Next Article in Journal
Numerical Investigation of Flow around Two Tandem Cylinders in the Upper Transition Reynolds Number Regime Using Modal Analysis
Next Article in Special Issue
Mixing Properties of Emulsified Fuel Oil from Mixing Marine Bunker-C Fuel Oil and Water
Previous Article in Journal
Post-Cyclic Drained Shear Behaviour of Fujian Sand under Various Loading Conditions
Previous Article in Special Issue
Study on the Cumulative Effects of Using a High-Efficiency Turbocharger and Biodiesel B20 Fuelling on Performance and Emissions of a Large Marine Diesel Engine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Techno-Economic Analysis of NH3 Fuel Supply and Onboard Re-Liquefaction System for an NH3-Fueled Ocean-Going Large Container Ship

1
Department of Naval Architecture and Offshore Engineering, Dong-A University, 37 Nakdong-daero 550 Beon-gil, Saha-gu, Busan 49315, Korea
2
Department of Aerospace Engineering, Republic of Korea Air Force Academy, 635 Danjae-ro, Sangdang-gu, Cheongju 28187, Korea
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2022, 10(10), 1500; https://doi.org/10.3390/jmse10101500
Submission received: 11 September 2022 / Revised: 10 October 2022 / Accepted: 11 October 2022 / Published: 15 October 2022
(This article belongs to the Special Issue Marine Fuels and Green Energy)

Abstract

:
This study proposed the integrated design of an NH3 fuel supply system and a re-liquefaction system for an ocean-going NH3-fueled ship. The target ship was a 14,000 TEU large container ship traveling from Asia to Europe. The NH3 fuel supply system was developed to feed the liquid fuel at 40 °C and 80 bar and cope with the re-circulated fuel with the sealing oil. Its power consumptions and SECs ranged from 56.4 to 157.5 kW and from 0.0063 to 0.009 kWh/kg, respectively. An onboard re-liquefaction system with a vapor compression refrigeration cycle was also designed to liquefy the BOG from the fuel tank. The re-liquefaction system’s exergy efficiency and SEC were 34.71% and 0.224 kWh/kg, respectively. The equipment with the most exergy destruction was the heat exchangers, accounting for 60% of the total exergy destruction. NPV analysis found that it is recommended to introduce the re-liquefaction system to the target ship. At the NH3 price of USD 250/ton, the reasonable cost of the re-liquefaction system is less than USD 1 million. According to LCC, NH3 fuel is economically feasible if the carbon tax is more than USD 80/ton and the NH3 price is around USD 250/ton.

1. Introduction

The International Maritime Organization (IMO) has imposed stringent environmental regulations on the shipping industry to control the Greenhouse Gas (GHG) emissions from international shipping. Since 1 January 2020, The IMO has set the 2020 sulfur cap, reducing global sulfur emissions to 0.5% from the previous level of 3.5% [1,2]. Consequently, the allowable sulfur content in marine fuels has decreased seven times, from 3.5% to 0.5% of the mass. In order to comply with the IMO low-sulfur policy, many shipping companies should adopt very low-sulfur fuel oil, SOx scrubbers, or LNG [3]. Additional regulations to reduce GHG emissions, such as the Energy Efficiency Design Index (EEDI) and Energy Efficiency Operations Index (EEOI), are being tightened. Along with these de-carbonization efforts, the shipping industry must reduce CO2 emissions by approximately 90% between 2010 and 2050 to keep the increase in global temperatures below 1.5 °C. In 2018, the IMO Marine Environment Protection Committee (MEPC) set a target to reduce the shipping sector’s CO2 emissions by 50% by 2050, recognizing the shipping sector’s enormous contribution to global CO2 emissions [4].
In 2018, GHG emissions from ships accounted for approximately 2.89% of global emissions. Various methods have been proposed to reduce CO2 emissions from shipping, including improved hull designs, enhanced power and propulsion systems, increased operational efficiencies, and the use of alternative energies [5]. Furthermore, alternative energy is a viable way to enhance international, national, and regional regulations [6]. H2 and NH3 are the most feasible solutions among various alternative energies. The International Transport Forum (ITF) [7] assumes that, in the case of an 80% carbon factor reduction, hydrogen and NH3 will account for approximately 70% of the fuel market. Moreover, Lewis, J. [8] suggested that H2 and NH3 are the most promising zero-carbon fuel options for de-carbonization in the transportation sector. The International Energy Agency (IEA) [9] estimates that H2 and NH3 have the potential to meet the environmental target in shipping, but their cost of production is high relative to oil-based fuels.
Table 1 shows the characteristics of H2 and NH3 as fuels compared with HFO. Although hydrogen can be obtained from various sources, such as biomass or electrolysis, it is mainly produced from NG [10]. Therefore, its key barriers are the high fuel price and limited availability for maritime operations. In addition, Table 1 shows that H2 liquefaction requires a relatively low temperature of −253 °C, which gives rise to the high costs of liquefaction and building of storage systems onboard. Furthermore, although H2 is an environmentally friendly fuel, it is difficult to store due to its low density. The density of LH2 is approximately 70.8 kg/m3, and that of heavy fuel oil (HFO) is approximately 1010 kg/m3. Therefore, NH3 is currently being discussed as an alternative fuel due to its higher volumetric energy density and ease of handling.
NH3 has a higher volumetric energy density than liquid hydrogen. Although NH3 has a lower gravimetric energy density (18.6 MJ/kg) compared to H2 (120 MJ/kg), the density of liquid NH3 (682 kg/m3) is significantly higher than that of liquid H2 (70.8 kg/m3). Therefore, the volumetric energy density of liquid NH3 (14,100 MJ/m3) is higher than liquid H2 (8500 MJ/m3), which is one of the advantages to fuel storage onboard. The storage requirements of NH3 are similar to those of propane; NH3 is in liquid form at room temperature when pressurized to approximately 10 bar or a temperature of −33.4 °C at 1.013 bar.
Several organizations predict that NH3 will shortly be considered as a promising alternative fuel for maritime transportation [12]. The American Bureau of Shipping (ABS) identified NH3 as a zero-carbon fuel that enters the global market relatively quickly and helps meet the GHG emissions profile, regardless of the fuel source [13]. The DNV-GL published a report about NH3 as a marine fuel, and it is expected that NH3 will potentially play an essential role in de-carbonizing deep-sea vessels. Although NH3 is toxic, with an energy density lower than oil-based fuels, it could be a suitable fuel for internal combustion engines [14]. The Korean Register (KR) published a technical report outlining the safety regulations and resulting design implications for NH3-fueled ships. The report also examines the development status of NH3 fuel cells and internal combustion engines, analyzing critical international requirements such as the IGC and IGF, which will further influence rule development [15]. In addition, the KR issued the guidelines for a ship using NH3 as fuel, describing the class society’s latest safety regulations and inspection standards for NH3-fueled vessels [16].
Many studies have been performed on the marine sector’s NH3-fueled internal combustion (IC) engine and fuel supply system. It is worth noting that the main engine and auxiliary engine manufacturers have already started developing new types of engines combusting NH3 fuel. In 2018, MAN ES announced that the first NH3 unit could be in operation in a short time based on their LPG engine. Furthermore, MAN ES released the principles of the NH3-fueled two-stroke engine and the fuel gas supply system [17] and is aiming for the first delivery of a new NH3-fueled two-stroke engine by 2024 [18]. Furthermore, in 2018, Wartsila signed a memorandum of understanding with Finland’s Lappeenranta University of Technology (LUT) and Nebraska Public Power District (NPPD) to develop a generator engine fueled by NH3 [19]. In 2021, Wartsila and Samsung Heavy Industries (SHI) signed a joint development program agreement to develop NH3-fueled vessels with four-stroke auxiliary engines [20].
Seo et al. [5] proposed two concepts for NH3 fuel storage for an NH3-fueled ammonia carrier and evaluated the concepts in economics. The first concept was to use NH3 in the cargo tank as fuel, and the second was to install an additional independent fuel tank in the vessel. Kim et al. [11] proposed four propulsion systems for a 2500 TEU container feeder ship, all fueled by NH3. They consisted of the main engine, diesel generator, proton-exchange membrane fuel cell (PEMFC), and solid oxide fuel cell (SOFC). Compared to the conventional main engine propulsion system with HFO, the SOFC power system was the most eco-friendly. Trivyza et al. [21] suggested the novel NH3-fueled fuel cell system, and a safety analysis and the preliminary HAZID were performed. In addition, the proposed system’s critical faults and functional failures were identified, and the system’s reaction to the identified hazards was assessed. Kjeld Aado [22] introduced the principles of the NH3-fueled MAN ES two-stroke dual-fuel engines. The NH3 fuel supply system was proposed, similar to the LPG supply system, and the NH3 fuel specifications were described for the two-stroke engine. Duong et al. [23] designed a novel integrated system with SOFCs and a gas turbine (GT) and evaluated it thermodynamically.
A survey of the existing literature and research shows that, despite optimistic demand forecasts and industrial interest in NH3-powered ships, there is a lack of comprehensive studies and analyses on the NH3 fuel supply system for NH3-powered applications. Therefore, the present study proposes a novel design of the NH3 fuel supply system with an onboard NH3 re-liquefaction system in an ocean-going 14,000 TEU container vessel, considering the technical and economic aspects of a deep-sea vessel. This study takes the following approach: First, the target vessel and its appropriate fuel tank are reasonably selected for oceanic conditions, and its potential operation profile is considered. Second, an integrated design of the NH3 fuel supply system and an NH3 re-liquefaction system is generated with the selected fuel tank. Third, the onboard re-liquefaction system suitable for NH3-powered ocean-going vessels is developed and thermodynamically evaluated. Fourth, the economic analysis in this study is performed considering only the annual fuel cost of LNG and NH3 fuel.
The rest of this paper is organized as follows: First, Section 2 clarifies the design basis and presents an NH3 fuel supply system and an onboard full re-liquefaction system. Then, the evaluation methodologies—thermodynamic performance and economic feasibility—are explained. Next, in Section 3, the results for the NH3 fuel supply system and onboard re-liquefaction system are described in detail. Finally, the summary and concluding remarks are presented in Section 4.

