
Both the total equipment cost and the electrolysers’ costs are updated to the same year (2022 in this study) by using the aforementioned Equation (87) for an accurate and precise comparison. As illustrated in Figure 12, excluding the H2 compression plant (SC-2) reduces the total investment cost by more than half for all of the three overall systems studied. Including the H2 compression system (SC-1) makes SOEC increase the plant investment cost by half a million compared to the PEME system, while excluding the compression plant plays in favour of SOEC by reducing its cost by almost USD 36,000. This is mainly due to the compressors’ costs affected by the H2 produced and the slight difference between the SOEC and PEME electrolyser’s costs. AE is the cheapest system.
According to the results illustrated in Figure 13, the H2 production cost including an AE electrolysis system is higher than PEME and SOEC, in this order, in both scenarios. In addition, in the case of SC-1, the H2 production cost by the SOEC system is cheaper than AE and PEME systems by more than 6 USD/kgH2 and 4 USD/kgH2, respectively. In the case of SC-2, there is a small difference between the LCOH registered by AE and PEME systems (0.24 USD/kgH2), while SOEC is cheaper by more than 2 USD/kgH2 compared to both systems. This is due to the higher H2 production by SOEC.

7. Conclusions

The article investigates the H2 production on board LNG carriers through three different electrolysis processes: AE, PEME, and SOEC. A comparison between the three electrolysers helps in ascertaining the most suitable and efficient technology for the onboard H2 production. The required energy by the electrolysis stacks to drive the water splitting reactions is secured by the total WH energy of the LNG carrier’s propulsion system through an ORC.
The electrolysis stacks are modelled through ACM V12.1 then exported to Aspen Hysys V12.1 for simulation with the rest of the plant component system. Aspen EDR V12.1 is used for the heat exchangers’ design, while APEA is used for the economic analysis. According to the results, the main conclusions are as follows:
  • SOEC consumes less energy (743.53 kW) than PEME (796.25 kW) and AE (797.69 kW), respectively, while producing more H2, 21.94 kg/h compared to 13.96 kg/h by the PEME and 10.74 kg/h by the AE;
  • The SOEC system electrolyser’s efficiency (64.34%LHV), ORC efficiency (12.8%), and overall system efficiency (10.59%) are all higher than those of PEME registering 58.33% for the electrolyser, 12.62% for the ORC, and 6.76% for the overall system; while AE registers 44.95% for the electrolyser, 12.62% for the ORC same as the PEME, and 5.2% for the overall system;
  • Although the total investment cost of the plant including the SOEC system is higher than both PEME and AE, the LCOH of the overall SOEC system is lower by almost double in cost compared to PEME and AE.
To conclude, H2 fuel is a promising alternative to fossil fuels in reducing emissions and aiming towards the decarbonisation of maritime transport. Among the three studied electrolysis systems, SOEC proves to be more advantageous than both PEME and AE for the onboard H2 production as it consumes less energy while producing almost double the mass flow of H2 per hr at a lower cost and being more efficient.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jmse12081287/s1, Thermodynamic properties of the different studied systems.

Author Contributions

Conceptualisation, D.E., M.N., and M.R.G.; methodology, D.E., M.N., and M.R.G.; software, D.E., M.N., and M.R.G.; validation, M.N. and M.R.G.; formal analysis, D.E.; investigation, D.E.; writing—original draft preparation, D.E.; writing—review and editing, M.N. and M.R.G.; visualisation, D.E.; supervision, M.N. and M.R.G.; formal analysis, D.E. All authors have read and agreed to the published version of the manuscript.

Funding

Research contract funded by the Ferrol Industrial Campus under the Call “Talent in Training 2023” through the Research Center for Naval and Industrial Technologies (CITENI). Convenio Xunta de Galicia-Universidad de A Coruña: 2022/CP/139, Spain.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

AclCell areaLNGLNG mass flow rate
ACMAspen custom modelerMSHEXMulti-streams heat exchanger
AEAlkaline electrolysisNPlant lifetime
anAnodeNclNumber of stack cells
APEAAspen process economic analyserNGNatural gas
BgFlow permeabilityNOXNitrogen oxides
BOGBoil-off gasO2Oxygen
BORDesign natural boil-off O2permO2 permeation coefficient
CCompressorOPOzone depletion potential
CatCathodeORCConventional organic Rankine cycle
CEPCIChemical plant cost indexOTHOxygen to hydrogen diffusion
CO2Carbon dioxidePPressure
CONDCondenserPEMEProton exchange membrane electrolysis
CRFCapital recovery factorPjPartial pressure of species j
daAnode thicknessQavAvailable heat
dcCathode thicknessQexcessHeat excess
DFDual fuelQin-heatHeat input
DH2-H2OMolecular diffusionQlossesHeat losses