2. Materials and Methods

2.1. System Design

This study designs and proposes an NH3 fuel supply system and an onboard re-liquefaction system using NH3 as a refrigerant in an NH3-powered large container ship equipped with an IMO type-A fuel tank. The NH3 IC engine utilizes data from the 2-stroke NH3 engine developed by MAN ES [17]. There is no NH3-fueled vessel currently in operation, and the LNG storage tank is currently being converted to an NH3 storage tank or manufactured as an ammonia-ready LNG-fueled vessel. Recently, there are similarities in use, such as approval for the use of NH3 of existing Mark III LNG systems in GTT [24]; therefore, this study assumed the NH3 fuel tank based on LNG tanks.
LNG is an alternative and bridge fuel for marine transportation due to its environmental and economic advantages. According to the DNV-GL [25], over 200 LNG-fueled ships have been operating, and 403 more have been on order worldwide since 2000. Currently, as LNG storage tanks for ships are developed to store and transport LNG, there are membrane-type LNG tanks and IMO A-, B-, and C-type independent tanks. Therefore, the membrane type, type-A, and type-B tanks, which can efficiently store a large amount of LNG, are being considered for ocean-going vessels. However, the fuel tanks are not designed to have a high internal pressure, so countermeasures for the generated BOG are required. Additionally, the venting of LNG is not allowed, except in emergencies. Therefore, LNG-fueled ships equipped with membrane, IMO type-A, or type-B fuel tanks handle the boil-off gases (BOGs) as fuel for engines or boilers using BOG compressors [26].
Aspen HYSYS V11 is used as a simulation tool for thermodynamic analysis, with extensive data and robust methods for computing physical properties. In addition, the Peng–Robinson equation of state is applied.

2.1.1. Basis of Design

A Target Ship and Main Engine

This study selected a target ship as a 14,000 TEU container ship sailing to the ocean [19]. Table 2 shows the details of the ship.
The propulsion engine used for this target ship is the 12G90ME-C10.5 engine of MAN ES. Table 3 describes the specifications of the engine. In addition, the nominal continuous revolution (NCR) was assumed to be 85% of the specific maximum continuous revolution (SMCR), and the ship’s speed in the NCR was set as 23.5 kts.

Specific Fuel Consumptions

The specific fuel consumptions of the engine load from the 12G90ME-C 10.5 engine are listed in Table 4 below. The specific fuel consumption (SFC) of NH3 and LNG was calculated based on the lower heating value (LHV) of HFO. It is assumed that the LHV of HFO is 42,700 kJ/kg, and the LHV of NH3 and LNG is 18,600 kJ/kg and 50,000 kJ/kg, respectively.

Operation Profile

The fuel’s yearly operational costs greatly depend on the ship’s load profile. Container ships typically sail on well-scheduled voyages because it is vital to maintain their schedules. The expected route of the target vessel is from Asia to Europe, and Table 5 proposes an operation profile to estimate the total operating costs per year [19]. The total number of operating days is 280, and since ocean-going container ships tend to run at high speed, more than 65% of the load of SMCR accounts for approximately 80% of the total operating days.

NH3 Fuel Tank

IMO type-A was selected as the NH3 fuel tank in this study, and the tank capacity was derived as follows: The design pressure of the NH3 fuel tank (IMO type-A) is 0.7 barg, which means the NH3 fuel must be carried in a fully refrigerated condition at or near atmospheric pressure (−33 °C of vapor temperature). With a capacity of 14,861 TEU, CMA CGM Tenere is the first of six Neo-Panamax containers [28]. It is equipped with a 12,000 m3 LNG fuel tank to complete a roundtrip voyage between Asia and Europe. Assuming the filling limit is 98%, the LNG density is 444.3 kg/m3, and the LNG LHV is 49,473 kJ/kg, the total energy of the LNG fuel in the tank is 258,492 GJ. Considering the difference in LHV and density between LNG and NH3 based on the total amount of energy required for the voyage, the mass and volume of NH3 required were estimated as follows (Table 6). The boil-off rate (BOR) of the 21,064 m3 NH3 fuel tank was estimated as 0.04%/day and BOG as 231.6 kg/h [29,30].