DH2O-KKnudsen diffusionQrecRecovered heat
Eact,iActivation energy at electrode iQstElectrolyser’s heat
ECOEconomiserRGas constant
EDRAspen exchanger design and ratingradAverage pore radius
EelRate of the electric exergy inputRelOverall ohmic resistance
EESEngineering Equation SolverSSeparator
EGExhaust gasSAScavenge air
EH2Standard chemical exergySMRSteam methane reforming
EheatRate of the thermal exergy inputSOECSolid oxide electrolysis cell
ESSEnergy storage systemTTemperature
FFaraday constantT0Reference environment temperature
Fr.WFreshwaterTW.hrWorking hours
FWGFreshwater generatorV0Standard potential
GCUGas combustion unitVactActivation overpotential
GENGeneratorVclCell voltage
gH2Green hydrogenVconcConcentration overpotential
GWPGlobal warming potentialVohmOhmic overpotential
H2HydrogenVrevReversible overpotential
H2permH2 permeation coefficient VstStack voltage
HEHeat exchangerVtankTotal cargo capacity
HFOHeavy fuel oilVthThermoneutral voltage
HHVHigher heating value of H2WelElectrical energy
HTOHydrogen to oxygen diffusionWHWaste heat
iInterest rateWHRWaste heat recovery
ICEInternal combustion engineZAEAE cost function
IMOInternational maritime organisationZO&MOperating and maintenance cost
IstStack currentZPEMEPEME cost function
JCurrent densityZSOECSOEC cost function
J0,iExchange current density at electrode iZtotalTotal investment cost
Jexp,iPre-exponential factor at electrode iεElectrode porosity
JWJacket waterεH2/kLennard-Jones potential of hydrogen
KOHPotassium hydroxide εH2O/kLennard-Jones potential of steam
LElectrolyte layer thicknessηEfficiency
LCOHLevelised cost of hydrogenλiWater content at anode and cathode
LHVLower heating value o H2μDynamic viscosity
LNGLiquefied natural gasξElectrode tortuosity
BOGBOG mass flow rate ρBOGBOG density
BOGNBOGN mass flow rateρLNGLNG density
MCRMaximum continuous rating power σH2Collision diameter of hydrogen
MDOMarine diesel oilσH2OCollision diameter of steam
MjMolecular mass of species jΩDDimensionless diffusion

References

  1. Kim, K.; Park, K.; Roh, G.; Chun, K. Case Study on Boil-Off Gas (BOG) Minimization for LNG Bunkering Vessel Using Energy Storage System (ESS). J. Mar. Sci. Eng. 2019, 7, 130. [Google Scholar] [CrossRef]
  2. Livaniou, S.; Papadopoulos, G.A. Liquefied Natural Gas (LNG) as a Transitional Choice Replacing Marine Conventional Fuels (Heavy Fuel Oil/Marine Diesel Oil), towards the Era of Decarbonisation. Sustainability 2022, 14, 16364. [Google Scholar] [CrossRef]
  3. Fernández, I.A.; Gomez, M.R.; Gómez, J.R.; López-González, L.M. Generation of H2 on board LNG vessels for consumption in the propulsion system. Pol. Marit. Res. 2020, 27, 83–95. [Google Scholar] [CrossRef]
  4. Aguilera, R.F.; Aguilera, R. World natural gas endowment as a bridge towards zero carbon emissions. Technol. Forecast. Soc. Chang. 2012, 79, 579–586. [Google Scholar] [CrossRef]
  5. Fernández, I.A.; Gómez, M.R.; Gómez, J.R.; López-González, L.M. H2 production by the steam reforming of excess boil off gas on LNG vessels. Energy Convers. Manag. 2017, 134, 301–313. [Google Scholar] [CrossRef]
  6. Maxwell, D.; Zhu, Z. Natural gas prices, LNG transport costs, and the dynamics of LNG imports. Energy Econ. 2011, 33, 217–226. [Google Scholar] [CrossRef]
  7. Kumar, S.; Kwon, H.-T.; Choi, K.-H.; Cho, J.H.; Lim, W.; Moon, I. Current status and future projections of LNG demand and supplies: A global prospective. Energy Policy 2011, 39, 4097–4104. [Google Scholar] [CrossRef]
  8. Fernández, I.A.; Gómez, M.R.; Gómez, J.R.; Insua, Á.B. Review of propulsion systems on LNG carriers. Renew. Sustain. Energy Rev. 2017, 67, 1395–1411. [Google Scholar] [CrossRef]
  9. Gómez, J.R.; Gómez, M.R.; Garcia, R.F.; Catoira, A.D.M. On board LNG reliquefaction technology: A comparative study. Pol. Marit. Res. 2014, 21, 77–88. [Google Scholar] [CrossRef]
  10. Tu, H.; Fan, H.; Lei, W.; Zhou, G. Options and Evaluations on Propulsion Systems of LNG Carriers. In Propulsion Systems; Alessandro, S., Mario, P., Eds.; IntechOpen: Rijeka, Croatia, 2019; p. Ch. 1. [Google Scholar] [CrossRef]
  11. Burel, F.; Taccani, R.; Zuliani, N. Improving sustainability of maritime transport through utilization of Liquefied Natural Gas (LNG) for propulsion. Energy 2013, 57, 412–420. [Google Scholar] [CrossRef]
  12. Yeo, D.; Ahn, B.; Kim, J.; Kim, I. Propulsion alternatives for modern LNG carriers. In Proceedings of the Gas Technology Institute-15th International Conference and Exhibition on Liquefied Natural Gas, Barcelona, Spain, 24–27 April 2007. [Google Scholar]
  13. Chang, D.; Rhee, T.; Nam, K.; Lee, S.; Kwak, B.; Ha, J. Economic evaluation of propulsion systems for LNG carriers: A comparative life cycle cost approach. Hydrocarb. Asia 2008, 18, 22–24,26,28. [Google Scholar]
  14. Gómez, J.R.; Gómez, M.R.; Bernal, J.L.; Insua, A.B. Analysis and efficiency enhancement of a boil-off gas reliquefaction system with cascade cycle on board LNG carriers. Energy Convers. Manag. 2015, 94, 261–274. [Google Scholar] [CrossRef]