2.1.2. Design of NH3 Fuel Supply System

The NH3 fuel supply system (FSS) is designed to supply liquid NH3 fuel at 80 bar and 40 °C to the injection valves of the NH3 engine [17]. Figure 1 shows the schematic diagram of the fuel supply system. The NH3 fuel is pressurized up to 25 bar by the submerged pump and heated up to 40 °C through the LP heater. Next, it is pressurized to 80 bar through the high-pressure (HP) pump to supply fuel to the engine. The heating medium of the LP heater is glycol water (water and ethylene glycol). The seawater coolant cools the re-circulated fuel to 38 °C in the return cooler. Finally, the re-circulated fuel with sealing oil is fed back to the fuel through the HP pump.
The NH3 fuel supply system consists of a fuel supply system, a re-circulation system, a fuel valve train system, a nitrogen system, a ventilation system, and an NH3 capture system [17,22]. This study focuses on the fuel supply system. NH3 is re-circulated through the re-circulation system to cool down the injection equipment. This re-circulated NH3 heats up in the engine during operation [17,22]. In addition, the re-circulated fuel contains some of the sealing oil from the injection equipment, which poses a risk of contaminating the fuel tank. The fuel supply system is designed to address the contamination problem by connecting the re-circulation line between the low-pressure (LP) heater and the HP pump.
Moreover, the amount of re-circulated NH3 varies according to the engine load [31]. When the engine is at a high load, the amount of re-circulated fuel is small compared to the fuel consumption. When the engine is at a low load, the amount of re-circulated fuel is relatively large compared to the fuel consumed. Table 7 shows the re-circulated fuels for fuel consumption. It is assumed that the re-circulated fuel is 20% of the fuel consumption in 100% SMCR, and the re-circulated fuel is 50% of the fuel consumption in 25% SMCR. The remaining re-circulated fuels are derived through the linear regression of these two points.
The following assumptions were made for the simulation of the fuel supply system.
  • The suction temperature and pressure of the submerged fuel pump are −33 °C and 2 bar, respectively.
  • The efficiency of the submerged pump, HP pump, and GW pump is 40%, 75%, and 75%, respectively.
  • The sea water temperature is 32 °C.
  • The temperature and pressure of the re-circulated NH3 are 50 °C and 80 bar.
  • The pressure drops of the heat exchangers are 0.1 bar.
  • The minimum approach temperature of the return cooler is 3 °C.
  • The GW system is designed to maintain the temperature of its GW outlet at 20 °C.
  • The hydraulic static pressure of the GW expansion tank is neglected.

2.1.3. Design of NH3 Re-Liquefaction System

BOG caused by heat penetration from the IMO type-A fuel tank during operation is inevitable. The fuel tank pressure should be effectively managed in terms of the fuel tank’s structural strength and the fuel quality, which should be maintained at low temperatures. The amounts of BOG generated during voyages are evaporation losses, which are a significant factor in the economics of NH3-powered fleets. Consequently, the BOG treatment of ships is critical in securing the fleet economy.
Ocean-going vessels have minimal space compared to onshore plants and have low accessibility due to their high sea movement frequency. Therefore, the criteria for an onboard re-liquefaction system are thermodynamic efficiency, compactness, operational simplicity, and easy maintenance. In addition, excessive BOG after the bunkering operation in a terminal is higher than the nominal BOG from the regular voyage. However, it is not considered a design point because the frequency of the bunkering operation is much lower. Table 8 presents the design assumptions for the process simulation of the re-liquefaction system.
The process flow diagram of the re-liquefaction system is illustrated in Figure 2.
The proposed re-liquefaction system adopts the vapor-compression refrigeration cycle with the NH3 refrigerant. The BOG is routed into the BOG compressor’s suction, so the BOG temperature in the normal voyage is assumed to be −20 °C, which is higher than the fuel storage temperature [32]. The compressed BOG passes through the aftercooler. The BOG with 5 bar and 40 °C passes through the condenser and goes to the tank’s bottom with −15.39 °C and 5 bar. In the refrigeration cycle, NH3 refrigerant is discharged with 15.5 bar and 40 °C by the compressor and the aftercooler. It is delivered to the condenser with −18.53 °C at 2 bar through the J/T valve. The NH3 refrigerant re-liquefies all NH3 BOG in this condenser using the latent heat.
For the re-liquefaction cycle analysis, the following assumptions are drawn:
  • The minimum temperature approach is 3 °C for the condenser.
  • The BOG is composed of 100% NH3.
  • The adiabatic efficiency of the BOG compressors is 75%.
  • The pressure ratios for the 1st- and 2nd-stage BOG compressors are 2.5 and 3.1, respectively.
  • The aftercooler discharge temperature is 40 °C.
  • The pressure drop in the heat exchangers is neglected.

2.2. System Evaluation Methodology

2.2.1. Thermodynamic Performance of NH3 Fuel Supply System

The NH3 fuel supply system is an open system that supplies fuel to the engine at the appropriate temperature and pressure with the amount of fuel required by the engine. The rate of NH3 fuel consumption in the engine varies depending on the ship’s speed. Hence, the power consumption of the fuel supply system varies according to the engine load.
The thermodynamic performance of the fuel supply system can be assessed using the required work per unit fuel mass. The specific electric consumption (SEC) is defined using Equation (1).
SEC = W ˙ t o t a l m ˙ f u e l
The total energy for the fuel gas supply is calculated using Equation (2).
W ˙ t o t a l = W ˙ S u b m e r g e d   + W ˙ H P + W ˙ G W

2.2.2. Thermodynamic Performance of NH3 Re-Liquefaction System

The thermodynamic performance of the re-liquefaction system can be evaluated using the required work per unit mass of liquefied BOG. SEC is defined by Equation (3) to assess the energy required to re-liquefy 1 kg of BOG.
SEC = W ˙ n e t m ˙ R L Q
The total energy for re-liquefaction is calculated using Equation (4).
W ˙ n e t = W ˙ B O G   + W ˙ R e f   1 + W ˙ R e f   2
In thermodynamic physical flow, exergy refers to the maximum useful work delivered to an external user as the stream reaches the dead state. The physical exergy flow is defined by Equation (5).
E ˙ = m ˙ · e = m ˙ · [ ( h 1 + h 0 ) T 0 · ( s 1 s 0 ) ]
where h1 and s1 represent the enthalpy and entropy in State 1, respectively, and subscript 0 indicates the standard environmental conditions (1 atm and 25 °C).
Refrigeration systems refer to the reversible and minimum work required for refrigeration to occur at a particular state. During re-liquefaction, the irreversibility between processes causes exergy loss. To estimate the exergy loss, the physical exergy difference between the inlets and outlets of a component can be used. The exergy loss makes the system less efficient and requires more work than the ideal amount. Table 9 presents the equations for calculating the exergy destruction for equipment.
From this point of view, the exergy efficiency is a thermodynamic performance evaluation of the re-liquefaction system, defined as the minimum work divided by the actual work in Equation (6).
η e x = E ˙ R L Q o u t E ˙ R L Q i n W ˙ n e t
Since no chemical changes are involved in the re-liquefaction system, the total exergy of the system is the physical exergy.