  15. Dincer, I. Green methods for hydrogen production. Int. J. Hydrogen Energy 2012, 37, 1954–1971. [Google Scholar] [CrossRef]
  16. Belz, S. A synergetic use of hydrogen and fuel cells in human spaceflight power systems. Acta Astronaut. 2016, 121, 323–331. [Google Scholar] [CrossRef]
  17. Sebbahi, S.; Nabil, N.; Alaoui-Belghiti, A.; Laasri, S.; Rachidi, S.; Hajjaji, A. Assessment of the three most developed water electrolysis technologies: Alkaline Water Electrolysis, Proton Exchange Membrane and Solid-Oxide Electrolysis. Mater. Today Proc. 2022, 66, 140–145. [Google Scholar] [CrossRef]
  18. Nejadian, M.M.; Ahmadi, P.; Houshfar, E. Comparative optimization study of three novel integrated hydrogen production systems with SOEC, PEM, and alkaline electrolyzer. Fuel 2023, 336, 126835. [Google Scholar] [CrossRef]
  19. Zaccara, A.; Petrucciani, A.; Matino, I.; Branca, T.A.; Dettori, S.; Iannino, V.; Colla, V.; Bampaou, M.; Panopoulos, K. Renewable Hydrogen Production Processes for the Off-Gas Valorization in Integrated Steelworks through Hydrogen Intensified Methane and Methanol Syntheses. Metals 2020, 10, 1535. [Google Scholar] [CrossRef]
  20. Nasser, M.; Hassan, H. Assessment of hydrogen production from waste heat using hybrid systems of Rankine cycle with proton exchange membrane/solid oxide electrolyzer. Int. J. Hydrogen Energy 2023, 48, 7135–7153. [Google Scholar] [CrossRef]
  21. Ferrero, D.; Lanzini, A.; Santarelli, M.; Leone, P. A comparative assessment on hydrogen production from low- and high-temperature electrolysis. Int. J. Hydrogen Energy 2013, 38, 3523–3536. [Google Scholar] [CrossRef]
  22. Dere, C.; Inal, O.B.; Zincir, B. Utilization of waste heat for onboard hydrogen production in ships. Int. J. Hydrogen Energy 2024, 75, 271–283. [Google Scholar] [CrossRef]
  23. Wang, F.; Wang, L.; Zhang, H.; Xia, L.; Miao, H.; Yuan, J. Design and optimization of hydrogen production by solid oxide electrolyzer with marine engine waste heat recovery and ORC cycle. Energy Convers. Manag. 2021, 229, 113775. [Google Scholar] [CrossRef]
  24. Delgado, M.S. Desarrollo y Validación de un Modelo para la Simulación de Sistemas de Electrólisis Alcalina para la Producción de Hidrógeno a Partir de Energías Renovables. Ph.D. Thesis, E.T.S.I. de Minas y Energía (UPM), Madrid, Spain, 2019. [Google Scholar]
  25. Kumar, S.S.; Lim, H. An overview of water electrolysis technologies for green hydrogen production. Energy Rep. 2022, 8, 13793–13813. [Google Scholar] [CrossRef]
  26. Burton, N.A.; Padilla, R.V.; Rose, A.; Habibullah, H. Increasing the efficiency of hydrogen production from solar powered water electrolysis. Renew. Sustain. Energy Rev. 2021, 135, 110255. [Google Scholar] [CrossRef]
  27. Al-Douri, A.; Groth, K.M. Hydrogen production via electrolysis: State-of-the-art and research needs in risk and reliability analysis. Int. J. Hydrogen Energy 2024, 63, 775–785. [Google Scholar] [CrossRef]
  28. Zhang, S.; Zhang, N. Review on integrated green hydrogen polygeneration system——Electrolysers, modelling, 4 E analysis and optimization. J. Clean. Prod. 2023, 414, 137631. [Google Scholar] [CrossRef]
  29. IRENA. Green Hydrogen Cost Reduction: Scaling Up Electrolysers to Meet the 1.5 °C Climate Goal; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2020; Available online: https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2020/Dec/IRENA_Green_hydrogen_cost_2020.pdf#page=12.00 (accessed on 4 April 2024).
  30. Schmidt, O.; Gambhir, A.; Staffell, I.; Hawkes, A.; Nelson, J.; Few, S. Future cost and performance of water electrolysis: An expert elicitation study. Int. J. Hydrogen Energy 2017, 42, 30470–30492. [Google Scholar] [CrossRef]
  31. Li, Y.; Lin, R.; O’Shea, R.; Thaore, V.; Wall, D.; Murphy, J.D. A perspective on three sustainable hydrogen production technologies with a focus on technology readiness level, cost of production and life cycle environmental impacts. Heliyon 2024, 10, e26637. [Google Scholar] [CrossRef] [PubMed]
  32. Vidas, L.; Castro, R. Recent Developments on Hydrogen Production Technologies: State-of-the-Art Review with a Focus on Green-Electrolysis. Appl. Sci. 2021, 11, 11363. [Google Scholar] [CrossRef]
  33. Buttler, A.; Spliethoff, H. Current status of water electrolysis for energy storage, grid balancing and sector coupling via power-to-gas and power-to-liquids: A review. Renew. Sustain. Energy Rev. 2018, 82, 2440–2454. [Google Scholar] [CrossRef]
  34. Blender, A. LNG Ship. Sketchfab, Ed. Available online: https://sketchfab.com/3d-models/lng-ship-fa335b96450d4344863bbf5d912a0288 (accessed on 19 March 2024).
  35. MAN. CEAS Engine Calculations. Available online: https://www.man-es.com/marine/products/planning-tools-and-downloads/ceas-engine-calculations (accessed on 19 March 2024).
  36. HYUNDAI. Project Guide HIMSEN H35DF for Marine. 2022. Available online: https://www.hyundai-engine.com/en/Customer/TechData?page=4&category=all (accessed on 19 March 2024).
  37. MAN. MAN B&W G70ME-C10.5-GI. Project Guide Electronically Controlled Dual Fuel Two-Stroke Engines. 2023. Available online: https://man-es.com/applications/projectguides/2stroke/content/printed/G70ME-C10_5-GI.pdf (accessed on 21 March 2024).
  38. GTT. Available online: https://gtt.fr/technologies/markiii-systems (accessed on 19 March 2024).
  39. Sprouse, C.; Depcik, C. Review of organic Rankine cycles for internal combustion engine exhaust waste heat recovery. Appl. Therm. Eng. 2013, 51, 711–722. [Google Scholar] [CrossRef]
  40. Konur, O.; Yuksel, O.; Korkmaz, S.A.; Colpan, C.O.; Saatcioglu, O.Y.; Muslu, I. Thermal design and analysis of an organic rankine cycle system utilizing the main engine and cargo oil pump turbine based waste heats in a large tanker ship. J. Clean. Prod. 2022, 368, 133230. [Google Scholar] [CrossRef]
  41. Singh, D.V.; Pedersen, E. A review of waste heat recovery technologies for maritime applications. Energy Convers. Manag. 2016, 111, 315–328. [Google Scholar] [CrossRef]
  42. Mondejar, M.E.; Andreasen, J.G.; Pierobon, L.; Larsen, U.; Thern, M.; Haglind, F. A review of the use of organic Rankine cycle power systems for maritime applications. Renew. Sustain. Energy Rev. 2018, 91, 126–151. [Google Scholar] [CrossRef]
  43. Baldasso, E.; Gilormini, T.J.A.; Mondejar, M.E.; Andreasen, J.G.; Larsen, L.K.; Fan, J.; Haglind, F. Organic Rankine cycle-based waste heat recovery system combined with thermal energy storage for emission-free power generation on ships during harbor stays. J. Clean. Prod. 2020, 271, 122394. [Google Scholar] [CrossRef]
  44. Konur, O.; Colpan, C.O.; Saatcioglu, O.Y. A comprehensive review on organic Rankine cycle systems used as waste heat recovery technologies for marine applications. Energy Sources Part A Recovery Util. Environ. Eff. 2022, 44, 4083–4122. [Google Scholar] [CrossRef]
  45. Zhu, S.; Zhang, K.; Deng, K. A review of waste heat recovery from the marine engine with highly efficient bottoming power cycles. Renew. Sustain. Energy Rev. 2020, 120, 109611. [Google Scholar] [CrossRef]
  46. Andreasen, J.G.; Meroni, A.; Haglind, F. A Comparison of Organic and Steam Rankine Cycle Power Systems for Waste Heat Recovery on Large Ships. Energies 2017, 10, 547. [Google Scholar] [CrossRef]
  47. Qyyum, M.A.; Khan, A.; Ali, S.; Khurram, M.S.; Mao, N.; Naquash, A.; Noon, A.A.; He, T.; Lee, M. Assessment of working fluids, thermal resources and cooling utilities for Organic Rankine Cycles: State-of-the-art comparison, challenges, commercial status, and future prospects. Energy Convers. Manag. 2022, 252, 115055. [Google Scholar] [CrossRef]
  48. Díaz-Secades, L.A.; González, R.; Rivera, N.; Quevedo, J.R.; Montañés, E. Parametric study of organic Rankine working fluids via Bayesian optimization of a preference learning ranking for a waste heat recovery system applied to a case study marine engine. Ocean Eng. 2024, 306, 118124. [Google Scholar] [CrossRef]
  49. Cataldo, F.; Mastrullo, R.; Mauro, A.W.; Vanoli, G.P. Fluid selection of Organic Rankine Cycle for low-temperature waste heat recovery based on thermal optimization. Energy 2014, 72, 159–167. [Google Scholar] [CrossRef]
  50. Ata, S.; Kahraman, A.; Şahin, R.; Aksoy, M. Thermo-enviro-economic analysis of different power cycle configurations for green hydrogen production from waste heat. Energy Convers. Manag. 2024, 301, 118072. [Google Scholar] [CrossRef]
  51. Larsen, U.; Pierobon, L.; Haglind, F.; Gabrielii, C. Design and optimisation of organic Rankine cycles for waste heat recovery in marine applications using the principles of natural selection. Energy 2013, 55, 803–812. [Google Scholar] [CrossRef]
  52. ASHRAE. Designation and Safety Classification of Refrigerants. 2019. Available online: https://www.ashrae.org/file%20library/technical%20resources/standards%20and%20guidelines/standards%20addenda/34_2019_f_20191213.pdf (accessed on 19 July 2024).