2.2.3. Economic Evaluation

A cost–benefit analysis was performed to assess the economic feasibility of the re-liquefaction system on the NH3-powered container ship. The results from the study are based on determining whether it is economically feasible. The cost–benefit analysis generally includes methods such as the B/C ratio, Net Present Value (NPV), and Internal Rate of Return (IRR). All three methods provide the same conclusion when there is no change in the discount rate. This study uses NPV to evaluate the economic feasibility of adopting the NH3 re-liquefaction system as a BOG treatment.
NPV is the difference between the present value of cash inflows and the present value of cash outflows over a period of time. Therefore, it is estimated by subtracting the cost from the benefit with the discount rate (5%) as present values. When the NPV is larger than 0, the profit of a project starts to be generated. Equation (7) shows the mathematical formulation of NPV [18].
NPV = t = 0 L B t ( 1 + r ) t t = 0 L C t ( 1 + r ) t
For the economic feasibility of the integrated system, the NH3 re-liquefaction system, and fuel supply system, the design of the LNG-powered ship is compared with one of the NH3-powered ships. For these systems, Life Cycle Cost (LCC) is an appropriate analysis that refers to the ownership cost during the lifetime of a system and is widely used for selecting design alternatives. LCC includes all costs involved in the design, construction, operation, and maintenance. It is calculated in Equation (8) [33].
LCC = t = 0 L C t ( 1 + r ) t = t = 0 0 CAPEX t ( 1 + r ) t + t = 0 L OPEX t ( 1 + r ) t
Capital Expenditure (CAPEX) is defined as the initial investment required to construct a plant, consisting of direct and indirect expenses, contingency, and fees, as shown in Equation (9). The direct expenses include the equipment, material, and labor costs required to install the equipment. The indirect project expenses include the freight, insurance, taxes, and overhead costs needed to construct the plant. The contingency is the cost that covers unforeseen circumstances, whereas the fee is related to the contractors. The direct and indirect expenses are collectively referred to as the bare module cost. The contingency and fee are assumed as 15% and 3% of the bare module cost, respectively. Richard Turton’s methodology is used for CAPEX and Operating Expenditure (OPEX) calculations [34].
CAPEX = C D + C I D + C C F = 1.18 · C B M
OPEX is the sum of maintenance, fuel costs, labor, and carbon expenses in Equation (10).
OPEX = C M + C F + C L + C C O 2
However, although NH3 fuel for ships is very promising and many research activities are in progress, no actual project has yet been launched. Due to this uncertainty, OPEX considers only the fuel cost, which is usually the most dominant factor in LCC, and CAPEX considers only the purchased equipment cost for fuel supply systems, including fuel tanks. The amount of carbon dioxide emitted when using LNG fuel uses the carbon factor, Cf, which is suggested by the IMO. Equation (11) shows the carbon factor of LNG [35]. Sensitivity analyses for NH3 prices and carbon dioxide taxes are also performed.
C f = 2.75 ( t o n · C O 2 / t o n · L N G )
The carbon emission costs are added to the NPV and LCC analysis to demonstrate how they would affect the economic viability. They may be estimated as shown in Equation (12).
C C O 2 = M C O 2 · R T A X
For the economic analysis, the following assumptions are drawn:
The power for the operation of the re-liquefaction system is generated by the NH3 internal combustion generator engine with 50% thermal efficiency.
  • The NH3 fuel consumption for the re-liquefaction system is neglected.
  • The LNG price is USD 10 per mmbtu, the price before the Ukraine–Russia conflict in 2022.
  • The amount of BOG is constant, 231.6 kg/h, regardless of the ship’s operation and the tank’s liquid level.
  • The lifetime of the target ship is 20 years.
  • The re-liquefaction system runs on all operation days, 280 days.
  • The discount rate is 5% [36].

3. Results and Discussion

3.1. Thermodynamic Performance of the NH3 Fuel Supply System

Figure 3 shows the FSS power consumption for the engine loads. The power consumption linearly increases as the engine load increases, ranging from 56.4 kW at 25% of SMRC to 157.5 kW at 100% of SMCR. In addition, the SECs tend to decrease as the engine load increases.
The power consumption of each component is listed in Table 10. The re-circulated NH3 fuel contaminated with the sealing oil is routed into the suction of the HP pump. This design can address the potential risk of sealing oil freezing in the fuel tank or other tanks, as well as improve the overall power consumption of the NH3 fuel supply system. The glycol water (GW) system was designed to keep the outlet temperature of the LP heater 20 °C without controlling the system resistance and mass flow. Therefore, the GW pump has a constant power consumption for the engine loads. The inlet GW temperature of the LP heater ranges from 40 °C to 60 °C, which is acceptable.

3.2. Thermodynamic Performance of NH3 Re-Liquefaction System

The energy and exergy efficiency were analyzed to evaluate the thermodynamic performance of the re-liquefaction systems. Table 11 shows the SEC and exergy efficiency of the proposed NH3 re-liquefaction system.
The amount of nominal BOG is selected as the design point of the re-liquefaction plant. In Figure 2, Stream 14, the suction temperature of the BOG compressor was assumed to be 1.4 bar and −20 °C due to the heat ingress through the pipelines between the top of the fuel tank and the suction of the BOG compressor. The BOG is compressed to 5 bar and fully re-liquefied at −15.4 °C. The re-liquefied NH3 is transmitted to the bottom of the fuel tank. The re-liquefied 5 bar-NH3 is expected to be more sub-cooled in the bottom of the fuel tank due to the effect of the hydrostatic pressure of the tank.
In the refrigeration loop, the mass flow rate of the NH3 refrigerant is 315 kg/h, and it is compressed to 15.5 bar with −40 °C. Next, it flows via the J/T valve. Then, the pressure and temperature are 2 bar and −18.5 °C, respectively. The latent heat of the refrigerant removes the heat, 93.4 kW, of the compressed BOG.
The exergy destruction of each component is essential to exergy analyses and controls the total exergy destruction. Figure 4 and Figure 5 show the exergy destruction and the percentage of the total for each piece of equipment. As expected, heat exchangers are the main components of exergy loss. They account for 60% of the total exergy destruction. The reason for this high exergy loss is the irreversibility of the heat transfer process. The compressors are the second most important factor from the perspective of the exergy loss. They are responsible for 30.1%, and the losses are caused by mechanical irreversibility. The heat exchangers and compressors are the main exergy-destructive components described in Figure 4 and Figure 5. Therefore, it is suggested that more attention should be paid to improving the performance of the heat exchangers and compressors to enhance the exergy efficiency of the re-liquefaction system.

3.3. Economic Evaluation

3.3.1. NH3 Re-Liquefaction System

For the economic feasibility of the re-liquefaction system, NPV was performed for the variable NH3 prices and the variable prices of the re-liquefaction system. The profit was estimated by subtracting the re-liquefaction system’s power generation costs from the re-liquefied NH3 costs. Table 12 shows the operation cost for the re-liquefaction system.
The amount of annual re-liquefied NH3 is 1556 tons per year. The cost of the re-liquefied NH3 varies according to the NH3 price. Therefore, Table 13 describes the NPV analysis for ten years in the case that the price of the re-liquefaction system is USD 1 million. When the NH3 is USD 250/ton and USD 500/ton, the profit is generated in the fourth and second years, respectively. If the NH3 is more than USD 500/ton, the profits are generated from the first year. Consequently, the re-liquefaction system is recommended when the price of the re-liquefaction system is USD 1 million.
Figure 6 shows the case when the payback for the prices of the re-liquefaction system and the NH3 fuel becomes positive. If the payback must start from the third year, it is always satisfied when the price of the re-liquefaction system is USD 0.5 million. If it is USD 1 million and USD 1.5 million and the price of NH3 is more than USD 500/ton, the payback is generated within the second and third years. If it is USD 2 million, it creates positive profits within the third year when the NH3 price is more than USD 750/ton. Therefore, it is very reasonable to introduce the re-liquefaction system if it is supplied for less than USD 1 million.
Figure 7 describes NPVs for the fuel and re-liquefaction system price. The NPV is positive when the re-liquefaction system price is less than USD 1.5 million and negative when the re-liquefaction system price is USD 2 million and the ammonia price is USD 250/ton.

3.3.2. LCC for NH3 and LNG

The annual fuel consumptions for HFO, LNG, and NH3 on the engine loads are indicated in Figure 8. The NH3 fuel has the lowest LHV, so the highest flow rate is required. In Table 14, the total annual fuel consumptions for each fuel are described as the NH3 fuel consumption of 106,953 tons, the HFO fuel consumption of 46,588 tons, and the LNG fuel consumption of 39,786 tons. These annual fuel consumptions were derived by multiplying the operation days per engine load in Table 5 by the fuel cost per engine load in Figure 8.
NH3 prices have changed rapidly since the Ukraine–Russia conflict, and to reflect this, sensitivity to changes in NH3 prices from USD 250/ton to USD 1500/ton was analyzed. Figure 9 shows the annual fuel cost of the target container ship according to the NH3 prices. If the price of NH3 is USD 250/ton, USD 26.42 million/year is spent on fuel, and in the case that it reaches USD 1500/ton, USD 16.43 million/year is spent on fuel.
The target ship’s NH3 and LNG fuel’s economic feasibility were investigated, considering CO2 emissions. Since ammonia fuel does not emit carbon dioxide, CO2 emissions are considered only for LNG fuel. The annual total CO2 from the LNG fuel emitted 109,413 tons in the target ship. The LNG price is fixed at the prices before the conflict between Ukraine and Russia. This is because LNG’s solid and stable supply chain will likely become stable after the Ukraine–Russia conflict.
Figure 10 shows a comparative analysis of the annual LNG and NH3 fuel costs, including the carbon tax. In particular, the annual LNG fuel cost rises due to the reflection of the carbon tax. When the carbon tax is USD 50/ton, the LNG fuel is always more economical than the NH3 fuel. If the carbon tax rises to USD 200/ton, the annual LNG fuel cost is estimated to be USD 40.8 million. Given that the annual fuel cost is USD 53.5 million in the case that the price of NH3 is USD 500/ton, the economic feasibility of the NH3 fuel is considered meaningful when the NH3 fuel price falls below USD 400/ton.
CAPEX is listed in Table 15. The price of a 15,000 TEU LNG-fueled container ship is approximately USD 142 million [37], and the supply cost of LNG fuel gas supply system (FGSS), including the fuel tank, is 10–25% of the ship’s price [38]. The LNG FGSS cost of the fuel tank was calculated to be USD 21.3 million, and the NH3 FSS cost was estimated at USD 10.65 million, half of the LNG FGSS cost.
In Figure 11, the results of the LCC for the carbon tax and the NH3 price are illustrated. It can be seen that there is little effect of CAPEX in the LCC analysis. If the NH3 price is more than USD 500/ton, it is not economical compared to LNG. When the NH3 price is USD 250/ton, and the carbon tax is over USD 80/ton, the NH3 system is more economically feasible than the LNG one.

4. Conclusions

This study proposed and economically evaluated the NH3 fuel supply system and the re-liquefaction system for the 14,000 TEU ocean-going container ship. To handle the BOG, the re-liquefaction system was adopted with the vapor compression refrigeration cycle of the NH3 refrigerant. The re-liquefaction system was assessed using thermodynamic performance analysis that estimated the energy and exergy efficiency and the economic feasibility that estimated the NPV. The exergy efficiency and SEC were 34.71% and 0.224 kWh/kg, respectively. In addition, the exergy destruction of each piece of equipment was reviewed. It was found that the exergy destruction of the heat exchangers and compressors accounted for 60% and 30% of the total exergy destruction, respectively. The NPV analysis revealed that if the NH3 price drops to USD 250/ton, the USD 100 million-re-liquefaction system can make a profit in four years. Additionally, if the NH3 price is USD 500/ton, the USD 200 million-re-liquefaction system can make a profit in four years. Considering that the cost of NH3 falls to USD 250–300/ton, a re-liquefaction system cost between USD 0.5 and USD 1 million is reasonable.
The proposed FSS was designed to feed the liquid NH3 at 80 bar and 40 °C and to re-feed the re-circulated NH3 fuel with the sealing oil to the engine. The SEC of the FSS was investigated from 25% SMCR to 100% SMCR. The SEC ranged from 0.009 kWh/kg at 25% SMRC to 0.0063 kW at 100% SMCR.
Finally, the LCC and annual fuel costs for the NH3 and LNG fuels were assessed. The carbon tax was included in the LNG fuel cost in this analysis. The annual LNG fuel cost for USD 50/ton, USD 80/ton, USD 100/ton, USD 150/ton, and USD 200/ton of the carbon tax was USD 24.4 million, USD 27.7 million, USD 29.9 million, USD 35.4 million, and USD 40.8 million, respectively. The average European carbon tax is USD 50/ton, and Switzerland and Sweden impose a carbon tax of USD 130/ton [39]. When the carbon tax is USD 50/ton, LNG fuel is always more economical than NH3 fuel. If the carbon tax soars to USD 200/ton, the annual fuel cost of LNG will rise to USD 40.8 million. Given that the annual fuel cost is USD 53.5 million if the price of NH3 is USD 500/ton, the economic feasibility of the NH3 fuel is considered to be meaningful when the NH3 fuel price falls below USD 400/ton. In addition, according to the results of the LCC analysis, the NH3 fuel is economically feasible in the case that the carbon tax is more than USD 80/ton, and the NH3 price is around USD 250/ton. According to reports, the future market price of NH3 by 2050 is expected to be USD 250–400/ton [9,15,40,41,42]. This shows that the results of this study are significant.
Based on this study, NH3 fuel is currently unattractive to economics. However, NH3 is still estimated as a good candidate for future marine fuel. Therefore, the use of NH3 as a marine fuel will increase as environmental regulations tighten. It is hoped that the results of this study will be an adequate reference for the research and development of NH3-fueled ships and will significantly contribute to their commercialization. Although this study rationally derived design parameters as much as possible, it contains a certain level of uncertainties in economic and technical analyses. Many additional studies are needed to address these uncertainties. In addition, the development and early commercialization of NH3 IC engines, NH3 equipment, FSS, etc., will address these uncertainties. The authors will conduct further research by way of a rigorous LCC analysis and an optimal integrated system of an FSS and BOG management system based on the risk of NH3.

Author Contributions

Conceptualization, J.L.; methodology, J.L.; software, J.L.; validation, J.L., Y.C. and J.C.; formal analysis, J.L.; investigation, J.L.; resources, J.L.; data curation, J.L.; writing—original draft preparation, J.L.; writing—review and editing, Y.C. and J.C.; visualization, J.L.; supervision, J.C.; project administration, J.C.; funding acquisition, J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) (Project No. 20223030030100, Development of web-based guidance platform technology for evaluation & design of fuel cell ships).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

W ˙ i n p u t Work input
W ˙ n e t Total work required to re-liquefy BOG
W ˙ C o m p   Work input for the BOG compressor
W ˙ R e f   1   Work input for no. 1 refrigerant compressor
W ˙ R e f   2   Work input for no. 2 refrigerant compressor
W ˙ t o t a l   Total work required to supply the NH3 fuel
W ˙ S u b m e r g e d Work input for the submerged pump
W ˙ H P   Work input for the HP pump
W ˙ G W   Work input for the GW pump
m ˙ Mass flowrate
m ˙ R L Q Mass flowrate of re-liquified NH3
m ˙ f u e l Mass flowrate of the NH3 fuel
η e x Exergy efficiency
e Specific exergy
E ˙ Physical flow exergy
E ˙ C o l d i n Physical cold flow-in exergy of heat exchanger
E ˙ C o l d o u t Physical cold flow-out exergy of heat exchanger
E ˙ H o t i n Physical hot flow-in exergy of heat exchanger
E ˙ C o l d o u t Physical hot flow-out exergy of heat exchanger
E ˙ R L Q i n Physical flow exergy of BOG of NH3
E ˙ R L Q o u t Physical flow exergy of liquified NH3
B t Benefit at t period in US dollars
C t Cost at t period in US dollars
C D Direct cost for CAPEX
C I D Indirect cost for CAPEX
C C F Contingency and fee for CAPEX
C M Cost of maintenance
C F Cost of fuel consumption
C C O 2 Cost of carbon dioxide
C L Cost of labor
L Time span in years
M C O 2 Mass of carbon dioxide
R T A X Carbon tax
r Discount rate in percent
t Years