  53. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar] [CrossRef]
  54. Abdin, Z.; Zafaranloo, A.; Rafiee, A.; Mérida, W.; Lipiński, W.; Khalilpour, K.R. Hydrogen as an energy vector. Renew. Sustain. Energy Rev. 2020, 120, 109620. [Google Scholar] [CrossRef]
  55. Wang, Z.; Wang, Y.; Afshan, S.; Hjalmarsson, J. A review of metallic tanks for H2 storage with a view to application in future green shipping. Int. J. Hydrogen Energy 2021, 46, 6151–6179. [Google Scholar] [CrossRef]
  56. Wang, F.; Wang, L.; Ou, Y.; Lei, X.; Yuan, J.; Liu, X.; Zhu, Y. Thermodynamic analysis of solid oxide electrolyzer integration with engine waste heat recovery for hydrogen production. Case Stud. Therm. Eng. 2021, 27, 101240. [Google Scholar] [CrossRef]
  57. Sánchez, M.; Amores, E.; Abad, D.; Rodríguez, L.; Clemente-Jul, C. Aspen Plus model of an alkaline electrolysis system for hydrogen production. Int. J. Hydrogen Energy 2020, 45, 3916–3929. [Google Scholar] [CrossRef]
  58. Bianchi, F.R.; Bosio, B. Operating Principles, Performance and Technology Readiness Level of Reversible Solid Oxide Cells. In Sustainability 2021, 13, 4777. [Google Scholar] [CrossRef]
  59. Sánchez, M.; Amores, E.; Rodríguez, L.; Clemente-Jul, C. Semi-empirical model and experimental validation for the performance evaluation of a 15 kW alkaline water electrolyzer. Int. J. Hydrogen Energy 2018, 43, 20332–20345. [Google Scholar] [CrossRef]
  60. AspenTech. Integrated Custom Model of Alkaline Electrolysis System for H2 Production; AspenTech: Bedford, MA, USA, 2021. [Google Scholar]
  61. Ulleberg, Ø. Modeling of advanced alkaline electrolyzers: A system simulation approach. Int. J. Hydrogen Energy 2003, 28, 21–33. [Google Scholar] [CrossRef]
  62. Dincer, I.; Rosen, M.A.; Ahmadi, P. Optimization of Energy Systems; John Wiley & Sons Ltd: Chichester, UK, 2017; p. 472. [Google Scholar] [CrossRef]
  63. Nami, H.; Akrami, E. Analysis of a gas turbine based hybrid system by utilizing energy, exergy and exergoeconomic methodologies for steam, power and hydrogen production. Energy Convers. Manag. 2017, 143, 326–337. [Google Scholar] [CrossRef]
  64. Ni, M.; Leung, M.K.H.; Leung, D.Y.C. Energy and exergy analysis of hydrogen production by a proton exchange membrane (PEM) electrolyzer plant. Energy Convers. Manag. 2008, 49, 2748–2756. [Google Scholar] [CrossRef]
  65. Mohtaram, S.; Wu, W.; Aryanfar, Y.; Yang, Q.; Alcaraz, J.L.G. Introducing and assessment of a new wind and solar-based diversified energy production system intergrading single-effect absorption refrigeration, ORC, and SRC cycles. Renew. Energy 2022, 199, 179–191. [Google Scholar] [CrossRef]
  66. Mohammadi, A.; Mehrpooya, M. Techno-economic analysis of hydrogen production by solid oxide electrolyzer coupled with dish collector. Energy Convers. Manag. 2018, 173, 167–178. [Google Scholar] [CrossRef]
  67. Ni, M.; Leung, M.K.H.; Leung, D.Y.C. Parametric study of solid oxide steam electrolyzer for hydrogen production. Int. J. Hydrogen Energy 2007, 32, 2305–2313. [Google Scholar] [CrossRef]
  68. Visitdumrongkul, N.; Tippawan, P.; Authayanun, S.; Assabumrungrat, S.; Arpornwichanop, A. Enhanced performance of solid oxide electrolysis cells by integration with a partial oxidation reactor: Energy and exergy analyses. Energy Convers. Manag. 2016, 129, 189–199. [Google Scholar] [CrossRef]
  69. Im-orb, K.; Visitdumrongkul, N.; Saebea, D.; Patcharavorachot, Y.; Arpornwichanop, A. Flowsheet-based model and exergy analysis of solid oxide electrolysis cells for clean hydrogen production. J. Clean. Prod. 2018, 170, 1–13. [Google Scholar] [CrossRef]
  70. Zhao, Y.; Xue, H.; Jin, X.; Xiong, B.; Liu, R.; Peng, Y.; Jiang, L.; Tian, G. System level heat integration and efficiency analysis of hydrogen production process based on solid oxide electrolysis cells. Int. J. Hydrogen Energy 2021, 46, 38163–38174. [Google Scholar] [CrossRef]
  71. AlZahrani, A.A. Investigation of a Novel High Temperature Solid Oxide Electrolyzer for Solar Hydrogen Production. Ph.D. Thesis, University of Ontario Institute of Technology, Oshawa, ON, Canada, 2017. [Google Scholar]
  72. Luo, Y.; Wu, X.-y.; Shi, Y.; Ghoniem, A.F.; Cai, N. Exergy analysis of an integrated solid oxide electrolysis cell-methanation reactor for renewable energy storage. Appl. Energy 2018, 215, 371–383. [Google Scholar] [CrossRef]
  73. Zhang, S.; Zhang, N.; Smith, R.; Wang, W. A zero carbon route to the supply of high-temperature heat through the integration of solid oxide electrolysis cells and H2–O2 combustion. Renew. Sustain. Energy Rev. 2022, 167, 112816. [Google Scholar] [CrossRef]
  74. Christensen, A. Assessment of Hydrogen Production Costs from Electrolysis: United States and Europe. 2020. Available online: https://theicct.org/wp-content/uploads/2021/06/final_icct2020_assessment_of-_hydrogen_production_costs-v2.pdf (accessed on 7 March 2024).
  75. Naeini, M.; Cotton, J.S.; Adams, T.A., II. An eco-technoeconomic analysis of hydrogen production using solid oxide electrolysis cells that accounts for long-term degradation. Front. Energy Res. 2022, 10, 1015465. [Google Scholar] [CrossRef]
  76. Querol, E.; Gonzalez-Regueral, B.; Perez-Benedito, J.L. Practical Approach to Exergy and Thermoeconomic Analyses of Industrial Processes; Springer: London, UK, 2013. [Google Scholar]
  77. The-University-of-Manchester-Department-of-Chemical-Engineering. Chemical Engineering Plant Cost Index. 2024. Available online: https://www.training.itservices.manchester.ac.uk/public/gced/CEPCI.html?reactors/CEPCI/index.html (accessed on 7 March 2024).