Abbreviations

ABSAmerican Bureau of Shipping
B/CBenefit–cost
BOGBoil-off Gas
BORBoil-off Rate
CAPEXCapital Expenditure
CIFCost, Insurance, and Freight
CO2Carbon Dioxide
DFDual Fuel
DNV-GLDet Norske Veritas and Germanischer Lloyd
EEDIEnergy Efficiency Design Index
EEOIEnergy Efficiency Operations Index
FGSSFuel Gas Supply System
FGFuel Gas Free on Board
FSSFuel Supply System
FVTFuel Valves Train
GHGGreen House Gas
GTGas Turbine
GWGlycol Water
HFOHeavy Fuel Oil
HPHigh Pressure
ICInternal Combustion
IEAInternational Energy Agency
IMFInternational Monetary Fund
IMOInternational Marine Organization
IRRInternal Rate of Return
ITFInternational Transport Forum
KRKorean Register
LCCLifecycle Cost
LH2Liquid Hydrogen
LHVLower Heating Value
LPLow pressure
LUTLappeenranta University of Technology
(S)MCR(Specific) Maximum Continuous Revolution
MEPCMarine Environment Protection Committee
NCRNominal Continuous Revolution
NPVNet Present Value
NPPDNebraska Public Power District
OPEXOperating Expenditure
PEMFCProton-Exchange Membrane Fuel Cell
RVTReturn Valves Train
SECSpecific Electric Consumption
SEEMPShip Energy Efficiency Management Plan
SFCSpecific Fuel Consumption
SFOCSpecific Fuel Oil Consumption
SHISamsung Heavy Industries
SOFCSolid Oxide Fuel Cell