Figure 1. Overall H2 production system. (Ship source: [34]).
Figure 1. Overall H2 production system. (Ship source: [34]).
Jmse 12 01287 g001
Figure 2. WHR system through an ORC.
Figure 2. WHR system through an ORC.
Jmse 12 01287 g002
Figure 3. AE layout for onboard H2 production.
Figure 3. AE layout for onboard H2 production.
Jmse 12 01287 g003
Figure 4. PEME layout for onboard H2 production.
Figure 4. PEME layout for onboard H2 production.
Jmse 12 01287 g004
Figure 5. SOEC layout for onboard H2 production.
Figure 5. SOEC layout for onboard H2 production.
Jmse 12 01287 g005
Figure 6. H2 compression system.
Figure 6. H2 compression system.
Jmse 12 01287 g006
Figure 7. Validation curve of AE cell voltage–current density with [57,60].
Figure 7. Validation curve of AE cell voltage–current density with [57,60].
Jmse 12 01287 g007
Figure 8. Validation curve of PEME cell voltage–current density with [65].
Figure 8. Validation curve of PEME cell voltage–current density with [65].
Jmse 12 01287 g008
Figure 9. Validation curve of SOEC cell voltage–current density with [66].
Figure 9. Validation curve of SOEC cell voltage–current density with [66].
Jmse 12 01287 g009
Figure 10. Cell voltage comparison between the AE, PEME, and SOEC stacks.
Figure 10. Cell voltage comparison between the AE, PEME, and SOEC stacks.
Jmse 12 01287 g010
Figure 11. Energy efficiency (LHV) comparison between AE, PEME, and SOEC.
Figure 11. Energy efficiency (LHV) comparison between AE, PEME, and SOEC.
Jmse 12 01287 g011
Figure 12. Investment cost of the H2 production by the different electrolysis overall systems.
Figure 12. Investment cost of the H2 production by the different electrolysis overall systems.
Jmse 12 01287 g012
Figure 13. H2 cost comparison.
Figure 13. H2 cost comparison.
Jmse 12 01287 g013
Table 1. Comparison between the different electrolysis technologies [24,25,26,27,28,29,30,31].
Table 1. Comparison between the different electrolysis technologies [24,25,26,27,28,29,30,31].
Low-Temperature ElectrolysisHigh-Temperature Electrolysis
AEPEMESOEC
Semi-reactionsAnode 2 O H H 2 O + 1 2 O 2 + 2 e H 2 O 2 H + + 1 2 O 2 + 2 e O 2 1 2 O 2 + 2 e
Cathode 2 H 2 O + 2 e H 2 + 2 O H 4 H + + 4 e 2 H 2 H 2 O 2 H + + O 2
Overall reaction 2 H 2 O 2 H 2 + O 2 2 H 2 O 2 H 2 + O 2 2 H 2 O 2 H 2 + O 2
ElectrolytePotassium hydroxide (KOH): 20–40 wt% [24,30]Solid polymer electrolyte (Perfluoro sulfonic acid (PFSA)), usually Nafion®Yttria stabilised zirconia (YSZ)
Sodium hydroxide (NaOH): 20 wt% [24]
Anode electrode Nickel coated perforated
Ni-Co alloys
Stainless steel
Metal oxides
Platinum carbon
RuO2, IrO2
Graphite-PTFE + Ti/RuO2, IrO2
Perovskites (LSCF, LSM)
Ceramics (Mn, La, Cr)
YSZ
Cathode electrodeNickel coated perforated
Ni-Mo alloys
Ni-Co alloys
Steel + Ni
Iridium oxide
Pt, Pt-Pd
Graphite + Pt/Pt
Ni/YSZ
Zr + Ni/CeOx
Operating temperature (°C)20–90 20–100600–1200
Operating pressure (bar)1–30<70 [25,27,29]1 [25,27,28,29]
70 [28]<20 [26]
<200 [26,30]1–5 [24]
1–350 [24]<25 [30]
Voltage range (V)1.4–3 1.4–2.50.7–1.5
Current density (A/m2)2000–8000 0–20,000 0–20,000
Cell area (m2)1–3 [25]<0.15 [25]0.02 [25]
<4 [26,30]<0.13 [26]<0.06 [26]
<0.3 [30]<0.01 [30]
Stack energy consumption (kWh/Nm3)4.2–5.9 4.2–5.5>3.2
System energy consumption (kWh/Nm3)4.5–6.6 [26,30]4.2–6.6 [26,30]3.7–3.9 [26,30]
5.55 [32]5.4 [32]3.8 [32]
Stack efficiency (%LHV)50–7850–8380–100
System efficiency (%LHV)51–6046–6076–81
Stack capital cost (USD/kW)Minimum 1 MW270400>2000
Minimum 10 MW500–1000700–1400Unknown
Maintenance cost (% of investment cost/year)2–3 [33]3–5 [33]Unknown [33]
Stack lifetime (h)60,000–120,00020,000–100,0008000–20,000
Maturity
(Technology readiness level TRL)
Mature and commercial
TRL 9
Commercial at small and medium scales
TRL 8 expected to reach 9 by 2050
Development phase
TRL 6 expected to reach 9 by 2050
AdvantagesCheap
No need for noble metal electrocatalysts
Long-term stability
Reliable functioning
Space saving configuration
High H2 purity
Fast start-up
Non-corrosive solid electrolyte
High current densities
High production rate
Cheap
High efficiency
No need for noble metal electrocatalysts
High working temperature
Low pressure
Low energy consumption
Non-corrosive solid electrolyte
Low pure water requirement
DisadvantagesLow H2 purity
Low current densities
Slow start-up
Corrosive
High energy consumption
Expensive membrane and electrodes
Acidic environment
High pressure
Ultra-pure feed water
Limited stability
Small cell area
Safety issues
Limited application
Under development
Table 2. Main characteristics of the model LNG carrier.