References

  1. Sultanbekov, R.; Islamov, S.; Mardashov, D.; Beloglazov, I.; Hemmingsen, T. Research of the Influence of Marine Residual Fuel Composition on Sedimentation Due to Incompatibility. J. Mar. Sci. Eng. 2021, 9, 1067. [Google Scholar] [CrossRef]
  2. Smyshlyaeva, K.I.; Rudko, V.A.; Povarov, V.G.; Shaidulina, A.A.; Efimov, I.; Gabdulkhakov, R.R.; Pyagay, I.N.; Speight, J.G. Influence of Asphaltenes on the Low-Sulphur Residual Marine Fuels’ Stability. J. Mar. Sci. Eng. 2021, 9, 1235. [Google Scholar] [CrossRef]
  3. Wu, P.C.; Lin, C.Y. Strategies for the Low Sulfur Policy of Imo—An Example of a Container Vessel Sailing through a European Route. J. Mar. Sci. Eng. 2021, 9, 1383. [Google Scholar] [CrossRef]
  4. Cheliotis, M.; Boulougouris, E.; Trivyza, N.L.; Theotokatos, G.; Livanos, G.; Mantalos, G.; Stubos, A.; Stamatakis, E.; Venetsanos, A. Review on the Safe Use of Ammonia Fuel Cells in the Maritime Industry. Energies 2021, 14, 3023. [Google Scholar] [CrossRef]
  5. Seo, Y.; Han, S. Economic Evaluation of an Ammonia-Fueled Ammonia Carrier Depending on Methods of Ammonia Fuel Storage. Energies 2021, 14, 8326. [Google Scholar] [CrossRef]
  6. Bakhtov, A. Alternative Fuels for Shipping in the Baltic Sea Region; Helsinki Commission(HELCOM): Helsinki, Finland, 2019. [Google Scholar]
  7. ITF/OECD. Decarbonising Maritime Transport. Pathways to Zero-Carbon Shipping by 2035; OECD: Paris, France, 2018. [Google Scholar]
  8. Lewis, J. Fuels Without Carbon-Prospects and the Pathway Forward for Zero-Carbon Hydrogen and Ammonia Fuel; Clean Air Task Force: Boston, MA, USA, 2018. [Google Scholar]
  9. IEA. The Future of Hydrogen-Seizing Today’s Opportunities; International Energy Agency(IEA): Paris, France, 2019. [Google Scholar]
  10. van Biert, L.; Godjevac, M.; Visser, K.; Aravind, P.V. A Review of Fuel Cell Systems for Maritime Applications. J. Power Sources 2016, 327, 345–364. [Google Scholar] [CrossRef] [Green Version]
  11. Kim, K.; Roh, G.; Kim, W.; Chun, K. A Preliminary Study on an Alternative Ship Propulsion System Fueled by Ammonia: Environmental and Economic Assessments. J. Mar. Sci. Eng. 2020, 8, 183. [Google Scholar] [CrossRef] [Green Version]
  12. Zincir, B. A Short Review of Ammonia as an Alternative Marine Fuel for Decarbonised Maritime Transportation. In Proceedings of the International Conference on Energy, Environment and Storage of Energy (ICEESEN2020), Kayseri, Turkey, 19–21 November 2020; pp. 19–21. [Google Scholar]
  13. The American Bureau of Shipping(ABS). Ammonia As Marine Fuel; ABS: Houston, TX, USA, 2020. [Google Scholar]
  14. Brinks, H.; Hektor, E.A. Ammonia as a Marine Fuel; DNV-GL: Hovik, Nrway, 2020. [Google Scholar]
  15. Korean Register(KR). Report on Ammonia-Fueled Ships; KR: Busan, Korea, 2021. [Google Scholar]
  16. Korean Register(KR). Guidelines for Ships Using Ammonia as Fuels; KR: Busan, Korea, 2021. [Google Scholar]
  17. MAN ES. MAN B&W Two-Stroke Engine Operating on Ammonia; MAN ES: Copenhagen SV, Denmark, 2019. [Google Scholar]
  18. MAN ES. MAN ES Targets 2024 for Delivery of First Ammonia Engine. Available online: https://www.motorship.com/man-es-targets-2024-for-delivery-of-first-ammonia-engine/1258700.article (accessed on 3 September 2022).
  19. MAN ES. Propulsion of 14,000 Teu Container Vessels; MAN ES: Copenhagen SV, Denmark, 2020. [Google Scholar]
  20. WARTSILA KOREA. Wärtsilä and SHI Agree to Collaborate on Ammonia Fuelled Engines for Future Newbuilds. Available online: https://www.wartsila.com/kor/media/news/22-09-2021-wärtsilä-and-shi-agree-to-collaborate-on-ammonia-fuelled-engines-for-future-newbuilds (accessed on 3 September 2022).
  21. Trivyza, N.L.; Cheliotis, M.; Boulougouris, E.; Theotokatos, G. Safety and Reliability Analysis of an Ammonia-Powered Fuel-Cell System. Safety 2021, 7, 80. [Google Scholar] [CrossRef]
  22. Aabo, K. Ammonia-Fuelled MAN B&W 2-Stroke Dual-Fuel Engines. Mar. Eng. 2020, 55, 737–744. [Google Scholar] [CrossRef]
  23. Duong, P.A.; Ryu, B.; Jung, J.; Kang, H. Thermal Evaluation of a Novel Integrated System Based on Solid Oxide Fuel Cells and Combined Heat and Power Production Using Ammonia as Fuel. Appl. Sci. 2022, 12, 6287. [Google Scholar] [CrossRef]
  24. SPLASH. Existing GTT LNG Containment System Can Be Used to Store Ammonia. Available online: https://splash247.com/existing-gtt-lng-containment-system-can-be-used-to-store-ammonia/ (accessed on 5 May 2022).
  25. SAFETY4SEA. 2021 Was a Record Year for LNG Fueled Ships, Says DNV. Available online: https://safety4sea.com/2021-was-a-record-year-for-lng-fueled-ships-says-dnv/ (accessed on 4 September 2022).
  26. Kim, T.W.; Suh, Y.S.; Jang, K.B.; Chun, M.S.; Lee, K.D.; Cha, K.H. A Study and Design on Tank Container for Fuel Tank of LNG Fueled Ship. J. Soc. Nav. Archit. Korea 2012, 49, 504–511. [Google Scholar] [CrossRef]
  27. MAN ES. CEAS Engine Data Report; MAN ES: Copenhagen SV, Denmark, 2022. [Google Scholar]
  28. Riviera. World’s First LNG-Fuelled ULCS Joins CMA CGM Fleet. Available online: https://www.rivieramm.com/news-content-hub/news-content-hub/worldrsquos-first-lng-fuelled-vlcs-joins-cma-cgm-fleet-60937 (accessed on 4 September 2022).
  29. Prem Baboo Ammonia Storage Tank Pre-Commissioning. Int. J. Eng. Res. 2019, 8, 612–624. [CrossRef]
  30. Blanchard, J. Storage of Liquefied Anhydrous Ammonia; CB&I: Houston, TX, USA, 2007. [Google Scholar]
  31. MAN ES. MAN B&W ME-LGIP Dual-Fuel Engines; MAN ES: Copenhagen SV, Denmark, 2018. [Google Scholar]
  32. Lee, H.; Shao, Y.; Lee, S.; Roh, G.; Chun, K.; Kang, H. Analysis and Assessment of Partial Re-Liquefaction System for Liquefied Hydrogen Tankers Using Liquefied Natural Gas (LNG)and H2 Hybrid Propulsion. Int. J. Hydrog. Energy 2019, 44, 15056–15071. [Google Scholar] [CrossRef]
  33. Adachi, M.; Kosaka, H.; Fukuda, T.; Ohashi, S.; Harumi, K. Economic Analysis of Trans-Ocean LNG-Fueled Container Ship. J. Mar. Sci. Technol. 2014, 19, 470–478. [Google Scholar] [CrossRef]
  34. Turton, R.; Bailie, R.C.; Whiting, W.B.; Shaeiwitz, J.A. Analysis, Systhesis, and Design of Chemical Processes, 4th ed.; Prentice Hall: Upper Saddle River, NJ, USA, 2012. [Google Scholar]
  35. IMO MEPC. RESOLUTION MEPC.281(70) (Adopted on 28 October 2016) AMENDMENTS; IMO: London, UK, 2016. [Google Scholar]
  36. Thoft-Christensen, P. Infrastructures and Life-Cycle Cost-Benefit Analysis. Struct. Infrastruct. Eng. 2012, 8, 507–513. [Google Scholar] [CrossRef] [Green Version]
  37. PULSE. Samsung Heavy Industries Grabs $710 Mn LNG-Fueled Container Ship Order. Available online: https://pulsenews.co.kr/view.php?sc=30800021&year=2021&no=151388 (accessed on 4 September 2022).
  38. Yoo, B.Y. Economic Assessment of Liquefied Natural Gas (LNG) as a Marine Fuel for CO2 Carriers Compared to Marine Gas Oil (MGO). Energy 2017, 121, 772–780. [Google Scholar] [CrossRef]
  39. Bray, S. Carbon Taxes in Europe. Available online: https://taxfoundation.org/carbon-taxes-in-europe-2022/ (accessed on 3 September 2022).
  40. Laval, A.; Hanfia; Topsøe, H.; Vestas; Gamesa, S. Ammonfuel—An Industrial View of Ammonia as a Marine Fuel; Alfa Laval: Lund, Sweden, 2020. [Google Scholar]
  41. Blanco, H.; Jinks, B.; Bianco, E.; Sezer, U. Innovation Outlook Renewable Ammonia in Partnership with about Irena Valuable Review and Feedback Was also Provided by Irena Colleagues Renewable Ammonia 3 Contents. 2022. Available online: https://www.irena.org/publications/2022/May/Innovation-Outlook-Renewable-Ammonia (accessed on 4 September 2022).
  42. Bunro Schiozawa. The Cost of CO2-Free Ammonia. Available online: https://www.ammoniaenergy.org/articles/the-cost-of-co2-free-ammonia/ (accessed on 4 September 2022).
Figure 1. Process flow diagram for NH3 fuel supply system.
Figure 1. Process flow diagram for NH3 fuel supply system.
Jmse 10 01500 g001
Figure 2. Process flow diagram for NH3 re-liquefaction system with FSS.
Figure 2. Process flow diagram for NH3 re-liquefaction system with FSS.
Jmse 10 01500 g002
Figure 3. FSS power consumption for % SMCR.
Figure 3. FSS power consumption for % SMCR.
Jmse 10 01500 g003
Figure 4. Exergy destruction for each equipment.
Figure 4. Exergy destruction for each equipment.
Jmse 10 01500 g004
Figure 5. Contribution of each piece of equipment to total exergy destruction.
Figure 5. Contribution of each piece of equipment to total exergy destruction.
Jmse 10 01500 g005
Figure 6. Payback for the NH3 and re-liquefaction prices.
Figure 6. Payback for the NH3 and re-liquefaction prices.
Jmse 10 01500 g006
Figure 7. NPV for the NH3 and re-liquefaction price.
Figure 7. NPV for the NH3 and re-liquefaction price.
Jmse 10 01500 g007
Figure 8. Fuel consumption for the engine loads for HFO, LNG, and NH3.
Figure 8. Fuel consumption for the engine loads for HFO, LNG, and NH3.
Jmse 10 01500 g008
Figure 9. Fuel cost for the NH3 price.
Figure 9. Fuel cost for the NH3 price.
Jmse 10 01500 g009
Figure 10. LNG and NH3 fuel cost for the carbon tax and NH3 price.
Figure 10. LNG and NH3 fuel cost for the carbon tax and NH3 price.
Jmse 10 01500 g010
Figure 11. LCC of LNG and NH3 for the carbon tax and NH3 price.
Figure 11. LCC of LNG and NH3 for the carbon tax and NH3 price.
Jmse 10 01500 g011
Table 1. Fuel properties of HFO, H2, and NH3 [11].
Table 1. Fuel properties of HFO, H2, and NH3 [11].
PropertyUnitHFOCompressed H2 (350 bar)LH2Liquid NH3
LHVMJ/kg40.212012018.6
Volumetric energy densityMJ/m339,564–42,0365040850014,100
Min. auto-ignition temperature°C250500–577500–577650–657
Boiling temperature at 1 bar°CN/AN/A−253−33.4
Condensation pressure at 25 °CbarN/AN/AN/A9.90
Hydrogen content% massN/A10010017.8
Table 2. Details of the target vessel [19].
Table 2. Details of the target vessel [19].
SpecificationUnitValue
Deadweight, maxDWT150,000
Scantling draughtm15.8
Design draughtm14.5
Length overallm368
Length between ppm352
BreadthM51
Sea margin%15
Engine margin%15
Light running margin%5
Design ship speedkn21.5/23.5
Type of propellermFPP
No. of propeller bladesEA5/6
Propeller diameterm9.6–10
Table 3. Specification of the main engine [27].
Table 3. Specification of the main engine [27].
Main EngineSMCR, kWNCR, kWSpeed at NCR, ktsSFOC at NCR, kWh
12G90ME-C10.566,35356,40023.5158.9
Table 4. Specific consumption of fuels [27].
Table 4. Specific consumption of fuels [27].
LoadPowerSFC, g/kWh
% SMCRkWHFOLNGNH3
10066,353165.1141.0379.0
9563,035162.9139.1374.0
9059,718160.9137.4369.4
8556,400158.9135.7364.8
8053,082158.1135.0363.0
7549,765157.7134.7362.0
7046,447155.2132.5356.3
6543,129153.5131.1352.4
6039,812154.2131.7354.0
5536,494155.2132.5356.3
5033,177156.2133.4358.6
4529,859157.6134.6361.8
4026,541159.1135.9365.2
3523,224160.6137.2368.7
3019,906162.2138.5372.4
2516,588163.8139.9376.0
Table 5. Operation profile [19].
Table 5. Operation profile [19].
Load, % SMCRPower, kWOperation Days
10066,35314
8556,40084
6543,129126
5033,17714
3523,22414
2516,58828
Total-280
Table 6. NH3 fuel tank sizing [28].
Table 6. NH3 fuel tank sizing [28].
PropertyValueRemark
Required energy, GJ258,49212,000 m3 LNG fuel tank
NH3 LHV, kJ/kg18,604
NH3 density, kg/m3673.1Liquid saturation at 1.013 bar
Required NH3 mass, kg13,894,267
Required NH3 volume, m320,643
NH3 fuel tank size, m321,06498% filling limit
Table 7. Fuel consumption and re-circulated fuel.
Table 7. Fuel consumption and re-circulated fuel.
% SMCRFuel Consumption, kg/hRe-circulated Fuel, kg/h
10025,149.11258
9523,573.21413
9022,058.51562
8520,574.01708
8019,266.11837
7518,016.51960
7016,548.72105
6515,198.22237
6014,093.32346
5513,002.52453
5011,896.92562
4510,803.12670
409694.02779
358562.42890
307412.23004
256237.73119
Table 8. Design assumptions of BOG re-liquefaction system.
Table 8. Design assumptions of BOG re-liquefaction system.
ItemUnitValue
Boil-off rate%/day0.4 [29,30]
BOGkg/h231.6 [32]
BOG feed temperature°C−20
Suction pressure of BOG compressorbar1.4
Composition of BOG%NH3 100
Table 9. Equation of exergy destruction.
Table 9. Equation of exergy destruction.
EquipmentExergy Destruction, kW
Compressor E ˙ i n + W ˙ i n p u t E ˙ o u t
Heat exchanger E ˙ C o l d i n + E ˙ H o t i n E ˙ C o l d o u t E ˙ H o t o u t
Separator, Valve E ˙ i n E ˙ o u t
Table 10. Power consumption for each unit.
Table 10. Power consumption for each unit.
% SMCRMass Flow, kg/hGW Pump, kWHP Pump, kWSubmerged Pump, kWTotal Power, kWSEC, kWh/kg
10025,149.18.999.349.3157.50.00626
9523,573.28.994.046.2149.10.00633
9022,058.58.988.943.2141.00.00639
8520,5748.983.840.3133.10.00647
8019,266.18.979.437.8126.10.00654
7518,016.58.975.235.3119.40.00663
7016,548.78.970.232.4111.50.00674
6515,198.28.965.629.8104.30.00686
6014,093.38.961.927.698.40.00698
5513,002.58.958.225.592.60.00712
5011,896.98.954.423.386.70.00728
4510,803.18.950.721.280.80.00748
4096948.947.019.074.90.00772
358562.48.943.116.868.80.00804
307412.28.939.214.562.70.00846
256237.78.935.312.256.40.00904
Table 11. Re-liquefaction performances.
Table 11. Re-liquefaction performances.
Performance IndexUnitValue
W ˙ n e t kW51.9
m ˙ R L Q Kg/h231.6
SECkWh/kg0.224
Exergy efficiency%34.71
Table 12. Cost for the re-liquefaction system’s operation.
Table 12. Cost for the re-liquefaction system’s operation.
NH3 Price, USD/tonPower Generation Cost, USD/year
25033,750
50067,501
750101,251
1000135,001
1250168,752
1500202,502
The power consumption for the re-liquefaction system is 51.9 kW.
Table 13. Annual profit and NPV (the re-liquefaction system: USD 1 million).
Table 13. Annual profit and NPV (the re-liquefaction system: USD 1 million).
YearNPV
NH3
USD 250/ton
NH3
USD 500/ton
NH3
USD 750/ton
NH3
USD 1000/ton
NH3
USD 1250/ton
NH3
USD 1500/ton
1−661,629−323,25915,112353,482691,8531,030,224
2−339,372324,172981,8851,642,5132,303,1422,963,770
3−32,460945,9361,902,6212,870,1623,837,7024,805,243
4259,8381,544,6242,779,5134,039,3515,299,1886,559,026
5538,2162,121,8003,614,6485,152,8646,691,0808,229,296
6803,3382,678,1864,410,0156,213,3538,016,6919,820,029
71,055,8363,213,9335,167,5077,223,3429,279,17811,335,013
81,296,3093,728,9135,888,9288,185,23710,481,54612,777,855
91,525,3324,222,9576,575,9959,101,32711,626,65914,151,991
101,743,4484,696,0187,230,3459,973,79412,717,24215,460,691
NPV6,188,85623,153,27938,566,56954,755,42570,944,28287,133,138
Table 14. Annual fuel consumption for fuels.
Table 14. Annual fuel consumption for fuels.
Annual Operation DaysNH3 Fuel Consumption, tonsHFO Fuel Consumption, tonsLNG Fuel Consumption, tons
280106,95346,58839,786
Table 15. CAPEX of LNG FGSS and NH3 FSS.
Table 15. CAPEX of LNG FGSS and NH3 FSS.
LNG FGSS, Million USDNH3 FSS Price, Million USD
21.310.65
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, J.; Choi, Y.; Choi, J. Techno-Economic Analysis of NH3 Fuel Supply and Onboard Re-Liquefaction System for an NH3-Fueled Ocean-Going Large Container Ship. J. Mar. Sci. Eng. 2022, 10, 1500. https://doi.org/10.3390/jmse10101500

AMA Style

Lee J, Choi Y, Choi J. Techno-Economic Analysis of NH3 Fuel Supply and Onboard Re-Liquefaction System for an NH3-Fueled Ocean-Going Large Container Ship. Journal of Marine Science and Engineering. 2022; 10(10):1500. https://doi.org/10.3390/jmse10101500

Chicago/Turabian Style

Lee, Jinkwang, Younseok Choi, and Jungho Choi. 2022. "Techno-Economic Analysis of NH3 Fuel Supply and Onboard Re-Liquefaction System for an NH3-Fueled Ocean-Going Large Container Ship" Journal of Marine Science and Engineering 10, no. 10: 1500. https://doi.org/10.3390/jmse10101500

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