Table 2. Main characteristics of the model LNG carrier.
Ship’s CharacteristicValue
Total cargo capacity ( V t a n k )1.7 × 105 m3
Propulsion systemMechanical 2-stroke
Principal engine type2 × MAN–5G70 ME-C10.5
Propulsion power (MCR)2 × 12,835 kW (70.8 rpm)
Auxiliary engines type2 × 6H35DF
2 × 7H35DF
Auxiliary power (MCR)2 × 2880 kW
2 × 3360 kW
Auxiliary power demand3448 kW
Total steam consumption1999.2 kg/h
Boil-off rate ( B O R )0.10%
Freshwater generation ( m ˙ F W G )20,000 kg/day
Table 3. Main characteristics of the working fluid R245fa [47,48,50,52].
Table 3. Main characteristics of the working fluid R245fa [47,48,50,52].
Working Fluid CharacteristicValue
ASHRAE codeR245fa
Chemical namePentafluoro-propane
Chemical formulaCF3CH2CHF2
TypeDry
ASHRAE safety groupB1
ODP0
GWP1030
Critical temperature (°C)154
Critical pressure (bar)36.5
Normal boiling point (°C)15.14
Table 4. Experimental parameters used for the calculation of the AE cell voltage [24,57,59].
Table 4. Experimental parameters used for the calculation of the AE cell voltage [24,57,59].
ParameterValueUnit
Cell voltage
r14.45153 × 10−5Ω m2
r26.88874 × 10−9Ω m2/°C
t1−0.01539m2/A
t22.00181m² °C/A
t315.24178m² °C2/A
d1−3.12996 × 10−6Ω m2
d24.47137 × 10−7Ω m2/bar
s0.33824V
Reversible voltage
a11.5184V
a20.0015421V/K
a30.00009523V
a40.0000000984V/K2
a50.000064629V/K bar
a60.000021946V/bar
a70.0055433V·K/bar
a80.0000095196V·K/bar2
a90.00013914V·K3/2/bar2
a100.0026144V·K2/bar2
a110.4953V·K3/bar2
Table 5. Faraday and HTO diffusion experimental parameters used for AE stack modelling [24,57,59].
Table 5. Faraday and HTO diffusion experimental parameters used for AE stack modelling [24,57,59].
ParameterValueUnit
Faraday efficiency
f11478,645.74A2/m4
f12−2953.15A2/m4 °C
f211.03960--
f22−0.00104°C−1
Hydrogen to oxygen diffusion
C10.09901--
C2−0.00207°C−1
C31.31064 × 10−5°C−2
C4−0.08483--
C50.00179°C−1
C6−1.13390 × 10−5°C−2
C71481.45A/m2
C8−23.60345A/m2 °C1
C9−0.25774A/m2 °C2
E13.71417--
E2−0.93063Bar−1
E30.05817Bar−2
E4−3.72068--
E50.93219Bar−1
E6−0.05826Bar−2
E7−18.38215A/m2
E85.87316A/m2 Bar
E9−0.46425A/m2 Bar2
Table 6. Input data for the AE simulation and validation [57,60].
Table 6. Input data for the AE simulation and validation [57,60].
VariableValueUnit
T75°C
P7Bar
Wel10kW
Ncl12--
Acl0.1m2
in900Kg/h
Inlet composition (H2O–KOH)35–65 % Mass fraction basis
Table 7. Parameters adopted for the PEME stack electrochemical modelling [63].
Table 7. Parameters adopted for the PEME stack electrochemical modelling [63].
ParameterValueUnit
J0,a1 × 10−5A/m2
J0,c10A/m2
Eact,a76 × 103J/mol
Eact,c18 × 103J/mol
L50 × 10−6m
λa14--
λc10--
Table 8. Input data for the PEME simulation [19].
Table 8. Input data for the PEME simulation [19].
VariableValueUnit
T90°C
PCathode 15-Anode 1.5 (10%)Bar
n ˙ H 2 O i n 352,080kg/h
Acl1 (*)m2
Ncl12 (*)--
Wel1,988,900kW
(*) Values estimated for modelling purpose.
Table 9. Parameters adopted for the SOEC stack electrochemical modelling [23,73].
Table 9. Parameters adopted for the SOEC stack electrochemical modelling [23,73].
ParameterValueUnit
L12.5 × 10−6m
da17.5 × 10−6m
dc12.5 × 10−6m
ε0.48--
ξ5.4--
rad1.385 × 10−6m
εH2O/k809.1K
εH2/k59.7K
σH2O2.641m
σH22.827m
Jexp,a2.051 × 109A/m2
Jexp,c1.344 × 1010A/m2
Eact,a1.2 × 105J/mol
Eact,c1 × 105J/mol
Table 10. Input data for the SOEC simulation [23].
Table 10. Input data for the SOEC simulation [23].
VariableValueUnit
T850°C
P1.01325bar
in36,000kg/h
J8000A/m2
Acl0.04m2
Ncl130,000--
Inlet composition (H2O–H2)80–20% Molar fraction basis
Table 11. Parameters adopted for the economic analysis [75,76,77].
Table 11. Parameters adopted for the economic analysis [75,76,77].
ParameterValue
T W . h r 8000 h
i10%
N20 years
α3%
CEPCI2019607.5
CEPCI2020596.2
CEPCI2022816
Table 12. Results comparison between the AE, PEME, and SOEC stacks.
Table 12. Results comparison between the AE, PEME, and SOEC stacks.
VariableUnitAEPEMESOEC
T°C8080800
Pbar1.013251.013251.01325
inkg/h36,00036,00036,000
Inlet composition (molar fraction basis)%90% H2O + 10% KOH100% H2O90% H2O + 10% H2
JA/m2500050005000
Aclm20.040.040.04
Ncl--500500500
H2-prodkg/h3.5623.7603.761
O2-prodkg/h28.27329.84429.848
WelkW199.97200.4698.93
VclV2.0002.0050.989
ηEN-LHV%59.3862.5297.44
ηEX%58.5461.6496.76
Table 13. WHR parameters at 70% of the engine’s load.
Table 13. WHR parameters at 70% of the engine’s load.
VariableUnitJWSAEG
Temperature°C90.0153.3196.1
Pressurebar1.013251.013251.01325
Mass flowkg/h144,842140,422142,579
Heat flowkW185647803237
Table 14. Comparison results of the overall H2 production system.
Table 14. Comparison results of the overall H2 production system.
VariableUnitAEPEMESOEC
Temperature°C8080800
Pressurebar7151.01325
Inlet composition (molar fraction basis)%90% H2O + 10% KOH100% H2O90% H2O + 10% H2
A c l m20.040.040.04
N c l --500500500
PORCkW921.4921.4940
Q a v kW987398739873
Q r e c kW688268826910
W e l kW796.25797.69743.53
JA/m214,94218,557.729,174.9
V c l V2.6642.1491.274
inkg/h143.5138.6248.2
n ˙ H 2 p r o d kg/h10.7413.9621.94
n ˙ O 2 p r o d kg/h85.22110.78174.161
ƞ e n L H V %44.9558.3364.34
ηORC-av%8.808.808.96
ηORC-rec%12.6212.6212.80
ƞ a v ,   L H V %3.634.717.41
ƞ r e c ,   L H V %5.206.7610.59
Table 15. Equipment costs of SOEC with and without compression plant (2022).
Table 15. Equipment costs of SOEC with and without compression plant (2022).
EquipmentSC-1SC-2
CompressorsC11,329,000
C21,375,100
C31,434,500
CondenserCOND538,100538,100
SOEC heat exchangersHE151,90051,900
HE278007800
HE315,40015,400
HE418,90018,900
HE511,00011,000
HE617,30017,300
H2 compression coolersHE720,700
HE817,200
HE918,300
Multi-streams heat exchangerMSHEX381,900381,900
ORC pumpP133,80033,800
ORC turbineT487,700487,700
Ztotal (USD)5,758,6001,563,800
Table 16. Equipment costs of PEME with and without a compression plant (2022).
Table 16. Equipment costs of PEME with and without a compression plant (2022).
EquipmentSC-1SC-2
CompressorsC11,232,100
C21,266,300
C31,289,700
CondenserCOND546,900546,900
PEME heat exchangerHE114,30014,300
H2 compression coolersHE714,300
HE815,000
HE916,100
Multi-streams heat exchangerMSHEX548,600548,600
ORC pumpP133,80033,800
PEME pumpP217,00017,000
TurbineT479,900479,900
Ztotal (USD)5,474,0001,640,500
Table 17. Equipment costs of AE with and without a compression plant (2022).
Table 17. Equipment costs of AE with and without a compression plant (2022).
EquipmentSC-1SC-2
CompressorsC11,232,300
C21,258,700
C31,289,700
CondenserCOND546,900546,900
AE heat exchangerHE113,30013,300
H2 compression coolersHE715,000
HE814,700
HE915,200
Multi-streams heat exchangerMSHEX548,600548,600
ORC pumpP133,80033,800
AE pumpP344004400
P449004900
TurbineT479,900479,900
Ztotal (USD)5,457,4001,631,800
Table 18. Electrolysers’ cost (2020).
Table 18. Electrolysers’ cost (2020).
Electrolyser TypePower (kW)Cost (USD)
SOEC743.531,698,975
PEME797.691,649,620
AE796.25100,964
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Elrhoul, D.; Naveiro, M.; Romero Gómez, M. Thermo-Economic Comparison between Three Different Electrolysis Technologies Powered by a Conventional Organic Rankine Cycle for the Green Hydrogen Production Onboard Liquefied Natural Gas Carriers. J. Mar. Sci. Eng. 2024, 12, 1287. https://doi.org/10.3390/jmse12081287

AMA Style

Elrhoul D, Naveiro M, Romero Gómez M. Thermo-Economic Comparison between Three Different Electrolysis Technologies Powered by a Conventional Organic Rankine Cycle for the Green Hydrogen Production Onboard Liquefied Natural Gas Carriers. Journal of Marine Science and Engineering. 2024; 12(8):1287. https://doi.org/10.3390/jmse12081287

Chicago/Turabian Style

Elrhoul, Doha, Manuel Naveiro, and Manuel Romero Gómez. 2024. "Thermo-Economic Comparison between Three Different Electrolysis Technologies Powered by a Conventional Organic Rankine Cycle for the Green Hydrogen Production Onboard Liquefied Natural Gas Carriers" Journal of Marine Science and Engineering 12, no. 8: 1287. https://doi.org/10.3390/jmse12081287

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop